Entry - *133430 - ESTROGEN RECEPTOR 1; ESR1 - OMIM
 
* 133430

ESTROGEN RECEPTOR 1; ESR1


Alternative titles; symbols

ESTROGEN RECEPTOR; ESR; ER
ESTROGEN RECEPTOR, ALPHA; ESRA


HGNC Approved Gene Symbol: ESR1

Cytogenetic location: 6q25.1-q25.2     Genomic coordinates (GRCh38): 6:151,656,672-152,129,619 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
6q25.1-q25.2 {Migraine, susceptibility to} 157300 AD 3
{Myocardial infarction, susceptibility to} 608446 3
Breast cancer, somatic 114480 3
Estrogen resistance 615363 AR 3

TEXT

Description

The estrogen receptor (ESR) is a ligand-activated transcription factor composed of several domains important for hormone binding, DNA binding, and activation of transcription. Alternative splicing results in several ESR1 mRNA transcripts, which differ primarily in their 5-prime untranslated regions. The translated receptors show less variability (Kos et al., 2001; Flouriot et al., 2000).


Cloning and Expression

Walter et al. (1985) cloned and Greene et al. (1986) sequenced a cDNA for the entire translated portion of the messenger RNA for the estrogen receptor of MCF-7 human breast cancer cells. Expression with production of a functional protein was accomplished in Chinese hamster ovary cells. The 1,785 nucleotides of the cDNA correspond to a polypeptide of 595 amino acids and a molecular weight of 66,200 (about that estimated from other studies of the estrogen receptor). Amino acid sequence comparisons showed considerable similarities between human ESR, human GCR (138040), and the putative v-erbA (190120) oncogene product. Both ESR and GCR exert their effects by binding directly to an intranuclear receptor molecule that is weakly associated with nuclear components in the absence of ligand. Binding of hormone to its receptor results in conversion of the receptor-steroid complex to a form that binds with high affinity to nuclear components. Green et al. (1986) also cloned and sequenced human estrogen receptor cDNA, using the breast cancer cell line MCF-7. They found extensive homology between ESR cDNA and the ERBA oncogene.

Flouriot et al. (2000) cloned a 46-kD isoform of ESR1 that lacks 173 N-terminal amino acids present in the major 66-kD isoform. These amino acids form a ligand-independent transactivation domain. The 46-kD isoform retains the DNA-binding domains and the hormone-binding domain, which includes the functional domain for hormone-inducible transcription activation.

Li et al. (2003) confirmed that the 46-kD endothelial cell protein (ESR46) is an N-terminal truncated product of full-length ESR-alpha (ESR66) that results from alternative splicing. ESR46 is expressed in the plasma membrane, cytosol, and nucleus of resting, estrogen-deprived cells. They found that ESR46 is localized and further dynamically targeted to the plasma membrane in a palmitoylation-dependent manner.


Biochemical Features

Crystal Structure

Shiau et al. (1998) reported the crystal structure of the human ESRA ligand-binding domain (LBD) bound to both the agonist diethylstilbestrol (DES) and a peptide derived from the nuclear receptor (NR) box II region of the coactivator GRIP1 (601993) as well as the crystal structure of the ESRA-LBD bound to the selective antagonist 4-hydroxytamoxifen (OHT). In the DES-LBD-peptide complex, the peptide binds as a short alpha helix to a hydrophobic groove on the surface of the LBD. In the OHT-LBD complex, helix 12 occludes the coactivator recognition groove by mimicking the interactions of the NR box peptide with the LBD.


Gene Structure

Ponglikitmongkol et al. (1988) showed that the human ESR gene is more than 140 kb long. It contains 8 exons, and the position of its introns has been highly conserved, being, for example, remarkably similar to those of one of the chicken thyroid hormone receptor genes.

Kos et al. (2001) reviewed the organization of the ESR1 gene. They described promoters used in the generation of ESR1 transcripts in human and other species and suggested a consistent nomenclature. The possible role of multiple promoters in the differential expression of ESR1 in tissues and during development was also discussed.


Mapping

Walter et al. (1985) determined that the human ESR gene maps to chromosome 6. By in situ hybridization, using a cDNA probe containing the coding sequence for the estrogen receptor, Gosden et al. (1986) assigned the gene to 6q24-q27. Zuppan et al. (1989) reported linkage data between the ESR locus and MYB (189990) and TCP10 (187020) on chromosome 6q. The estimated recombination fractions were 0.19 between ESR and MYB and 0.14 between ESR and TCP10, assuming equal male and female recombination fractions. To localize ESR more precisely, Menasce et al. (1993) isolated a YAC containing the gene and mapped it to 6q25.1 by fluorescence in situ hybridization (FISH) and a new simple method of post-FISH chromosome banding. Using a single interspecific backcross, Justice et al. (1990) demonstrated the genetic location of the Esr gene in relation to other loci on mouse chromosome 10.


Gene Function

Using a construct containing the human estrogen receptor cDNA with a yeast PGK promoter, Metzger et al. (1988) demonstrated that the human estrogen receptor can be expressed in yeast. Initiation of transcription in the system operated in a strictly hormone-dependent manner, documenting an extraordinarily high level of conservation of this activation mechanism.

Issa et al. (1994) reported that CpG island methylation, an epigenetic modification of DNA known to correlate closely with silencing of gene transcription, appears in the ESR gene in a subpopulation of cells that increases as a direct function of age in human colonic mucosa. They found that the same methylation change characterized almost all cells in the 45 colorectal tumors examined, including the earliest stages of tumor formation. ESR gene expression is diminished or absent in colorectal tumors, and introduction of an exogenous ESR gene in cultured colon carcinoma cells was found to result in marked growth suppression. Issa et al. (1994) concluded that methylation-associated inactivation of the ESR gene in aging colorectal mucosa may be one of the earliest events that predispose to sporadic colorectal tumorigenesis.

One mechanism suggested to play a role in the progression of human breast cancer from hormone dependence to independence is the expression or altered expression of mutant and/or variant forms of the estrogen receptor. Murphy et al. (1996) stated that 2 major types of variant ESR mRNA had been reported in human breast biopsy samples so far: truncated transcripts and exon-deleted transcripts. Murphy et al. (1996) provided data on a novel type of abnormal ESR mRNA. They found larger-than-wildtype ESR mRNA RT-PCR products in 9.4% of 212 human breast tumors analyzed. Cloning and sequencing of these larger RT-PCR products showed 3 different types: complete duplication of exon 6 in 7.5%; complete duplication of exons 3 and 4 in 1 tumor; and a 69-bp insertion between exons 5 and 6 in 3 tumors. While it is unknown if these novel ESR-like mRNAs are stably translated in vivo, any resulting protein would be structurally altered, possibly resulting in altered function.

Two isoforms of the human ESR, ESRA and ESR-beta (ESR2; 601663), occur, each with distinct tissue and cell patterns of expression. Additional ESR isoforms, generated by alternative mRNA splicing, have been defined in several tissues and are postulated to play a role in tumorigenesis or in modulating the estrogen response. By RT-PCR and hybridization blotting analysis, Shupnik et al. (1998) examined 71 human pituitary adenomas of varying phenotypes and 6 normal pituitary specimens for ESR mRNA isoforms. All 14 prolactinomas contained ESRA, and 5 of 14 contained ESRB mRNA. In comparison, 6 tumors that expressed prolactin (176760) and growth hormone (GH; 139250) expressed ESRB (4 of 6) more frequently than ESRA (3 of 6). ESRB mRNA was also found more frequently in null cell (8 of 24 ESRA and 14 of 24 ESRB) and gonadotrope (13 of 21 ESRA and 18 of 21 ESRB) tumors. Additionally, ESRB was found in 4 of 6 tumors that contained only GH, although ESRA was not observed in this tumor type. Expression of the 2 ESR isoforms within a tumor type was overlapping, but some tumors contained only 1 isoform. Expression of ESRA mRNA splice variants also varied with cell type. All normal pituitaries contained ESRA deletions of exon 4, 5, and 7, whereas only 2 of 6 samples contained the exon 2 deletion variant. Although the same ESRA mRNA variants were observed among the various tumor types, the proportion of specific splice variants expressed varied. The authors concluded that expression of the ESR isoforms, as well as of the mRNA splice variants, may influence the biologic properties of these tumors and affect their ability to respond to estrogen and antiestrogen therapies.

Vaults are large ribonucleoprotein particles composed largely of MVP (605088). Abbondanza et al. (1998) found that MVP coprecipitated with ER from nuclear extracts of MCF-7 human breast cancer cells and that ER associated with intact vaults. Mutation analysis showed that a central region of ER containing nuclear localization signals was involved in the interaction. A limited amount of ER molecules in the nuclear extract appeared to be associated with MVP. Physiologic concentrations of estradiol increased the amount of MVP present in MCF-7 nuclear extracts and coimmunoprecipitated with ER. The hormone-dependent interaction of vaults with ER was reproduced in vitro.

Lawson et al. (1999) studied ESR1 expression in normal breast tissue biopsies obtained by open surgical and core biopsy of benign lesions of the breast from 74 white Australian women and 92 Japanese women. Among Australians, the median of the percentage of ESR1-positive cells is about 12% for women younger than 50 years and 17% among women 50 years or older. Among Japanese, the overall median is about 9% and does not vary substantially with age. The authors concluded that their results are compatible with the hypothesis that the expression of estrogen receptors in normal breast tissue increases the risk of breast cancer and explains the 5-fold contrast of breast cancer incidence between white and Asian women. The difference in estrogen receptor expression as well as in breast cancer incidence between white and Japanese is more evident among older women.

Primary transcripts of the human ESR and progesterone receptor (PGR; 607311) genes undergo a number of alternative-splicing events that result in a range of variant mRNA isoforms in receptor-positive tissues. Despite in vitro demonstrations of a possible role for some of these isoforms in hormonal sensitivity, the clinical significance of this process is uncertain. By RT-PCR and Southern blot analysis, Balleine et al. (1999) documented the coexpression of variant ESR and PGR transcripts in a series of receptor-positive breast tumors. In 35 ESR-positive tumors, a common profile of variant ESR transcripts was present, with all tumors containing the exon 2-deleted and exon 7-deleted ESR variants, 94% containing the exon 4-deleted ESR variant, and 83% containing the exon 5-deleted ESR variant. In 25 of these cases, which were also PGR positive, the most highly expressed PGR variants, the exon 4-deleted PGR, exon 6-deleted PGR, and delta-4/2PGR, a transcript from which a 126-bp portion of PGR exon 4 was deleted, were detected in more than 90% of the cases. The alternatively spliced ESR isoforms were expressed at higher relative levels than the PGR isoforms, which had mean levels of expression less than 10% that of wildtype PGR. The most abundant isoform was the exon 7-deleted ESR, which was present at levels ranging from 29 to 83% of the wildtype. Balleine et al. (1999) concluded that the common profile of alternatively spliced ESR and PGR transcripts seen in breast tumors precludes its use as a discriminator of hormone responsiveness or other clinical end points. Furthermore, the low level of expression of the majority of variant isoforms called into question their potential for impacting significantly on receptor function.

Using transient transfection assays, Fan et al. (1999) demonstrated that BRCA1 inhibits signaling by the ligand-activated estrogen receptor ESR-alpha through the estrogen-responsive enhancer element and blocks the C-terminal transcriptional activation function AF2 of ESR-alpha. These results suggested that wildtype BRCA1 protein may function, in part, to suppress estrogen-dependent mammary epithelial proliferation by inhibiting ESR-alpha-mediated transcriptional pathways related to cell proliferation, and that loss of this ability may contribute to tumorigenesis.

Chaidarun and Alexander (1998) examined the effects of a human ESR1 isoform on estrogen-mediated gene activation in U2-OS osteosarcoma cells. ESR1-5, an ESR1 variant generated by an alternate splice event that omits exon 5 and alters the reading frame of the resulting mRNA, was coexpressed with normal ESR1 in several estrogen-responsive neoplastic tissues. ESR1-5 encodes the hormone-independent trans-activating function (AF-1), as well as the constitutive receptor dimerization and DNA-binding domains. The ESR1-5 protein is prematurely terminated and lacks the majority of the hormone-binding and activating function-2 (AF-2) domains. When ESR1-5 was cotransfected with normal ESR1, both basal and estrogen-stimulated reporter activation were increased approximately 500% over the levels observed when cells were transfected with ESR1 alone. Electromobility shift/supershift assays using nuclear extracts of U2-OS cells stably transfected with ESR1 and ESR1-5 confirmed the constitutive binding of ESR1-5 and ESR1 protein to the estrogen-response element (ERE) sequence independent of estrogen, and also showed an increase in ESR1-5/ESR1-ERE complexes with estrogen treatment. These data were considered to be consistent with the interactive effects of normal ESR1 and ESR1-5 on transcription from classic ERE gene promoters. Chaidarun and Alexander (1998) concluded that ESR1-5 acts as a dominant-positive receptor that increases both basal and estrogen-stimulated gene transactivation of normal ESR1.

Ye et al. (2000) investigated the expression of ESR1 in prostate cancer cell lines and unexpectedly found a FASN/ESR1 fusion transcript. Using semi-nested RT-PCR analysis of ESR1 and its variants, Ye et al. (2000) found that the N-terminal coding region of FASN containing domain 1 was fused to the C-terminal coding region of the ESR1 ligand binding domain. Nested RT-PCR also detected the fusion transcript in breast, cervical, and bladder cancer cell lines.

Esmaeli et al. (2000) found immunohistochemical evidence that estrogen receptors are present in the meibomian glands of the upper eyelid. Unlike sebaceous glands elsewhere on the skin, the meibomian glands lack androgen receptors. Esmaeli et al. (2000) suggested that these eyelid estrogen receptors may play a role in modulation of the tear film lipid layer. They concluded that estrogen receptor activity may be linked to meibomian gland dysfunction and dry eye syndrome.

ESR1 is downregulated in the presence of its cognate ligand, estradiol, through the ubiquitin proteasome pathway. Lonard et al. (2000) showed that ubiquitin proteasome function is required for ESR1 to serve as a transcriptional activator. Deletion of the last 61 amino acids of ESR1, including residues that form helix 12, abolished ligand-mediated downregulation of the receptor, as did point mutations in the ligand-binding domain that impaired coactivator binding. In addition, the authors found that coactivators also are subject to degradation by the 26S proteasome, but their intrinsic transcriptional activity is not affected. These data provided evidence that protein interactions with ESR1 coactivator binding surfaces are important for ligand-mediated receptor downregulation and suggested that receptor and coactivator turnover contributes to ESR1 transcriptional activity.

Simoncini et al. (2000) showed that the estrogen receptor isoform ESR-alpha binds in a ligand-dependent manner to the p85-alpha (171833) regulatory subunit of the phosphatidylinositol-3-hydroxy kinase (PI3K). Stimulation with estrogen increases ESR-alpha-associated PI3K activity, leading to the activation of protein kinase B/AKT (164730) and endothelial nitric oxide synthase (eNOS; 163729). Recruitment and activation of PI3K by ligand-bound ESR-alpha are independent of gene transcription, do not involve phosphotyrosine adaptor molecules or src homology domains of p85-alpha, and extend to other steroid hormone receptors. Mice treated with estrogen showed increased eNOS activity and decreased vascular leukocyte accumulation after ischemia and reperfusion injury. This vascular protective effect of estrogen was abolished in the presence of PI3K or eNOS inhibitors. Simoncini et al. (2000) concluded that their findings defined a physiologically important nonnuclear estrogen-signaling pathway involving the direct interaction of ESR-alpha with PI3K.

Pelletier and El-Alfy (2000) studied the immunocytochemical localization of ESRA and ESRB in human reproductive tissues. In the ovary, ESRB immunoreactivity was found in nuclei of granulosa cells of growing follicles at all stages from primary to mature follicles, interstitial gland, and germinal epithelium cells. Nuclear staining for ESRA occurred in thecal, interstitial gland, and germinal epithelium cells. In the uterus, strong ESRA immunoreactivity was detected in nuclei of epithelial, stromal, and muscle cells. Similar localization was obtained for ESRB, although the staining was much weaker. In the vagina, only ESRA could be detected; a nuclear reaction was observed in deep layers of the stratified epithelium as well as in stromal and muscle cells. In the mammary gland, both ESR subtypes were observed in epithelial and stromal cells. In the testis, ESRB was detected in the nuclei of Sertoli and Leydig cells, whereas ESRA immunoreactivity was only observed in Leydig cells, with no tubular labeling. In the efferent ducts, only ESRB could be detected, whereas neither ESRB nor ESRA could be found in the epididymis. In the prostate, ESRB nuclear immunolabeling was observed in both basal and secretory cells in alveoli as well as in stromal cells, whereas ESRA could not be detected. The authors concluded that there is a cell-specific localization for each of the ESR subtypes in the majority of the reproductive organs studied.

Chiang et al. (2000) examined the regulation of ESRA and ESRB expression by human chorionic gonadotropin (CG; see 118850) in human granulosa-luteal cells. CG treatment significantly attenuated the ESRA (45%) and ESRB (40%) mRNA levels. The CG-induced decrease in ESRA and ESRB expression was mimicked by 8-bromo-cAMP and forskolin treatment. Next, the effect of gonadotropin-releasing hormone (GNRH; 152760) on estrogen receptor expression was studied. Sixty-eight percent and 60% decreases in ESRA and ESRB mRNA levels, respectively, were observed after treatment with GNRH agonist. Pretreatment of the cells with a protein kinase C (PKC; see 176960) inhibitor completely reversed the GNRH agonist-induced downregulation of ESRA and ESRB expression, suggesting the involvement of PKC in GNRH signal transduction in granulosa-luteal cells. Chiang et al. (2000) observed a differential expression of ESRA and ESRB mRNA in granulosa-luteal cells in vitro. The authors concluded that the demonstration of CG- and GNRH agonist-induced downregulation of ESRA and ESRB gene expression suggests that CG and GNRH may contribute to the control of granulosa-luteal cell function. Furthermore, they inferred that the effects of CG and GNRH on ESRA and ESRB expression in granulosa-luteal cells are mediated in part by activation of PKA (see 176911) and PKC signaling pathways, respectively.

Bord et al. (2001) examined the expression of ESR1 and ESR2 in neonatal human rib bone. ESR1 and ESR2 immunoreactivity was seen in proliferative and prehypertrophic chondrocytes in the growth plate, with lower levels of expression in the late hypertrophic zone. Different patterns of expression of the 2 estrogen receptors were seen in bone. In cortical bone, intense staining for ESR1 was observed in osteoblasts and osteocytes adjacent to the periosteal-forming surface and in osteoclasts on the opposing resorbing surface. In cancellous bone, ESR2 was strongly expressed in both osteoblasts and osteocytes, whereas only low expression of ESR1 was seen in these areas. Nuclear and cytoplasmic staining for ESR2 was apparent in osteoclasts. The authors concluded that these observations demonstrate distinct patterns of expression for the 2 ER subtypes in developing human bone and indicate functions in both the growth plate and mineralized bone. In the latter, ESR1 is predominantly expressed in cortical bone, whereas ESR2 shows higher levels of expression in cancellous bone.

Takeyama et al. (2001) examined the expression and cellular localization of ESR1 and ESR2 in various human fetal tissues by semiquantitative RT-PCR (13 and 20 gestational weeks) and immunohistochemistry (13, 20, and 38 gestational weeks), respectively, to study the possible effects of estrogens on human fetal tissues during development. Relatively high levels of ESR2 expression were detected in various human fetal tissues, whereas those tissues expressing ESR2 had markedly lower levels of ESR1 expression. ESR2 mRNA expression was especially high in the adrenal gland. ESR2-immunoreactive protein was localized to the definitive zone, but not in the fetal zone, of the adrenal cortex. Although low levels of ESR2 mRNA were present in the brain, heart, lung, and kidney, ESR2 immunoreactivity was not detected in these tissues. These results suggested that the effects of estrogens in these tissues are predominantly mediated through ESR2. ESR2 immunoreactivity was detected in Sertoli cells and spermatogonia in the male reproductive tract and in germ cells in the fetal testis and epididymis. In the female reproductive tract, both ESR1 and ESR2 were immunopositive in the epithelium of the oviduct. These results demonstrated the possible sites for estrogenic action in the human fetus and suggested that the effects of estrogen via ESR2 may play important roles in human fetal development, especially in the definitive zone of the adrenal cortex and in the reproductive tissues of the developing fetus.

The therapeutic effectiveness of selective estrogen receptor modulators such as tamoxifen and raloxifene in breast cancer depends on their antiestrogenic activity. In the uterus, however, tamoxifen is estrogenic. Shang and Brown (2002) showed that both tamoxifen and raloxifene induce the recruitment of corepressors to target gene promoters in mammary cells. In endometrial cells, tamoxifen, but not raloxifene, acts like estrogen by stimulating the recruitment of coactivators to a subset of genes. The estrogen-like activity of tamoxifen in the uterus requires a high level of steroid receptor coactivator-1 (SRC1; 602691) expression. Thus, Shang and Brown (2002) concluded that cell type- and promoter-specific differences in coregulator recruitment determine the cellular response to selective estrogen receptor modulators.

Kumar et al. (2002) identified a naturally occurring short form of MTA1 (603526), which they called MTA1s, that contains a theretofore unknown sequence of 33 amino acids with an estrogen receptor-binding motif, leu-arg-ile-leu-leu (LRILL). MTA1s localizes in the cytoplasm, sequesters estrogen receptor in the cytoplasm, and enhances nongenomic responses of estrogen receptor. Deletion of the LRILL motif in MTA1s abolished its corepressor function and its interaction with estrogen receptor, and restored nuclear localization of estrogen receptor. Dysregulation of human epidermal growth factor receptor-2 (ERBB2; 164870) in breast cancer cells enhanced the expression of MTA1s and the cytoplasmic sequestration of estrogen receptor. Expression of MTA1s in breast cancer cells prevented ligand-induced nuclear translocation of estrogen receptor and stimulated malignant phenotypes. MTA1s expression is increased in human breast tumors with no or low nuclear estrogen receptor. The regulation of the cellular localization of estrogen receptor by MTA1s represents a mechanism for redirecting nuclear receptor signaling by nuclear exclusion.

Auboeuf et al. (2002) examined the impact of transcription mediated by steroid receptors, including progesterone and estrogen receptors, on RNA processing using reporter genes subject to alternative splicing driven by steroid-sensitive promoters. Steroid hormones affected the processing of pre-mRNA synthesized from steroid-sensitive promoters, but not from steroid unresponsive promoters, in a steroid receptor-dependent and receptor-selective manner. Several nuclear receptor coregulators showed differential splicing effects, suggesting that steroid hormone receptors may simultaneously control gene transcription activity and exon content of the product mRNA by recruiting coregulators involved in both processes.

Flouriot et al. (2000) showed that the 46-kD isoform of ESR1 can heterodimerize with the 66-kD isoform and can competitively inhibit the ligand-independent transactivation function of the larger receptor.

Kos et al. (2002) noted that the various ESR1 promoters are utilized in a tissue-specific manner to control the levels of mRNA variants in individual tissues. They looked specifically at the effect of upstream AUGs on the translation of major ESR1 variants. The 5-prime untranslated region of 1 mRNA species significantly suppressed the translation of a reporter gene in transient transfection assays. The 5-prime untranslated region of 3 others showed moderate negative effects. Toeprinting analysis revealed that leaky scanning occurs with mRNAs bearing the most inhibitory 5-prime untranslated regions. Kos et al. (2002) concluded that short upstream open reading frames in the 5-prime untranslated regions of ESR1 mRNAs can regulate receptor expression.

Reid et al. (2003) established the ubiquitination status and subnuclear distribution of ESR, its mobility, the kinetics of transcriptional activation, and the cyclic recruitment of E3 ligases and the 19S regulatory component of the proteasome. They demonstrated that proteasome-mediated degradation and ESR-mediated transactivation are linked and act to turn over ESR continuously on responsive promoters, and concluded that the cyclic turnover of ESR permits continuous responses to changes in the concentration of estradiol.

Zhao et al. (2003) noted that the crystal structure of the ER-alpha ligand-binding domain suggests that leu536 may be involved in hydrophobic interactions at the start of helix 12. They found that certain mutations of leu536 increased the ligand-independent activity of ER-alpha and reduced or eliminated the agonist activity of 17-beta-estradiol and 4-hydroxytamoxifen in a reporter assay. Zhao et al. (2003) concluded that leu536 is critical in coupling the binding of ligand to changes in the conformation and transcriptional activity of ER-alpha. A heterodimer of the dioxin receptor (AHR; 600253) and ARNT (126110), which are basic helix-loop-helix/PAS family transcription factors, mediates most of the toxic effects of dioxins.

Li et al. (2003) showed that the ESR46 isoform more efficiently modulates membrane-initiated estrogen actions, including activation of eNOS, than full-length ESR66. Conversely, ESR66 more efficiently mediates estrogen response element reporter-gene transactivation than ESR46.

Ohtake et al. (2003) demonstrated that the agonist-activated AHR/ARNT heterodimer directly associates with the estrogen receptors ESR-alpha and ESR-beta. They showed that this association results in the recruitment of unliganded estrogen receptor and the coactivator p300 (602700) to estrogen-responsive gene promoters, leading to activation of transcription and estrogenic effects. The function of liganded estrogen receptor was found to be attenuated. Estrogenic actions of AHR agonists were detected in wildtype ovariectomized mouse uteri, but were absent in Ahr -/- or Er-alpha -/- ovariectomized mice. Ohtake et al. (2003) concluded that their findings suggest a novel mechanism by which estrogen receptor-mediated estrogen signaling is modulated by a coregulatory-like function of activated AHR/ARNT, giving rise to adverse estrogen-related actions of dioxin-type environmental contaminants.

Garcia-Morales et al. (1994) found that cadmium is a potent stimulator of the estrogen receptor independent of estradiol. Stoica et al. (2000) found that cadmium activates ESR1 at concentrations as low as 10(-11) M. Cadmium was found to block the binding of estradiol to ESR1 in a noncompetitive manner, suggesting that the heavy metal interacts with the hormone-binding domain of the receptor. Stoica et al. (2000) showed that cadmium activates ESR1 through an interaction with the hormone-binding domain of the receptor. Transfection and binding assays with ESR1 mutants identified cys381, cys447, glu523, his524, and asp538 as possible interaction sites of cadmium with the hormone-binding domain of ER-alpha. Johnson et al. (2003) showed that cadmium has a potent estrogen-like activity in vivo. Female rats were ovariectomized on postnatal day 28 and allowed to recover for 3 weeks. Rats were then given a single intraperitoneal dose of cadmium (5 microgram/per kg body weight or approximately 27 nmol/kg). This exposure to cadmium increased uterine weight, promoted growth and development of mammary glands, and induced hormone-regulated genes in ovariectomized animals. In the uterus, the increase in wet weight was accompanied by proliferation of the endometrium and induction of progesterone receptor (Pgr; 607311), and complement component C3 (120700). In the mammary gland, cadmium promoted an increase in the formation of side branches and alveolar buds and the induction of casein, whey acidic protein, Pgr, and C3. In utero exposure to the metal also mimicked the effects of estrogens. Female offspring experienced an earlier onset of puberty and an increase in the epithelial area and the number of terminal end buds in the mammary gland.

Using chromatin immunoprecipitation assays, Metivier et al. (2003) identified protein complexes recruited by ESR1 to the pS2 (TFF1; 113710) promoter in a breast carcinoma cell line following estradiol treatment, and they determined the order in which the complexes were recruited.

By immunoprecipitation of human breast cancer cell lines and protein pull-down assays with in vitro translated proteins, Wada-Hiraike et al. (2005) demonstrated that ESR1 interacted with MSH2 (609309) in a ligand-dependent manner, whereas ESR2 (601663) and MSH2 interacted in a ligand-independent manner. Both receptors bound MSH2 through its MSH3 (600887)/MSH6 (600678)-interaction domain. In a transient expression assay, MSH2 potentiated the transactivation function of ligand-activated ESR1 but not ESR2. Wada-Hiraike et al. (2005) concluded that MSH2 may be a coactivator of ESR1-dependent gene expression.

Wei et al. (2006) found that the MUC1 (158340) C-terminal subunit associated with ESR1 and that the interaction was stimulated by 17-beta-estradiol in human breast carcinoma cell lines. MUC1 bound directly to the ESR1 DNA-binding domain and stabilized ESR1 by blocking its ubiquitination and degradation. Chromatin immunoprecipitation assays demonstrated that MUC1 associated with ESR1 complexes on estrogen-responsive promoters, enhanced ESR1 promoter occupancy, and increased recruitment of p160 (PELP1; 609455) coactivators SRC1 and GRIP1 (604597). MUC1 stimulated ESR1-mediated transcription and contributed to estradiol-mediate d growth and survival of breast cancer cells.

Aguirre et al. (2007) demonstrated that extracellular signal-regulated kinases (see ERKs, 176948) are not activated by stretching in osteocytic and osteoblastic cells in which both ESR1 and ESR2 have been knocked out or knocked down; this effect was partially reversed by transfection of either of the 2 human ESRs, and fully by transfection of both receptors. ERK activation in response to stretching was also recovered by transfecting the ligand-binding domain of either receptor or an ESR1 mutant that does not bind estrogens. Mechanoresponsiveness was restored by transfecting ESR1 targeted to the plasma membrane but not to the nucleus, and ESR1 mutants with impaired plasma membrane localization or binding to caveolin-1 (601047) failed to confer ERK activation in response to stretching. An ESR antagonist abrogated ERK activation as well as the antiapoptotic effect of mechanical stimulation. Aguirre et al. (2007) concluded that in addition to their role as ligand-dependent mediators of the effects of estrogens, the ESRs participate in the transduction of mechanical forces into prosurvival signaling in bone cells in a ligand-independent manner.

Perillo et al. (2008) analyzed how H3 histone methylation and demethylation control expression of estrogen-responsive genes and showed that a DNA-bound estrogen receptor directs transcription by participating in bending chromatin to contact the RNA polymerase II (see 180660) recruited to the promoter. This process is driven by receptor-targeted demethylation of H3K9 (see 602810) at both enhancer and promoter sites and is achieved by activation of resident LSD1 (AOF2; 609132) demethylase. Localized demethylation produces hydrogen peroxide, which modifies the surrounding DNA and recruits 8-oxoguanine-DNA glycosylase-1 (601982) and topoisomerase II-beta (126431), triggering chromatin and DNA conformational changes that are essential for estrogen-induced transcription. Perillo et al. (2008) concluded that their data showed a strategy that uses controlled DNA damage and repair to guide productive transcription.

Hurtado et al. (2008) showed that estrogen-estrogen receptor (ER) and tamoxifen-ER complexes directly repress ERBB2 (164870) transcription by means of a cis-regulatory element within the ERBB2 gene in human cell lines. Hurtado et al. (2008) implicated the paired box-2 gene product (PAX2; 167409) in a previously unrecognized role, as a crucial mediator of ER repression of ERBB2 by the anticancer drug tamoxifen. Hurtado et al. (2008) showed that PAX2 and the ER coactivator AIB1/SRC3 (601937) compete for binding and regulation of ERBB2 transcription, the outcome of which determines tamoxifen response in breast cancer cells. The repression of ERBB2 by ER-PAX2 links these 2 breast cancer subtypes and suggests that aggressive ERBB2-positive tumors can originate from ER-positive luminal tumors by circumventing this repressive mechanism. Hurtado et al. (2008) concluded that their data provided mechanistic insight into the molecular basis of endocrine resistance in breast cancer.

Wang et al. (2008) found that human HPIP (PBXIP1; 618819) interacted with both ER-alpha and ER-beta in mammalian cells. Overexpression and knockdown analyses revealed that HPIP interaction increased expression of ER-alpha target genes by enhancing phosphorylation of ER-alpha at ser167 by MAPK and AKT. Immunoprecipitation experiments demonstrated that ER-beta also interacted with ER-alpha, thereby decreasing binding of ER-alpha to HPIP and inhibiting expression of ER-alpha target genes.

Fullwood et al. (2009) described the development of a new strategy, which they called chromatin interaction analysis by paired-end tag sequencing (ChIA-PET), for the de novo detection of global chromatin interactions, with which they comprehensively mapped the chromatin interaction network bound by estrogen receptor-alpha (ER-alpha) in the human genome. Fullwood et al. (2009) found that most high-confidence remote ER-alpha-binding sites are anchored at gene promoters through long-range chromatin interactions, suggesting that ER-alpha functions by extensive chromatin looping to bring genes together for coordinated transcriptional regulation. Fullwood et al. (2009) proposed that chromatin interactions constitute a primary mechanism for regulating transcription in mammalian genomes.

Using breast cancer and other cancer cell lines, Hurtado et al. (2011) showed that FOXA1 (602294) mediated ER binding and function. Almost all ER-chromatin interactions and gene expression changes depended on FOXA1, and FOXA1 influenced genomewide chromatin accessibility. FOXA1 was also required for the inhibitory activity of tamoxifen against ER.

The classical sex steroid receptors, including ER-alpha, translocate from the plasma membrane to the nucleus following ligand binding and influence gene transcription. Sex steroid receptors also engage in cell signaling at the plasma membrane, and this function requires both ligand binding and receptor palmitoylation at a conserved cysteine residue. Palmitoylacyltransferases contain an asp-his-his-cys (DHHC) signature sequence. By expressing plasmids encoding each mouse DHHC protein in MCF-7 human breast cancer cells, Pedram et al. (2012) found that only Dhhc7 (614604) and Dhhc21 (614605) functioned in ER-alpha palmitoylation. Knockdown of endogenous DHHC7 and DHHC21 via small interfering RNA reduced ER-alpha localization at the plasma membrane and impaired intracellular signaling by 17-beta-estradiol via ERK (see 176948) and AKT kinases and cAMP generation. Nuclear ER-alpha localization and transcriptional activation were unaffected. Mutation of a cysteine in the C-terminal region of mouse ER-alpha abrogated palmitoylation by Dhhc7 and Dhhc21 and inhibited cell signaling from the plasma membrane.

Using a constitutively active mutant of the mouse nonreceptor tyrosine phosphatase Shp2 (PTPN11; 176876), He et al. (2012) found that Shp2 integrated leptin (LEP; 164160) and estrogen signaling in transgenic female mice. Transgenic females, but not males, were resistant to high-fat diet-induced obesity and liver steatosis via enhanced leptin and insulin sensitivity and downstream ERK activation. SHP2 and ESR1 interacted directly in MCF-7 cells and female mouse tissues, and the interaction was enhanced by estrogen stimulation. Ovariectomy of transgenic mice reversed their resistance to high-fat diet-induced obesity.

Hypercholesterolemia is a risk factor for ER-positive breast cancers and is associated with a decreased response of tumors to endocrine therapies. Nelson et al. (2013) showed that 27-hydroxycholesterol (27HC), a primary metabolite of cholesterol and an ER and liver X receptor (see LXRA, 602423) ligand, increases ER-dependent growth and LXR-dependent metastasis in mouse models of breast cancer. The effects of cholesterol on tumor pathology required its conversion to 27HC by the cytochrome P450 oxidase CYP27A1 (606530) and were attenuated by treatment with CYP27A1 inhibitors. In human breast cancer specimens, CYP27A1 expression levels correlated with tumor grade. In high-grade tumors, both tumor cells and tumor-associated macrophages exhibited high expression levels of the enzyme. Thus, Nelson et al. (2013) concluded that lowering circulating cholesterol levels or interfering with its conversion to 27HC may be a useful strategy to prevent and/or treat breast cancer.

Kim et al. (2013) showed that CAC1 (CACUL1; 618764) and ER-alpha interacted in a ligand-independent manner and colocalized to nucleus in transacted H1299 cells. Interaction with CAC1 repressed ER-alpha transcriptional activity, and CoRNR box-2 of CAC1 was required for both ER-alpha binding and repression. CAC1 interacted with ER-alpha coactivator LSD1 and inhibited its binding to the ER-alpha target promoter, thereby repressing LSD1-enhanced ER-alpha activity. In addition, CAC1 increased paclitaxel-induced cell death, as overexpression of CAC1 in paclitaxel-resistant MCF7 cells increased their sensitivity to paclitaxel by suppressing ER-alpha activation.

The ventrolateral subdivision of the murine ventromedial hypothalamus (VMHvl) contains neurons whose activity increases during male-male and male-female social encounters. Non-cell-type-specific optogenetic activation of this region elicited attack behavior, but not mounting. Lee et al. (2014) identified a subset of VMHvl neurons marked by Esr1, and investigated their role in male social behavior. Optogenetic manipulations indicated that Esr1-positive (but not Esr1-negative) neurons are sufficient to initiate attack, and that their activity is continuously required during ongoing agonistic behavior. Surprisingly, weaker optogenetic activation of these neurons promoted mounting behavior, rather than attack, toward both males and females, as well as sniffing and close investigation. Increasing photostimulation intensity could promote a transition from close investigation and mounting to attack, within a single social encounter. Importantly, time-resolved optogenetic inhibition experiments revealed requirements for Esr1-positive neurons in both the appetitive (investigative) and the consummatory phases of social interactions. Combined optogenetic activation and calcium imaging experiments in vitro, as well as c-Fos analysis in vivo, indicated that increasing photostimulation intensity increases both the number of active neurons and the average level of activity per neuron. Lee et al. (2014) concluded that these data suggested that Esr1-positive neurons in VMHvl control the progression of a social encounter from its appetitive through its consummatory phases, in a scalable manner that reflects the number or type of active neurons in the population.

Cho et al. (2015) used ribosome profiling and RNA sequencing to quantify changes in gene and protein expression in mouse hippocampus following contextual fear conditioning. They identified 3 phases of alterations: an initial wave of translational regulation at 5 to 10 minutes, a second wave of induction of immediate early genes at 10 to 30 minutes, and suppression of genes through decreased mRNA levels after 30 minutes, which continued through 4 hours. Ingenuity pathway analysis identified Esr1 as the most prominent upstream regulator of differentially expressed genes (DEGs) at 4 hours, and Esr1 was predicted to be inhibited. Half of the decreased DEGs at 4 hours were putative downstream targets of Esr1, and several other DEGs were downstream targets of Otx1 (600036), an Esr1 downstream target. Quantitative RT-PCR on hippocampal RNAs following administration of an Esr1 antagonist showed marked decreases in Otx1 and other putative Esr1 targets. Administration of an Esr1 agonist into mouse hippocampus after learning significantly impaired memory formation compared with controls in 2 hippocampus-dependent tasks. Cho et al. (2015) concluded that downregulation of ESR1 signaling is important for memory formation.

Mohammed et al. (2015) showed that the progesterone receptor (PR; 607311) is not merely an ER-alpha-induced gene target, but is also an ER-alpha-associated protein that modulates its behavior. In the presence of agonist ligands, PR associates with ER-alpha to direct ER-alpha chromatin binding events within breast cancer cells, resulting in a unique gene expression program that is associated with good clinical outcome. Progesterone inhibited estrogen-mediated growth of ER-alpha(+) cell line xenografts and primary ER-alpha(+) breast tumor explants, and had increased antiproliferative effects when coupled with an ER-alpha antagonist. Copy number loss of the PGR gene is a common feature in ER-alpha(+) breast cancers, explaining lower PR levels in a subset of cases. Mohammed et al. (2015) concluded that their findings indicated that PR functions as a molecular rheostat to control ER-alpha chromatin binding and transcriptional activity.


Molecular Genetics

Estrogen Resistance

In a 28-year-old man with estrogen resistance (ESTRR; 615363), Smith et al. (1994) performed single-strand conformation polymorphism analysis of the ESR1 gene and observed a variant banding pattern in exon 2. Direct sequencing revealed a homozygous nonsense mutation (R157X; 133430.0002).

McInerney et al. (1996) characterized a human ESR1 variant, V364E (133430.0003), which they demonstrated to be a strong dominant-negative inhibitor of wildtype ESR1.

In an 18-year-old woman with estrogen resistance, Quaynor et al. (2013) identified homozygosity for a missense mutation in the ESR1 gene (Q375H; 133430.0006). Clegg and Palmer (2013) noted that the Esr1 -/AA knockin mouse model (see ANIMAL MODEL) recapitulates the phenotype of the female patient studied by Quaynor et al. (2013), in that they have hypoplastic mammary glands, anovulation, and altered steroidogenesis, but have normal body weight, adiposity, locomotor activity, and glucose homeostasis. These findings supported the proposal by Quaynor et al. (2013) that nonclassical regulation of ESR1 is sufficient to protect against the obesity-metabolic syndrome phenotype associated with total loss of activity of ESR1, but is not sufficient to rescue the infertility phenotype.

In 3 sibs from a consanguineous Algerian family with estrogen resistance, Bernard et al. (2017) identified homozygosity for a missense mutation in the ESR1 gene (R394H; 133430.0007) that segregated with disease.

Association with Breast Cancer

Zuppan et al. (1991) reported a lod score of 1.85 for linkage of ESR at zero recombination with late-onset breast cancer in 1 extended family with 8 affected members. Simulation of this pedigree assuming independent inheritance of breast cancer and ESR genotypes led to a lod score equal to or greater than 1.85 only once in 2,000 replicates. Zuppan et al. (1991) suggested testing linkage in other breast cancer families of the late-onset type.

McGuire et al. (1991) used the screening techniques of chemical mismatch cleavage, single-strand conformation polymorphism (SSCP), and gel retardation to discover a number of estrogen receptor mRNA variants in tissues from breast cancers. They identified basepair insertions, transitions, and deletions as well as alternative splicing, yielding deletions of exon 3, 5, or 7. Using a yeast transactivation assay, they discovered receptors with 'outlaw' function, including both a dominant-positive receptor which was transcriptionally active in the absence of estrogen, and a dominant-negative receptor, which was itself transcriptionally inactive but prevented the action of normal estrogen receptor. McGuire et al. (1991) concluded that these variants could have clinical significance, helping to explain differences in the behavior of breast tumors and patient outcome.

Ponglikitmongkol et al. (1988) found that the ESR isolated from a human breast cancer cell line contained a gly400-to-val mutation in the hormone-binding domain. A highly conserved 66-amino acid region of the estrogen and glucocorticoid receptors, which corresponds to part of the receptor DNA-binding domain (region C), determines the specificity of target gene recognition. This region contains 2 subregions (CI and CII), encoded by 2 separate exons that are analogous to 'zinc fingers.' By the study of chimeric estrogen receptor, Mader et al. (1989) showed that 3 amino acids located on the C-terminal side of the CI finger play a key role in the specificity.

It is accepted that the presence of estrogen receptor identifies those breast cancer patients with a lower risk of relapse and better overall survival (Clark and McGuire, 1988), and the measurement of ESR has become a standard assay in the clinical management of breast cancer. Receptor status also provides a guideline for those tumors that may be responsive to hormonal intervention. But only about half of ESR-positive patients respond to various hormonal therapies and of those who do respond initially, most will eventually develop hormonally unresponsive disease following a period of treatment even though ESR is often still present. Sluyser and Mester (1985) hypothesized that the loss of hormone dependence of certain breast tumors may be due to the presence of mutated or truncated steroid receptors that activate transcription even in the absence of hormone. Fuqua et al. (1993) reviewed ESR mutations that may be important in breast cancer progression. Scott et al. (1991), for example, had found truncated forms of DNA-binding ESR in human breast cancer. To better understand structure-activity relationships of the human estrogen receptor, Weis et al. (1996) examined the role of tyrosine-537 in transcriptional response of the receptor. This residue is close to a region of the hormone-binding domain shown previously to be important in hormone-dependent transcriptional activity; it also has been proposed to be a tyrosine kinase phosphorylation site important in ESR activity. Weis et al. (1996) substituted 5 amino acids at this position (alanine, phenylalanine, glutamic acid, lysine, or serine) and screened these mutants for their biologic activities in the presence and absence of estradiol. Two of the ESR mutants, tyr537 to ala (Y537A) and tyr537 to ser (Y537S), displayed estrogen-independent constitutive activity that was approximately 20% or 100%, respectively, of the activity of the wildtype receptor with estradiol. In some circumstances, the tyr537-to-glu (Y537E) and tyr537-to-lys (Y537K) proteins also exhibited some low level of constitutive activity. Their findings indicated that tyrosine-537 is in a region important in the ligand regulation of ESR transcriptional activity and that certain amino acid substitutions at this position can shift ESR into a conformation that is active even without ligand.

Sluyser (1995) reviewed the somatically generated mutations in ESR that had been found at the mRNA/cDNA level in human breast cancer biopsies and in established breast cancer cell lines. Aberrantly spliced ESR mRNA causes the appearance of truncated or internally deleted ESR protein forms. Studies on the functional activity of ESR variants in expression systems demonstrated dominant-positive receptors that are transcriptionally active in the absence of estrogen, and dominant-negative receptors that are themselves transcriptionally inactive but that prevent the action of the normal receptor. The ESR variants are believed to confer resistance to endocrine therapy in breast cancer patients. Abnormally spliced forms of ESR, similar to those in breast cancer, were reported by McGuire et al. (1992) and by others. In all, 19 somatic mutations were tabulated and mapped on a diagram of the structural organization of the ESR gene.

Andersen et al. (1997) studied leukocyte DNA from 143 patients with familial clustering of breast and/or ovarian cancer and tumor DNA from 96 breast carcinomas for base mutations in the ESR gene. Three patients with a family history of cancer were carrying a gly160-to-cys germline substitution, which they concluded represents a polymorphism because it was detected in 4 females and 4 males of 729 controls, split about equally between males and females. However, in the 229 female controls in whom family history of cancer was known, 1 of 2 who had a sister with breast cancer was carrying the variant allele; hence, a possible clinical significance of the gly160-to-cys change should be further investigated. Somatic mutations were not detected in any of the tumors studied, and the data did not provide support for somatic ESR base mutations as an important mechanism for hormonal therapy resistance in estrogen receptor-positive breast carcinomas.

Using an Affymetrix 10K SNP array to screen for gene copy number changes in breast cancer, Holst et al. (2007) detected a single-gene amplification of the ESR1 gene. A subsequent tissue microarray analysis of more than 2,000 clinical breast cancer samples showed ESR1 amplification in 20.6% of breast cancers. In 99% of tumors with ESR1 amplification, overexpression of estrogen receptor protein was demonstrated, compared with 66.6% of cancers without ESR1 amplification. In 175 women who had received adjuvant tamoxifen monotherapy, survival was significantly longer for women with cancer with ESR1 amplification than for women with estrogen receptor-expressing cancers without ESR1 amplification (P = 0.023). Notably, they also found ESR1 amplification in benign and precancerous breast diseases, suggesting that ESR1 amplification may be a common mechanism in proliferative breast disease and a very early genetic alteration in a large subset of breast cancers.

In correspondences, Brown et al. (2008), Horlings et al. (2008), Vincent-Salomon et al. (2008), and Reis-Filho et al. (2008) reported attempts to replicate the finding of Holst et al. (2007) of a high frequency of ESR1 amplification in breast cancer. No group was able to replicate the results of Holst et al. (2007), using a variety of methods including array comparative genomic hybridization (CGH), FISH, and quantitative PCR. Amplification was found at a frequency of approximately 10% or less (Albertson, 2008). In a discussion of the findings of all of these groups, Albertson (2008) noted that although Holst et al. (2007) reported to have followed the standard procedure for scoring FISH, i.e., to count closely spaced signals as 1 signal, in their reply to the contesting groups Holst et al. (2008) emphasized the importance of scoring clusters of signals. Holst et al. (2008) stated, 'In our laboratory, most ESR1-amplified tumors have small gene clusters that could be considered as one signal if 'ERBB criteria' were applied....We therefore feel that estimating the ESR1 gene copy number may--given the currently available reagents--enable a more reliable identification of amplified cancers than classical counting.' Albertson (2008) concluded that this and other discrepancies, including that involving the clinical significance concerning prognosis, indicated that 'the jury is still out on the question of ESR1 amplification and its clinical significance.'

Certain malignant breast tumors (see 114480) are characterized by a high prediction uncertainty ('low-confidence') with respect to ESR status. Kun et al. (2003) analyzed these 'low-confidence' tumors and determined that their uncertain prediction status arose as a result of widespread perturbations in multiple genes whose expression is important for ESR-subtype discrimination. Patients with 'low-confidence' ESR-positive tumors exhibited a significantly worse overall survival (p = 0.03) and shorter time to distant metastasis (p = 0.004) compared with their 'high-confidence' ESR-positive counterparts. Elevated expression of ERBB2 (164870) was significantly correlated with a breast tumor exhibiting a 'low-confidence' prediction. Although ERBB2 signaling has been proposed to inhibit the transcriptional activity of ESR1, a large proportion of the perturbed genes in the 'low-confidence'/ERBB2-positive samples are not known to be estrogen responsive. Kun et al. (2003) proposed that a significant portion of the effect of ERBB2 on ESR-positive breast tumors may involve ESR-independent mechanisms of gene activation, which may contribute to the clinically aggressive behavior of the 'low-confidence' breast tumor subtype.

Toy et al. (2013) conducted a comprehensive genetic analysis of 2 independent cohorts of metastatic ER-positive breast tumors and identified mutations in ESR1 affecting the ligand-binding domain (LBD) in 14 of 80 cases. These included highly recurrent mutations encoding Y537S, Y537N, and asp538 to gly (D538G) alterations. Molecular dynamics simulations suggested that the structures of the Y537S and D538G mutants involve hydrogen bonding of the mutant amino acids with asp351, thus favoring the agonist conformation of the receptor. Consistent with this model, mutant receptors drive ER-dependent transcription and proliferation in the absence of hormone and reduce the efficacy of ER antagonists.

Robinson et al. (2013) enrolled 11 patients with ER-positive metastatic breast cancer in a prospective clinical sequencing program for advanced cancers. Whole-exome and transcriptome analysis identified 6 cases that harbored mutations of ESR1 affecting its LBD, all of whom had been treated with antiestrogens and estrogen deprivation therapies. A survey of The Cancer Genome Atlas identified 4 endometrial cancers with similar mutations of ESR1. The 5 LBD-localized ESR1 mutations identified, encoding L536Q, Y537S, Y537C, Y537, and D538G, were shown to result in constitutive activity and continued responsiveness to antiestrogen therapies in vitro.

Association with Bone Mineral Density Variation

Lorentzon et al. (1999) investigated the influence of ESRA gene polymorphism and estradiol on height and bone density during and after puberty in males. Using the restriction enzymes XbaI and PvuII, the allelic variants XX, Xx, xx, PP, Pp, and pp were identified in 90 Caucasian boys. In a multivariate analysis including pubertal development, physical activity, and body weight, the XbaI genotype independently predicted total body BMD, head BMD, and spine volumetric BMD (P less than 0.05). The PvuII genotype independently predicted spine volumetric BMD (pp greater than PP; P of 0.01). The 20 boys with the PP allelic variant were found to have a greater body height than the other 70 boys (182 cm vs 179 cm; P of 0.03). At a 2-year follow-up, the XbaI genotype was still independently related to total body BMD, head BMD, and spine volumetric BMD. The authors concluded that ESRA polymorphism is related to bone density and height during late puberty and at attainment of peak bone density in young men.

Bone mineral density, the major determinant of osteoporotic fracture risk, has a strong genetic component. The discovery that inactivation of ESR1 is associated with low BMD indicated ESR1 as a candidate gene for osteoporosis (166710). Becherini et al. (2000) genotyped 610 postmenopausal women for 3 ESR1 gene polymorphisms (intron 1 RFLPs PvuII and XbaI, and a (TA)n repeat 5-prime upstream of exon 1). Although no significant relationship between intron 1 RFLPs and BMD was observed, a statistically significant correlation between (TA)n-repeat allelic variants and lumbar BMD was observed (P = 0.04, ANCOVA), with subjects having a low number of repeats (TA less than 15) showing the lowest BMD values. The authors observed a statistically significant difference in the mean +/- SD number of (TA)n repeats between 73 analyzed women with a vertebral fracture and the nonfracture group, equivalent to a 2.9-fold increased fracture risk in women with a low number of repeats. The authors concluded that in their large sample the (TA)n polymorphism in ESR1 accounts for part of the heritable component of BMD and may prove useful in the prediction of vertebral fracture risk in postmenopausal osteoporosis.

Colin et al. (2003) studied the combined influence of polymorphisms in the ESR1 and the VDR (601769) genes on the susceptibility to osteoporotic vertebral fractures in 634 women aged 55 years and older. Three VDR haplotypes (1, 2, and 3) of the BsmI, ApaI, and TaqI restriction fragment length polymorphisms and 3 ESR1 haplotypes (1, 2, and 3) of the PvuII and XbaI RFLPs were identified. ESR1 haplotype 1 was dose-dependently associated with increased vertebral fracture risk corresponding to an odds ratio of 1.9 (95% confidence interval, 0.9-4.1) per copy of the risk allele. VDR haplotype 1 was overrepresented in vertebral fracture cases. There was a significant interaction (P = 0.01) between ESR1 haplotype 1 and VDR haplotype 1 in determining vertebral fracture risk. The association of ESR1 haplotype 1 with vertebral fracture risk was present only in homozygous carriers of VDR haplotype 1. The risk of fracture was 2.5 for heterozygous and 10.3 for homozygous carriers of ESR1 haplotype 1. These associations were independent of BMD. The authors concluded that interaction between ESR1 and VDR gene polymorphisms leads to increased risk of osteoporotic vertebral fractures in women, largely independent of BMD.

Van Meurs et al. (2003) investigated the influence of genetic variation in ESR1 on several bone parameters in 2,042 individuals of the Rotterdam Study, a prospective population-based cohort study of elderly individuals. They analyzed 3 polymorphic sites in the 5-prime region of the ESR1 gene: a (TA)n repeat in the promoter region, molecular haplotypes of the PvuII and XbaI RFLPs in intron 1, and inferred long-range haplotypes thereof. Linkage disequilibrium (LD) analysis between the PvuII-XbaI haplotype and the (TA)n repeat showed strong LD between the 2 sites. Reconstruction of long-range haplotypes over the entire 5-prime region revealed 6 frequent long-range haplotypes. In women only, there was an allele dosage effect of haplotype 'px' (p = 0.003) and a low number of (TA)n repeats (p = 0.008) with decreased lumbar spine BMD and decreased vertebral bone area (p = 0.016). There was also an increased vertebral fracture risk with evidence for an allele dosage effect with an odds ratio of 2.2 (95% CI, 1.3-3.5) for haplotype 'px' and an odds ratio of 2.0 (95% CI, 1.5-3.2) for a low number of (TA)n repeats. The ESR1 genotype-dependent fracture risk was largely independent of BMD and bone area. Van Meurs et al. (2003) concluded that ESR1 polymorphisms in the 5-prime promoter region are associated with vertebral fracture risk, lumbar spine BMD, and vertebral bone area in postmenopausal women, but not in men.

Khosla et al. (2004) studied relationships between polymorphisms of the ESRA and ESRB genes, BMD, and rates of bone loss in an age-stratified random sample of 283 Rochester, Minnesota, men aged 22 to 90 years. DNA was analyzed for the XbaI and PvuII ESRA and AluI ESRB polymorphisms. The X/P and x/p alleles of the ESRA gene were in strong linkage disequilibrium. BMD at multiple sites did not differ as a function of either the ESRA or ESRB genotype. However, the ESRA (but not ESRB) genotypes did modulate the relationship between BMD or rates of bone loss and bioavailable estradiol levels in these men. The authors concluded that the ESRA genotype may modulate the relationship between BMD or rates of bone loss and estrogen levels in men and that bone mass in men with the X or P alleles may be more susceptible to the consequences of estrogen deficiency (and conversely, benefit most from estrogen sufficiency) than in men with the xx or pp genotypes.

Sowers et al. (2004) conducted a 10-year prospective study of peak bone mass and its change in 604 women, aged 24 to 44 years at study initiation, and related changes in bone mineral density (BMD) and osteocalcin (OCN; 112260) concentrations to ESR1 gene polymorphisms (the XbaI and PvuII RFLPs) in 442 of these women. The authors concluded that while ESR1 genotype associations were statistically significant in explaining the rate of perimenopausal bone loss and its turnover, baseline BMI or becoming postmenopausal contributed more to the magnitude of the difference in bone change.

In 945 postmenopausal Scottish women who had not received hormone replacement therapy (non-HRT), Albagha et al. (2005) found that annual rates of femoral neck bone loss were 14% higher in subjects who carried 1 copy of the 'px' allele and 22% higher in those who carried 2 copies compared to those who did not carry the px haplotype (p = 0.009). The px haplotype was associated with lower femoral neck BMD in non-HRT postmenopausal women (p = 0.02), and with reduced calcaneal broadband ultrasound attenuation in the whole study population of 3,054 Scottish women (p = 0.005). Albagha et al. (2005) concluded that the ESR1 px haplotype is associated with reduced femoral neck BMD and increased rates of femoral neck bone loss in non-HRT postmenopausal women, and suggested that the association with broadband ultrasound attenuation may explain the fact that ESR1 intron 1 alleles predict osteoporotic fractures by a mechanism partly independent of differences in BMD.

Tobias et al. (2007) investigated whether the gain in area-adjusted bone mineral content (ABMC) in girls occurs in late puberty and examined whether the magnitude of this gain is related to ESR1 polymorphisms. For rs2234693 (PvuII) and rs9340799 (XbaI) polymorphisms, differences in spinal ABMC in late puberty were 2-fold greater in girls who were homozygous for the C and G alleles, respectively (P = 0.001). For rs7757956, the difference in total body less head ABMC in late puberty was 50% less in individuals homozygous or heterozygous for the A allele (P = 0.006). Tobias et al. (2007) concluded that gains in ABMC in late pubertal girls are strongly associated with ESR1 polymorphisms, suggesting that estrogen contributes to this process via an estrogen receptor-alpha-dependent pathway.

In a case-control study of 70 osteoporotic Mexican women and 70 nonosteoporotic female controls, Gomez et al. (2007) analyzed the (TA)n repeat and 2014G-A polymorphisms of the ESR1 gene and found that, with correction for population stratification, the 2014G allele was associated with osteoporosis (OR, 4.34; p = 0.006) in the Mexican population, whereas the TA repeat polymorphism was not.

In a genomewide association study to find common sequence variants that influence bone mineral density and low-trauma fractures in 3 populations of European descent, Styrkarsdottir et al. (2008) identified a complex pattern of association in the 6q25 region (see BMND11, 612114). SNPs in this region showed an association with BMD of both hip and spine, although no single SNP could fully explain the association. At least 3 SNPs were required to account for the overall association; one of these was in an intron of an ESR1 splice variant, and the other 2 were in the nearby C6ORF97 gene.

Association with Myocardial Infarction or Cardiovascular Risk Factors

In a study of 309 postmenopausal women with coronary artery disease, Herrington et al. (2002) found that women who were of a particular genotype at the ESR1 locus had an augmented response of HDL cholesterol to hormone replacement therapy (see 133430.0004).

In a study of 2,617 men and 3,791 postmenopausal women, aged 55 years and older, followed up over a period of 7 years or more, Schuit et al. (2004) found that postmenopausal women who carried the ESR1 haplotype 1 had an increased risk of myocardial infarction (MI; 608446) and ischemic heart disease, independent of known cardiovascular risk factors. In men, no association was observed. Haplotype 1 is comprise of 2 polymorphisms located in the first intron of the ESR1 gene, 397 bp (PvuII; rs2234693) and 351 bp (XbaI; rs9340799) upstream of exon 2.

Taguchi (2004) raised the question of whether the connection between ESR1 and MI might not be direct. Since Mattila et al. (1989) reported an association between poor dental health and acute MI in men, several studies had shown such an association in women as well (Emingil et al., 2000); Taguchi et al. (2001) reported a significant association between ESR1 polymorphisms and tooth loss in postmenopausal women. Thus, the association between ESR1 polymorphisms and MI in postmenopausal women could actually be due to an association between ESR1 polymorphisms and tooth loss. Another possibility that Taguchi (2004) raised was that endothelial dysfunction that may lead to MI contributes to periodontitis and subsequent tooth loss in postmenopausal women. A third possibility was that the association between ESR1 polymorphisms and tooth loss and between ESR1 polymorphisms and MI are both real, so that the association between tooth loss and MI in postmenopausal women in previous studies is spurious and actually due to the ESR1 polymorphisms.

Shearman et al. (2005) tested for interaction between smoking and ESR1 variation in association with plasma concentration of atherogenic small, low density lipoprotein (LDL) particles and LDL particle size. Among 1,727 unrelated subjects from the population-based Framingham Heart Study, women who smoked and had the common ESR1 rs2234693 TT genotype had more than 1.7-fold higher levels of small LDL particles than women with the alternative genotypes (P-value for smoking-genotype interaction = 0.001). Similar results were obtained for 3 other ESR1 variants including rs9340799, in the same linkage disequilibrium block. A similar substantial gender-specific result was also evident with a 975C-G variant, in a separate linkage disequilibrium block, in exon 4 (P = 0.003). Women who smoked and had specific, common ESR1 genotypes had a substantially higher plasma concentration of atherogenic small LDL particles. Significant results revealed a dose-dependent effect of smoking and were evident in both pre- and postmenopausal women. The reported association has the potential to explain the risks associated with estrogen use in certain women and an association between the ESR1 haplotype 1 (rs2234693 T allele and rs9340799 A allele) with increased MI and ischemic heart disease, independent of the standard, established cardiovascular risk factors.

Associations Pending Confirmation

For discussion of a possible association between variation in the ESR1 gene and osteoarthritis of the knee, see 165720.

For discussion of a possible association between variation in the ESR1 gene and the waist-to-hip ratio in women, see 605552.

For discussion of a possible association between variation in the ESR1 gene and age-related macular degeneration, see ARMD1 (603075).

For discussion of a possible association between variation in the ESR1 gene and age of onset of menopause, see 300488.


Animal Model

Korach (1994) investigated hormone responsiveness in genetic mutant mice without a functional estrogen receptor, created through gene knockout techniques. Both sexes of these mutant animals were infertile and showed a variety of phenotypic changes associated with the gonads, mammary glands, reproductive tracts, and skeletal tissues.

To clarify the role of estrogen signaling in reproductive tract development and function, Couse et al. (1999) generated mice lacking ESRA and ESRB by targeted disruption. ESRA/ESRB knockout males were infertile but possessed a grossly normal reproductive tract. They exhibited various stages of spermatogenesis, but the numbers and motility of epididymal sperm were reduced significantly. ESRA/ESRB knockout females exhibited proper differentiation of the mullerian-derived structures of the uterus, cervix, and upper vagina, but these structures were severely hypoplastic in adults. Similar uterine hypoplasia was observed in ESRA, but not in ESRB, knockout mice. The ovaries of adult ESRA/ESRB knockout females exhibited morphologic phenotypes that were clearly distinct from those of the prepubertal ESRA/ESRB knockout females and the individual estrogen receptor knockout mice. The double-knockout female ovaries had structures resembling seminiferous tubules of the testis. Within the lumen of the tubule-like structures were degenerating granulosa cells and cells resembling Sertoli cells of the testis. Couse et al. (1999) argued that certain characteristics of the adult ESRA/ESRB knockout ovary indicated redifferentiation of varying components rather than a developmental phenomenon: the absence of similar structures in prepubertal ESRA/ESRB knockout ovaries; the consistent spherical shape of the tubules, suggesting origination from a once healthy follicle; and age-related increases in the area of transdifferentiation. The ovaries of adult ESRA/ESRB knockout females expressed mullerian-inhibiting substance (600957), sulfated glycoprotein-2 (185430), and Sox9 (608160). Couse et al. (1999) concluded that the loss of both receptors leads to an ovarian phenotype that is distinct from that of the individual estrogen receptor knockout mutants, which indicates that both receptors are required for the maintenance of germ and somatic cells in the postnatal ovary.

Heine et al. (2000) found that male and female Esr1 knockout mice had hyperplasia and hypertrophy of adipocytes, insulin resistance, and glucose intolerance. The results provided evidence that estrogen/ESR1 signaling is critical in female and male white adipose tissue; obesity in the knockout males involved a mechanism of reduced energy expenditure rather than increased energy intake. Similar results were obtained by Jones et al. (2000) studying aromatase (CYP19; 107910) knockout mice, which cannot synthesize endogenous estrogens. Both male and female aromatase knockout mice progressively accumulated significantly more intraabdominal adipose tissue than their wildtype littermates and had elevated circulating levels of leptin and cholesterol, as well as elevated insulin levels and a striking accumulation of lipid droplets in the livers.

Davis et al. (2002) noted that studies in humans and rodent models had suggested that estrogen may provide protection against age-related cataracts. The presence of estrogen receptors in the eye indicates that estrogen protection may result from direct interactions with its receptors in the eye, instead of being an indirect consequence from effects on another tissue. Davis et al. (2002) validated the concept that estrogen is beneficial for the eye. In transgenic mice expressing ESR-delta-3, a natural variant of ESR1 with an in-frame deletion of exon 3 resulting from alternative splicing, they found that cortical cataracts spontaneously formed in females after puberty and progressed with age. ESR-delta-3 is a dominant-negative form of ESR-alpha that inhibits ESR-alpha function. Cataract formation could be prevented if the females were ovariectomized before, but not after, sexual maturity. Both male and female ESR-delta-3 mice developed cataracts after neonatal treatment with the potent estrogen DES. The incidence of spontaneous and DES-induced cataracts in ESR-delta-3 mice was 100%, whereas such cataracts were absent from wildtype mice. The data suggested that repression of estrogen action induces cortical cataract formation because estrogen is required to activate the ESR-delta-3 repressor. Evidence of DES-induced cataracts in the ESR-delta-3 males as well as the females suggested that estrogen is important in lens physiology in both sexes.

Using mice lacking functional Esr1, Lee et al. (2003) showed that bone in vivo undergoes an adaptive response to loading that is less effective in the absence of Esr1 and that osteoblast-like cells require Esr1 to proliferate in response to mechanical strain in vitro. Lee et al. (2003) speculated that as ESR1 expression in osteoblasts and osteocytes depends on estrogen concentration, a failure to maintain bone strength after menopause might be due to reduction in the activity of ESR1 in bone cells, thereby limiting their anabolic response to mechanical loading and allowing a loss of bone tissue comparable to that associated with disuse.

Garey et al. (2003) identified a generalized arousal component in the behavior of mice. Analyzed by mathematical/statistical approaches across experiments, investigators, and mouse populations, it accounted for approximately one-third of the variance in arousal-related measures. Knockout of the Esr1 gene greatly reduced arousal responses. In contrast, disrupting the Esr2 gene (601663), a likely gene duplication product which encodes ER-beta, did not have these effects.

In mice, ovariectomy accelerates the progression of the end-stage renal disease glomerulosclerosis. In women, the incidence of this disorder increases after menopause, and estrogen alters its progression. Polymorphisms in the ESR1 gene may constitute a genetic predisposition for lupus nephritis (Liu et al., 2002). Shim et al. (2004) showed that by 1 year of age, mice lacking ER-alpha, but not those lacking ER-beta, exhibited immune complex-type glomerulonephritis, proteinuria, and destruction of tubular cells. The mice also showed spontaneous formation of germinal centers in the spleen in the absence of antigen challenge and infiltration of plasma cells in the kidney and spleen. The results indicated that ER-alpha has indispensable functions in the kidney and in germinal centers, and that defective ER-alpha signaling results in glomerulonephritis.

In female mice and rats, Musatov et al. (2007) used RNAi to focally silence ER-alpha in the ventromedial nucleus of the hypothalamus and observed development of a phenotype characteristic of metabolic syndrome, marked by obesity, hyperphagia, impaired glucose tolerance, and reduced energy expenditure; this phenotype persisted despite normal ER-alpha levels elsewhere in the brain. Although an increase in food intake preceded weight gain, the authors stated that their data suggested that a leading factor of obesity in this model was likely a decline in energy expenditure with all 3 major constituents being affected, including voluntary activity, basal metabolic rate, and diet-induced thermogenesis. Musatov et al. (2007) concluded that ER-alpha in the ventromedial nucleus of the hypothalamus neurons plays an essential role in the control of energy balance and maintenance of normal body weight.

Nakamura et al. (2007) selectively ablated Esr1 in differentiated mouse osteoclasts and found that females, but not males, exhibited trabecular bone loss, similar to the osteoporotic bone phenotype in postmenopausal women. Furthermore, estrogen induced apoptosis and upregulation of Fas ligand (FASL, or TNFSF6; 134638) in osteoclasts of trabecular bones of wildtype mice, but not mutant mice. Expression of Esr1 was also required for induction of apoptosis by tamoxifen and estrogen in cultured osteoclasts. Nakamura et al. (2007) concluded that estrogen regulates the life span of mature osteoclasts via induction of the FAS (TNFRSF6; 134637)/FASL system.

To examine the ability of estrogen response element (ERE)-independent ESR1 signaling pathways to convey estrogen feedback regulation of the female hypothalamic-pituitary axis, Glidewell-Kenney et al. (2007) bred knockin mice expressing a mutant form of Esr1 ('AA') that has ablated ERE activity but intact ERE-independent activity with Esr1 -/- mice. The Esr1 -/AA mice exhibited 70% lower serum LH levels compared to Esr1 -/- mice. In addition, like wildtype mice, Esr1 -/AA mice exhibited elevated LH after ovariectomy, and the postovariectomy rise in LH was significantly suppressed by estrogen treatment in the ovariectomized Esr1 -/AA mice. However, unlike wildtype, both Esr1 -/AA and Esr1 -/- mice failed to exhibit estrous cyclicity, spontaneous ovulation, or an afternoon LH surge response to estrogen. Glidewell-Kenney et al. (2007) concluded that ERE-independent ESR1 signaling is sufficient to convey the majority of estrogen's negative feedback actions, whereas positive feedback and spontaneous ovulatory cyclicity require ERE-dependent ESR1 signaling.

Park et al. (2011) used Esr1 -/AA mice to assess the role of noncanonical ESR1 signaling in energy homeostasis and found that nonclassical ESR1 signaling restored metabolic parameters dysregulated in Esr1 -/- mice to normal or near-normal values. The rescue of body weight and metabolic function by nonclassical ESR1 signaling was mediated by normalization of energy expenditure, including voluntary locomotor activity. Park et al. (2011) concluded that nonclassical ESR1 signaling mediates major effects of estradiol-17-beta on energy balance.


ALLELIC VARIANTS ( 7 Selected Examples):

.0001 ESTROGEN RECEPTOR MUTANT, TEMPERATURE-SENSITIVE

ESR1, CYS447ALA
  
RCV000018058

Reese and Katzenellenbogen (1991) identified an estrogen receptor mutant that had a similar binding affinity for estradiol as wildtype ESR but displayed a dose-response shift for estradiol in transactivation studies. The mutant contained an alanine substitution for cysteine at amino acid 447 in the hormone binding domain of the receptor. Reese and Katzenellenbogen (1992) showed by hormone binding studies that the C447A receptor is a temperature-sensitive mutant, whose instability is only apparent at elevated temperatures, and that ligand can stabilize the mutant receptor. The mutant also showed a temperature-sensitive loss in the DNA binding ability of the receptor. (The mutant was one of several created by Reese and Katzenellenbogen (1991) by in vitro oligonucleotide site-directed mutagenesis of human ESR cDNA. Function of the mutant forms was tested in Chinese hamster ovary (CHO) cells, an estrogen receptor-deficient cell line.)


.0002 ESTROGEN RESISTANCE

ESR1, ARG157TER
  
RCV000018060

In a 28-year-old man with estrogen resistance (ESTRR; 615363), Smith et al. (1994) identified homozygosity for a C-T transition in exon 2 of the ESR1 gene, resulting in an arg157-to-ter (R157X) substitution. Both parents were heterozygous carriers of the R157X mutation, and pedigree analysis showed that they were related as second cousins; 3 sisters were also heterozygous for the mutation.


.0003 ESTROGEN RESISTANCE

ESR1, VAL364GLU
  
RCV000018061

McInerney et al. (1996) characterized a human ESR mutant, val364 to glu, which has a single amino acid substitution in its hormone-binding domain. While this mutant is fully active or even superactive at saturating levels of estradiol it also acts as a strong dominant-negative inhibitor of the wildtype ESR and it is able to repress ESR-mediated transcription when the mutant and wildtype ESR are present together in cells, even without DNA binding. It is probable that altered interactions with proteins important in ESR-mediated transcription play a key role in the repression of transcription by val364 to glu.


.0004 MYOCARDIAL INFARCTION, SUSCEPTIBILITY TO

ESR1, IVS1AS, T-C, -401
  
RCV000018062

Myocardial Infarction, Susceptibility to

In a study of atherosclerotic cardiovascular events in 1,739 unrelated men and women from the Framingham Heart Study, Shearman et al. (2003) found that after adjustment for covariates, the ESR1 C/C genotype was significantly associated with major atherosclerotic cardiovascular disease and myocardial infarction (608446) (odds ratios of 2.0 and 3.0, respectively, compared to individuals with the C/T or T/T genotypes). The results remained significant when analyses were restricted to men; too few women had events to study them separately. Shearman et al. (2003) concluded that individuals with the common ESR1 C/C genotype have a substantial increase in risk of myocardial infarction.

Atherosclerosis, Susceptibility to

Lehtimaki et al. (2002) examined coronary artery specimens from 300 Finnish white men aged 33 to 69 years included in the Helsinki Sudden Death Study and determined the ESR1 IVS1 -401T/C (or PvuII) genotype. After adjusting for age and BMI, men aged 53 years or over with C/T and C/C genotypes had areas of complicated lesions on average 2- and 5-fold larger, respectively, than subjects with the T/T genotype. The age and BMI-adjusted odds ratios for coronary thrombosis were 6.2 for C/T and 10.6 for C/C compared to men with the T/T genotype. After additional adjustment for diabetes and hypertension, ESR1 genotype persisted as an independent predictor of complicated lesions and coronary thrombosis. Lehtimaki et al. (2002) concluded that the ESR1 gene is a potential candidate behind the pathogenesis of acute coronary events.

HDL Cholesterol, Augmented Response of, to Hormone Replacement

Herrington et al. (2002) characterized 309 women with coronary artery disease with respect to 8 previously described and 2 novel ESR1 polymorphisms, and examined the association between these polymorphisms and the response of HDL cholesterol and other lipids to treatment with estrogen alone or estrogen plus progestin. They found that postmenopausal women who had the ESR1 C/C genotype at the -401 position in intron 1, or several other closely related genotypes, had an augmented response of HDL cholesterol to hormone replacement therapy.


.0005 MIGRAINE, SUSCEPTIBILITY TO

ESR1, 594G-A
  
RCV001674740...

Migraine (157300) is a painful and debilitating disorder with a significant genetic component. Steroid hormones, in particular estrogen, have long been considered to play a role in migraine, as variations in hormone levels are associated with migraine onset in many individuals with the disorder. Steroid hormones mediate their activity via hormone receptors, which have a wide tissue distribution. Estrogen receptors have been localized to the brain in regions considered to be involved in migraine pathogenesis. Colson et al. (2004) examined the ESR1 gene for a potential role in migraine pathogenesis and susceptibility. A population-based cohort of 224 patients with migraine and 224 matched controls were genotyped for the 594G-A polymorphism located in exon 8 of the ESR1 gene. Statistical analysis indicated a significant difference between patients with migraine and those without migraine in both the allele frequencies (P = 0.003) and genotype distributions (P = 0.008). An independent follow-up study using this marker in an additional population-based cohort of 260 patients with migraine and 260 matched controls resulted in a significant association between the 2 groups with regard to allele frequencies and genotype distributions. The findings supported the hypothesis that genetic variation in hormone receptors, in particular the ESR1 gene, may play a role in migraines.


.0006 ESTROGEN RESISTANCE

ESR1, GLN375HIS
  
RCV000054445

In an 18-year-old woman with estrogen resistance (ESTRR; 615363), Quaynor et al. (2013) identified homozygosity for a c.1125G-T transversion in exon 5 of the ESR1 gene, resulting in a gln375-to-his (Q375H) substitution at a highly conserved residue within the ligand-binding domain. The patient had been adopted, thus DNA was not available from her parents; however, high-density microarray analysis revealed an approximately 11% region of homozygosity, suggesting that her biologic parents were second-degree relatives and likely heterozygous carriers of the mutation. Functional analysis in transfected COS-7 cells showed greatly reduced activity in the mutated estrogen receptor, with an EC50 that was 240-fold greater than that of the wildtype receptor.


.0007 ESTROGEN RESISTANCE

ESR1, ARG394HIS
  
RCV000495838

In 3 sibs from a consanguineous Algerian family with estrogen resistance (ESTRR; 615363), Bernard et al. (2017) identified homozygosity for a c.1181G-A transition (c.1181G-A, NM_000125.3) in exon 5 of the ESR1 gene, resulting in an arg394-to-his (R394H) substitution at a highly conserved residue in the ligand-binding domain. Their unaffected parents and an unaffected sister were heterozygous for the mutation, which was not found in the Exome Variant Server or ExAC databases. Analysis of transiently transfected HEK293T cells demonstrated highly reduced sensitivity of the mutant receptor to 17-beta-estradiol stimulation, with an ED50 that was 65-fold higher than that of wildtype.


REFERENCES

  1. Abbondanza, C., Rossi, V., Roscigno, A., Gallo, L., Belsito, A., Piluso, G., Medici, N., Nigro, V., Molinari, A. M., Moncharmont, B., Puca, G. A. Interaction of vault particles with estrogen receptor in the MCF-7 breast cancer cell. J. Cell Biol. 141: 1301-1310, 1998. [PubMed: 9628887, images, related citations] [Full Text]

  2. Aguirre, J. I., Plotkin, L. I., Gortazar, A. R., Martin Millan, M., O'Brien, C. A., Manolagas, S. C., Bellido, T. A novel ligand-independent function of the estrogen receptor is essential for osteocyte and osteoblast mechanotransduction. J. Biol. Chem. 282: 25501-25508, 2007. [PubMed: 17609204, related citations] [Full Text]

  3. Albagha, O. M. E., Pettersson, U., Stewart, A., McGuigan, F. E. A., MacDonald, H. M., Reid, D. M., Ralston, S. H. Association of oestrogen receptor alpha gene polymorphisms with postmenopausal bone loss, bone mass, and quantitative ultrasound properties of bone. J. Med. Genet. 42: 240-246, 2005. [PubMed: 15744038, related citations] [Full Text]

  4. Albertson, D. G. Conflicting evidence on the frequency of ESR1 amplification in breast cancer. Nature Genet. 40: 821-822, 2008. [PubMed: 18583976, related citations] [Full Text]

  5. Andersen, T. I., Wooster, R., Laake, K., Collins, N., Warren, W., Skrede, M., Eeles, R., Tveit, K. M., Johnston, S. R. D., Dowsett, M., Olsen, A. O., Moller, P., Stratton, M. R., Borresen-Dale, A.-L. Screening for ESR mutations in breast and ovarian cancer patients. Hum. Mutat. 9: 531-536, 1997. [PubMed: 9195227, related citations] [Full Text]

  6. Auboeuf, D., Honig, A., Berget, S. M., O'Malley, B. W. Coordinate regulation of transcription and splicing by steroid receptor coregulators. Science 298: 416-419, 2002. [PubMed: 12376702, related citations] [Full Text]

  7. Balleine, R. L., Hunt, S. M. N., Clarke, C. L. Coexpression of alternatively spliced estrogen and progesterone receptor transcripts in human breast cancer. J. Clin. Endocr. Metab. 84: 1370-1377, 1999. [PubMed: 10199781, related citations] [Full Text]

  8. Becherini, L., Gennari, L., Masi, L., Mansani, R., Massart, F., Morelli, A., Falchetti, A., Gonnelli, S., Fiorelli, G., Tanini, A., Brandi, M. L. Evidence of a linkage disequilibrium between polymorphisms in the human estrogen receptor-alpha gene and their relationship to bone mass variation in postmenopausal Italian women. Hum. Molec. Genet. 9: 2043-2050, 2000. [PubMed: 10942433, related citations] [Full Text]

  9. Bernard, V., Kherra, S., Francou, B., Fagart, J., Viengchareun, S., Guechot, J., Ladjouze, A., Guiochon-Mantel, A., Korach, K. S., Binart, N., Lombes, M., Christin-Maitre, S. Familial multiplicity of estrogen insensitivity associated with a loss-of-function ESR1 mutation. J. Clin. Endocr. Metab. 102: 93-99, 2017. [PubMed: 27754803, related citations] [Full Text]

  10. Bord, S., Horner, A., Beavan, S., Compston, J. Estrogen receptors alpha and beta are differentially expressed in developing human bone. J. Clin. Endocr. Metab. 86: 2309-2314, 2001. [PubMed: 11344243, related citations] [Full Text]

  11. Brown, L. A., Hoog, J., Chin, S.-F., Tao, Y., Zayed, A. A., Chin, K., Teschendorff, A. E., Quackenbush, J. F., Marioni, J. C., Leung, S., Perou, C. M., Neilsen, T. O., Ellis, M., Gray, J. W., Bernard, P. S., Huntsman, D. G., Caldas, C. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 806-807, 2008. [PubMed: 18583964, related citations] [Full Text]

  12. Castagnoli, A., Maestri, I., Bernardi, F., Del Senno, L. PvuII RFLP inside the human estrogen receptor gene. Nucleic Acids Res. 15: 866 only, 1987. [PubMed: 2881267, related citations] [Full Text]

  13. Chaidarun, S. S., Alexander, J. M. A tumor-specific truncated estrogen receptor splice variant enhances estrogen-stimulated gene expression. Molec. Endocr. 12: 1355-1366, 1998. [PubMed: 9731704, related citations] [Full Text]

  14. Chiang, C.-H., Cheng, K. W., Igarashi, S., Nathwani, P. S., Leung, P. C. K. Hormonal regulation of estrogen receptor alpha and beta gene expression in human granulosa-luteal cells in vitro. J. Clin. Endocr. Metab. 85: 3828-3839, 2000. [PubMed: 11061546, related citations] [Full Text]

  15. Cho, J., Yu, N.-K., Choi, J.-H., Sim, S.-E., Kang, S. J., Kwak, C., Lee, S.-W., Kim, J., Choi, D. I., Kim, V. N., Kaang, B.-K. Multiple repressive mechanisms in the hippocampus during memory formation. Science 350: 82-87, 2015. [PubMed: 26430118, related citations] [Full Text]

  16. Clark, G. M., McGuire, W. L. Steroid receptors and other prognostic factors in primary breast cancer. Semin. Oncol. 15 (suppl. 1): 20-25, 1988. [PubMed: 3285482, related citations]

  17. Clegg, D. J., Palmer, B. F. Effects of an estrogen receptor alpha variant. (Letter) New Eng. J. Med. 369: 1663-1664, 2013. [PubMed: 24152274, related citations] [Full Text]

  18. Colin, E. M., Uitterlinden, A. G., Meurs, J. B. J., Bergink, A. P., van de Klift, M., Fang, Y., Arp, P. P., Hofman, A., van Leeuwen, J. P. T. M., Pols, H. A. P. Interaction between vitamin D receptor genotype and estrogen receptor alpha genotype influences vertebral fracture risk. J. Clin. Endocr. Metab. 88: 3777-3784, 2003. [PubMed: 12915669, related citations] [Full Text]

  19. Colson, N. J., Lea, R. A., Quinlan, S., MacMillan, J., Griffiths, L. R. The estrogen receptor 1 G594A polymorphism is associated with migraine susceptibility in two independent case/control groups. Neurogenetics 5: 129-133, 2004. [PubMed: 15133719, related citations] [Full Text]

  20. Couse, J. F., Hewitt, S. C., Bunch, D. O., Sar, M., Walker, V. R., Davis, B. J., Korach, K. S. Postnatal sex reversal of the ovaries in mice lacking estrogen receptors alpha and beta. Science 286: 2328-2331, 1999. [PubMed: 10600740, related citations] [Full Text]

  21. Davis, V. L., Chan, C.-C., Schoen, T. J., Couse, J. F., Chader, G. J., Korach, K. S. An estrogen receptor repressor induces cataract formation in transgenic mice. Proc. Nat. Acad. Sci. 99: 9427-9432, 2002. [PubMed: 12082175, images, related citations] [Full Text]

  22. Emingil, G., Buduneli, E., Aliyev, A., Akilli, A., Atilla, G. Association between periodontal disease and acute myocardial infarction. J. Periodont. 71: 1882-1886, 2000. [PubMed: 11156045, related citations] [Full Text]

  23. Esmaeli, B., Harvey, J. T., Hewlett, B. Immunohistochemical evidence for estrogen receptors in meibomian glands. Ophthalmology 107: 180-184, 2000. [PubMed: 10647739, related citations] [Full Text]

  24. Fan, S., Wang, J.-A., Yuan, R., Ma, Y., Meng, Q., Erdos, M. R., Pestell, R. G., Yuan, F., Auborn, K. J., Goldberg, I. D., Rosen, E. M. BRCA1 inhibition of estrogen receptor signaling in transfected cells. Science 284: 1354-1356, 1999. [PubMed: 10334989, related citations] [Full Text]

  25. 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-alpha) that is encoded by distinct transcripts and that is able to repress hER-alpha activation function 1. EMBO J. 19: 4688-4700, 2000. [PubMed: 10970861, images, related citations] [Full Text]

  26. Fullwood, M. J., Liu, M. H., Pan, Y. F., Liu, J., Xu, H., Mohamed, Y. B., Orlov, Y. L., Velkov, S., Ho, A., Mei, P. H., Chew, E. G. Y., Huang, P. Y. H., and 29 others. An oestrogen-receptor-alpha-bound human chromatin interactome. Nature 462: 58-64, 2009. [PubMed: 19890323, images, related citations] [Full Text]

  27. Fuqua, S. A. W., Chamness, G. C., McGuire, W. L. Estrogen receptor mutations in breast cancer. J. Cell. Biochem. 51: 135-139, 1993. [PubMed: 8440747, related citations] [Full Text]

  28. Garcia-Morales, P., Saceda, M., Kenney, N., Kim, N., Salomon, D. S., Gottardis, M. M., Solomon, H. B., Sholler, P. F., Jordan, V. C., Martin, M. B. Effect of cadmium on estrogen receptor levels and estrogen-induced responses in human breast cancer cells. J. Biol. Chem. 269: 16896-16901, 1994. [PubMed: 8207012, related citations]

  29. Garey, J., Goodwillie, A., Frohlich, J., Morgan, M., Gustafsson, J.-A., Smithies, O., Korach, K. S., Ogawa, S., Pfaff, D. W. Genetic contributions to generalized arousal of brain and behavior. Proc. Nat. Acad. Sci. 100: 11019-11022, 2003. [PubMed: 12930901, images, related citations] [Full Text]

  30. Glidewell-Kenney, C., Hurley, L. A., Pfaff, L., Weiss, J., Levine, J. E., Jameson, J. L. Nonclassical estrogen receptor alpha signaling mediates negative feedback in the female mouse reproductive axis. Proc. Nat. Acad. Sci. 104: 8173-8177, 2007. [PubMed: 17470805, images, related citations] [Full Text]

  31. Gomez, R., Magana, J. J., Cisneros, B., Perez-Salazar, E., Faugeron, S., Veliz, D., Castro, C., Rubio, J., Casas, L., Valdes-Flores, M. Association of the estrogen receptor alpha gene polymorphisms with osteoporosis in the Mexican population. Clin. Genet. 72: 574-581, 2007. [PubMed: 17953702, related citations] [Full Text]

  32. Gosden, J. R., Middleton, P. G., Rout, D. Localization of the human oestrogen receptor gene to chromosome 6q24-q27 by in situ hybridization. Cytogenet. Cell Genet. 43: 218-220, 1986. [PubMed: 3802924, related citations] [Full Text]

  33. Green, S., Walter, P., Kumar, V., Krust, A., Bornert, J.-M., Argos, P., Chambon, P. Human oestrogen receptor cDNA: sequence, expression and homology to v-erb-A. Nature 320: 134-139, 1986. [PubMed: 3754034, related citations] [Full Text]

  34. Greene, G. L., Gilna, P., Waterfield, M., Baker, A., Hort, Y., Shine, J. Sequence and expression of human estrogen receptor complementary DNA. Science 231: 1150-1154, 1986. [PubMed: 3753802, related citations] [Full Text]

  35. He, Z., Zhang, S. S., Meng, Q., Li, S., Zhu, H. H., Raquil, M.-A., Alderson, N., Zhang, H., Wu, J., Rui, L., Cai, D., Feng, G.-S. Shp2 controls female body weight and energy balance by integrating leptin and estrogen signals. Molec. Cell. Biol. 32: 1867-1878, 2012. [PubMed: 22431513, images, related citations] [Full Text]

  36. Heine, P. A., Taylor, J. A., Iwamoto, G. A., Lubahn, D. B., Cooke, P. S. Increased adipose tissue in male and female estrogen receptor-alpha knockout mice. Proc. Nat. Acad. Sci. 97: 12729-12734, 2000. [PubMed: 11070086, images, related citations] [Full Text]

  37. 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. New Eng. J. Med. 346: 967-974, 2002. [PubMed: 11919305, related citations] [Full Text]

  38. Holst, F., Stahl, P., Hellwinkel, O., Dancau, A.-M., Krohn, A., Wuth, L., Heupel, C., Lebeau, A., Terracciano, L., Al-Kuraya, K., Janicke, F., Sauter, G., Simon, R. Reply to 'ESR1 gene amplification in breast cancer: a common phenomenon?' (Letter) Nature Genet. 40: 810-812, 2008.

  39. Holst, F., Stahl, P. R., Ruiz, C., Hellwinkel, O., Jehan, Z., Wendland, M., Lebeau, A., Terracciano, L., Al-Kuraya, K., Janicke, F., Sauter, G., Simon, R. Estrogen receptor alpha (ESR1) gene amplification is frequent in breast cancer. Nature Genet. 39: 655-660, 2007. [PubMed: 17417639, related citations] [Full Text]

  40. Horlings, H. M., Bergamaschi, A., Nordgard, S. H., Kim, Y. H., Han, W., Noh, D.-Y., Salari, K., Joosse, S. A., Reyal, F., Lingjaerde, O. C., Kristensen, V. N., Borresen-Dale, A.-L., Pollack, J., van de Vijver, M. J. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 807-808, 2008. [PubMed: 18583965, related citations] [Full Text]

  41. Hurtado, A., Holmes, K. A., Geistlinger, T. R., Hutcheson, I. R., Nicholson, R. I., Brown, M., Jiang, J., Howat, W. J., Ali, S., Carroll, J. S. Regulation of ERBB2 by oestrogen receptor-PAX2 determines response to tamoxifen. Nature 456: 663-666, 2008. Note: Erratum: Nature 457: 1168 only, 2009. [PubMed: 19005469, images, related citations] [Full Text]

  42. Hurtado, A., Holmes, K. A., Ross-Innes, C. S., Schmidt, D., Carroll, J. S. FOXA1 is a key determinant of estrogen receptor function and endocrine response. Nature Genet. 43: 27-33, 2011. [PubMed: 21151129, images, related citations] [Full Text]

  43. Issa, J.-P. J., Ottaviano, Y. L., Celano, P., Hamilton, S. R., Davidson, N. E., Baylin, S. B. Methylation of the oestrogen receptor CpG island links ageing and neoplasia in human colon. Nature Genet. 7: 536-540, 1994. [PubMed: 7951326, related citations] [Full Text]

  44. Jeltsch, J. M., Roberts, M., Schatz, C., Garnier, J. M., Brown, A. M. C., Chambon, P. Structure of the human oestrogen-responsive gene pS2. Nucleic Acids Res. 15: 1401-1414, 1987. [PubMed: 3822834, related citations] [Full Text]

  45. Johnson, M. D., Kenney, N., Stoica, A., Hilakivi-Clarke, L., Singh, B., Chepko, G., Clarke, R., Sholler, P. F., Lirio, A. A., Foss, C., Reiter, R., Trock, B., Paik, S., Martin, M. B. Cadmium mimics the in vivo effects of estrogen in the uterus and mammary gland. Nature Med. 9: 1081-1084, 2003. [PubMed: 12858169, related citations] [Full Text]

  46. Jones, M. E. E., Thorburn, A. W., Britt, K. L., Hewitt, K. N., Wreford, N. G., Proietto, J., Oz, O. K., Leury, B. J., Robertson, K. M., Yao, S., Simpson, E. R. Aromatase-deficient (ArKO) mice have a phenotype of increased adiposity. Proc. Nat. Acad. Sci. 97: 12735-12740, 2000. [PubMed: 11070087, images, related citations] [Full Text]

  47. Justice, M. J., Siracusa, L. D., Gilbert, D. J., Heisterkamp, N., Groffen, J., Chada, K., Silan, C. M., Copeland, N. G., Jenkins, N. A. A genetic linkage map of mouse chromosome 10: localization of eighteen molecular markers using a single interspecific backcross. Genetics 125: 855-866, 1990. [PubMed: 1975791, related citations] [Full Text]

  48. Khosla, S., Riggs, B. L., Atkinson, E. J., Oberg, A. L., Mavilia, C., Del Monte, F., Melton, L. J., III, Brandi, M. L. Relationship of estrogen receptor genotypes to bone mineral density and to rates of bone loss in men. J. Clin. Endocr. Metab. 89: 1808-1816, 2004. [PubMed: 15070949, related citations] [Full Text]

  49. Kim, J., Park, U.-H., Moon, M., Um, S.-J., Kim, E.-J. Negative regulation of ER-alpha by a novel protein CAC1 through association with histone demethylase LSD1. FEBS Lett. 587: 17-22, 2013. [PubMed: 23178685, related citations] [Full Text]

  50. Korach, K. S. Insights from the study of animals lacking functional estrogen receptor. Science 266: 1524-1527, 1994. [PubMed: 7985022, related citations] [Full Text]

  51. Kos, M., Denger, S., Reid, G., Gannon, F. Upstream open reading frames regulate the translation of the multiple mRNA variants of the estrogen receptor-alpha. J. Biol. Chem. 277: 37131-37138, 2002. [PubMed: 12147702, related citations] [Full Text]

  52. Kos, M., Reid, G., Denger, S., Gannon, F. Minireview: genomic organization of the human ER-alpha gene promoter region. Molec. Endocr. 15: 2057-2063, 2001. [PubMed: 11731608, related citations] [Full Text]

  53. Kumar, R., Wang, R.-A., Mazumdar, A., Talukder, A. H., Mandal, M., Yang, Z., Bagheri-Yarmand, R., Sahin, A., Hortobagyi, G., Adam, L., Barnes, C. J., Vadlamudi, R. K. A naturally occurring MTA1 variant sequesters oestrogen receptor-alpha in the cytoplasm. Nature 418: 654-657, 2002. [PubMed: 12167865, related citations] [Full Text]

  54. Kun, Y., How, L. C., Hoon, T. P., Bajic, V. B., Lam, T. S., Aggarwal, A., Sze, H. G., Bok, W. S., Yin, W. C., Tan, P. Classifying the estrogen receptor status of breast cancers by expression profiles reveals a poor prognosis subpopulation exhibiting high expression of the ERBB2 receptor. Hum. Molec. Genet. 12: 3245-3258, 2003. [PubMed: 14570715, related citations] [Full Text]

  55. Lawson, J. S., Field, A. S., Champion, S., Tran, D., Ishikura, H., Trichopoulos, D. Low oestrogen receptor alpha expression in normal breast tissue underlies low breast cancer incidence in Japan. (Letter) Lancet 354: 1787-1788, 1999. [PubMed: 10577642, related citations] [Full Text]

  56. Lee, H., Kim, D.-W., Remedios, R., Anthony, T. E., Chang, A., Madisen, L., Zeng, H., Anderson, D. J. Scalable control of mounting and attack by Esr1+ neurons in the ventromedial hypothalamus. Nature 509: 627-632, 2014. [PubMed: 24739975, images, related citations] [Full Text]

  57. Lee, K., Jessop, H., Suswillo, R., Zaman, G., Lanyon, L. Bone adaptation requires oestrogen receptor-alpha. Nature 424: 389 only, 2003. [PubMed: 12879058, related citations] [Full Text]

  58. Lehtimaki, T., Kunnas, T. A., Mattila, K. M., Perola, M., Penttila, A., Koivula, T., Karhunen, P. J. Coronary artery wall atherosclerosis in relation to the estrogen receptor 1 gene polymorphism: an autopsy study. J. Molec. Med. 80: 176-180, 2002. [PubMed: 11894143, related citations] [Full Text]

  59. Li, L., Haynes, M. P., Bender, J. R. Plasma membrane localization and function of the estrogen receptor alpha variant (ER46) in human endothelial cells. Proc. Nat. Acad. Sci. 100: 4807-4812, 2003. [PubMed: 12682286, images, related citations] [Full Text]

  60. Liu, Z.-H., Cheng, Z.-H., Gong, R.-J., Liu, H., Liu, D., Li, L.-S. Sex differences in estrogen receptor gene polymorphism and its association with lupus nephritis in Chinese. Nephron 90: 174-180, 2002. [PubMed: 11818702, related citations] [Full Text]

  61. Lonard, D. M., Nawaz, Z., Smith, C. L., O'Malley, B. W. The 26S proteasome is required for estrogen receptor-alpha and coactivator turnover and for efficient estrogen receptor-alpha transactivation. Molec. Cell 5: 939-948, 2000. [PubMed: 10911988, related citations] [Full Text]

  62. Lorentzon, M., Lorentzon, R., Backstrom, T., Nordstrom, P. Estrogen receptor gene polymorphism, but not estradiol levels, is related to bone density in healthy adolescent boys: a cross-sectional and longitudinal study. J. Clin. Endocr. Metab. 84: 4597-4601, 1999. [PubMed: 10599725, related citations] [Full Text]

  63. Mader, S., Kumar, V., de Verneuil, H., Chambon, P. Three amino acids of the oestrogen receptor are essential to its ability to distinguish an oestrogen from a glucocorticoid-responsive element. Nature 338: 271-274, 1989. [PubMed: 2922054, related citations] [Full Text]

  64. Mattila, K. J., Nieminen, M. S., Valtonen, V. V., Rasi, V. P., Kesaniemi, Y. A., Syrjala, S. L., Jungell, P. S., Isoluoma, M., Hietaniemi K., Jokinen, M. J., Huttunen, J. K. Association between dental health and acute myocardial infarction. Brit. Med. J. 298: 779-782, 1989. [PubMed: 2496855, related citations] [Full Text]

  65. McGuire, W. L., Chamness, G. C., Fuqua, S. A. W. Estrogen receptor variants in clinical breast cancer. Molec. Endocr. 5: 1571-1577, 1991. [PubMed: 1779964, related citations] [Full Text]

  66. McGuire, W. L., Chamness, G. C., Fuqua, S. A. W. Abnormal estrogen receptor in clinical breast cancer. J. Steroid Biochem. Molec. Biol. 43: 243-247, 1992. [PubMed: 1525064, related citations] [Full Text]

  67. 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. Molec. Endocr. 10: 1519-1526, 1996. [PubMed: 8961262, related citations] [Full Text]

  68. Menasce, L. P., White, G. R. M., Harrison, C. J., Boyle, J. M. Localization of the estrogen receptor locus (ESR) to chromosome 6q25.1 by FISH and a simple post-FISH banding technique. Genomics 17: 263-265, 1993. [PubMed: 8406468, related citations] [Full Text]

  69. Metivier, R., Penot, G., Hubner, M. R., Reid, G., Brand, H., Kos, M., Gannon, F. Estrogen receptor-alpha directs ordered, cyclical, and combinatorial recruitment of cofactors on a natural target promoter. Cell 115: 751-763, 2003. [PubMed: 14675539, related citations] [Full Text]

  70. Metzger, D., White, J. H., Chambon, P. The human oestrogen receptor functions in yeast. Nature 334: 31-36, 1988. [PubMed: 3290686, related citations] [Full Text]

  71. Mohammed, H., Russell, I. A., Stark, R., Rueda, O. M., Hickey, T. E., Tarulli, G. A., Serandour, A. A., Birrell, S. N., Bruna, A., Saadi, A., Menon, S., Hadfield, J., and 12 others. Progesterone receptor modulates ER-alpha action in breast cancer. Nature 523: 313-317, 2015. Note: Erratum: Nature 523: 144 only, 2015. [PubMed: 26153859, images, related citations] [Full Text]

  72. Murphy, L. C., Wang, M., Coutt, A., Dotzlaw, H. Novel mutations in the estrogen receptor messenger RNA in human breast cancers. J. Clin. Endocr. Metab. 81: 1420-1427, 1996. [PubMed: 8636345, related citations] [Full Text]

  73. Musatov, S., Chen, W., Pfaff, D. W., Mobbs, C. V., Yang, X.-J., Clegg, D. J., Kaplitt, M. G., Ogawa, S. Silencing of estrogen receptor-alpha in the ventromedial nucleus of hypothalamus leads to metabolic syndrome. Proc. Nat. Acad. Sci. 104: 2501-2506, 2007. [PubMed: 17284595, images, related citations] [Full Text]

  74. Nakamura, T., Imai, Y., Matsumoto, T., Sato, S., Takeuchi, K., Igarashi, K., Harada, Y., Azuma, Y., Krust, A., Yamamoto, Y., Nishina, H., Takeda, S., Takayanagi, H., Metzger, D., Kanno, J., Takaoka, K., Martin, T. J., Chambon, P., Kato, S. Estrogen prevents bone loss via estrogen receptor alpha and induction of Fas ligand in osteoclasts. Cell 130: 811-823, 2007. [PubMed: 17803905, related citations] [Full Text]

  75. Nelson, E. R., Wardell, S. E., Jasper, J. S., Park, S., Suchindran, S., Howe, M. K., Carver, N. J., Pillai, R. V., Sullivan, P. M., Sondhi, V., Umetani, M., Geradts, J., McDonnell, D. P. 27-hydroxycholesterol links hypercholesterolemia and breast cancer pathophysiology. Science 342: 1094-1098, 2013. [PubMed: 24288332, images, related citations] [Full Text]

  76. Ohtake, F., Takeyama, K., Matsumoto, T., Kitagawa, H., Yamamoto, Y., Nohara, K., Tohyama, C., Krust, A., Mimura, J., Chambon, P., Yanagisawa, J., Fujii-Kuriyama, Y., Kato, S. Modulation of oestrogen receptor signalling by association with the activated dioxin receptor. Nature 423: 545-550, 2003. [PubMed: 12774124, related citations] [Full Text]

  77. Park, C. J., Zhao, Z., Glidewell-Kenney, C., Lazic, M., Chambon, P., Krust, A., Weiss, J., Clegg, D. J., Dunaif, A., Jameson, J. L., Levine, J. E. Genetic rescue of nonclassical ER-alpha signaling normalizes energy balance in obese Er-alpha-null mutant mice. J. Clin. Invest. 121: 604-612, 2011. [PubMed: 21245576, images, related citations] [Full Text]

  78. Pedram, A., Razandi, M., Deschenes, R. J., Levin, E. R. DHHC-7 and -21 are palmitoylacyltransferases for sex steroid receptors. Molec. Biol. Cell 23: 188-199, 2012. [PubMed: 22031296, images, related citations] [Full Text]

  79. Pelletier, G., El-Alfy, M. Immunocytochemical localization of estrogen receptors alpha and beta in the human reproductive organs. J. Clin. Endocr. Metab. 85: 4835-4840, 2000. [PubMed: 11134151, related citations] [Full Text]

  80. Perillo, B., Ombra, M. N., Bertoni, A., Cuozzo, C., Sacchetti, S., Sasso, A., Chiariotti, L., Malorni, A., Abbondanza, C., Avvedimento, E. V. DNA oxidation as triggered by H3K9me2 demethylation drives estrogen-induced gene expression. Science 319: 202-206, 2008. [PubMed: 18187655, related citations] [Full Text]

  81. Ponglikitmongkol, M., Green, S., Chambon, P. Genomic organization of the human oestrogen receptor gene. EMBO J. 7: 3385-3388, 1988. [PubMed: 3145193, related citations] [Full Text]

  82. Quaynor, S. D., Stradtman, E. W., Jr., Kim, H.-G., Shen, Y., Chorich, L. P., Schreihofer, D. A., Layman, L. C. Delayed puberty and estrogen resistance in a woman with estrogen receptor alpha variant. New Eng. J. Med. 369: 164-171, 2013. [PubMed: 23841731, images, related citations] [Full Text]

  83. Reese, J. C., Katzenellenbogen, B. S. Mutagenesis of cysteines in the hormone binding domain of the human estrogen receptor: alterations in binding and transcriptional activation by covalently and reversibly attaching ligands. J. Biol. Chem. 266: 10880-10887, 1991. [PubMed: 2040605, related citations]

  84. Reese, J. C., Katzenellenbogen, B. S. Characterization of a temperature-sensitive mutation in the hormone binding domain of the human estrogen receptor: studies in cell extracts and intact cells and their implications for hormone-dependent transcriptional activation. J. Biol. Chem. 267: 9868-9873, 1992. [PubMed: 1577818, related citations]

  85. Reid, G., Hubner, M. R., Metivier, R., Brand, H., Denger, S., Manu, D., Beaudouin, J., Ellenberg, J., Gannon, F. Cyclic, proteasome-mediated turnover of unliganded and liganded ER-alpha on responsive promoters is an integral feature of estrogen signaling. Molec. Cell 11: 695-707, 2003. [PubMed: 12667452, related citations] [Full Text]

  86. Reis-Filho, J. S., Drury, S., Lambros, M. B., Marchio, C., Johnson, N., Natrajan, R., Salter, J., Levey, P., Fletcher, O., Peto, J., Ashworth, A., Dowsett, M. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 809-810, 2008. [PubMed: 18583966, related citations] [Full Text]

  87. Robinson, D. R., Wu, Y.-M., Vats, P., Su, F., Lonigro, R. J., Cao, X., Kalyana-Sundaram, S., Wang, R., Ning, Y., Hodges, L., Gursky, A., Siddiqui, J., and 13 others. Activating ESR1 mutations in hormone-resistant metastatic breast cancer. Nature Genet. 45: 1446-1451, 2013. [PubMed: 24185510, images, related citations] [Full Text]

  88. Schuit, S. C. E., Oei, H.-H. S., Witteman, J. C. M., Geurts van Kessel, C. H., van Meurs, J. B. J., Nijhuis, R. L., van Leeuwen, J. P. T. M., de Jong, F. H., Zillikens, M. C., Hofman, A., Pols, H. A. P., Uitterlinden, A. G. Estrogen receptor alpha gene polymorphisms and risk of myocardial infarction. JAMA 291: 2969-2977, 2004. [PubMed: 15213208, related citations] [Full Text]

  89. Scott, G. K., Kushner, P., Vigne, J.-L., Benz, C. C. Truncated forms of DNA-binding estrogen receptors in human breast cancer. J. Clin. Invest. 88: 700-706, 1991. [PubMed: 1864980, related citations] [Full Text]

  90. Shang, Y., Brown, M. Molecular determinants for the tissue specificity of SERMs. Science 295: 2465-2468, 2002. [PubMed: 11923541, related citations] [Full Text]

  91. Shearman, A. M., Cupples, L. A., Demissie, S., Peter, I., Schmid, C. H., Karas, R. H., Mendelsohn, M. E., Housman, D. E., Levy, D. Association between estrogen receptor alpha gene variation and cardiovascular disease. JAMA 290: 2263-2270, 2003. Note: Erratum: JAMA 291: 186 only, 2004. [PubMed: 14600184, related citations] [Full Text]

  92. Shearman, A. M., Demissie, S., Cupples, L. A., Peter, I., Schmid, C. H., Ordovas, J. M., Mendelsohn, M. E., Housman, D. E. Tobacco smoking, estrogen receptor alpha gene variation and small low density lipoprotein level. Hum. Molec. Genet. 14: 2405-2413, 2005. [PubMed: 16014638, related citations] [Full Text]

  93. 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 95: 927-937, 1998. [PubMed: 9875847, related citations] [Full Text]

  94. Shim, G.-J., Kis, L. L., Warner, M., Gustafsson, J.-A. Autoimmune glomerulonephritis with spontaneous formation of splenic germinal centers in mice lacking the estrogen receptor alpha gene. Proc. Nat. Acad. Sci. 101: 1720-1724, 2004. Note: Erratum: Proc. Nat. Acad. Sci. 103: 8298 only, 2006. [PubMed: 14745006, images, related citations] [Full Text]

  95. Shupnik, M. A., Pitt, L. K., Soh, A. Y., Anderson, A., Lopes, M. B., Laws, E. R., Jr. Selective expression of estrogen receptor alpha and beta isoforms in human pituitary tumors. J. Clin. Endocr. Metab. 83: 3965-3972, 1998. [PubMed: 9814476, related citations] [Full Text]

  96. Simoncini, T., Hafezi-Moghadam, A., Brazil, D. P., Ley, K., Chin, W. W., Liao, J. K. Interaction of oestrogen receptor with the regulatory subunit of phosphatidylinositol-3-OH kinase. Nature 407: 538-541, 2000. [PubMed: 11029009, images, related citations] [Full Text]

  97. Sluyser, M., Mester, J. Oncogenes homologous to steroid receptors? (Letter) Nature 315: 546 only, 1985. [PubMed: 2989691, related citations] [Full Text]

  98. Sluyser, M. Mutations in the estrogen receptor gene. Hum. Mutat. 6: 97-103, 1995. [PubMed: 7581411, related citations] [Full Text]

  99. 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. New Eng. J. Med. 331: 1056-1061, 1994. Note: Erratum: New Eng. J. Med. 332: 131 only, 1995. [PubMed: 8090165, related citations] [Full Text]

  100. Sowers, M., Jannausch, M. L., Liang, W., Willing, M. Estrogen receptor genotypes and their association with the 10-year changes in bone mineral density and osteocalcin concentrations. J. Clin. Endocr. Metab. 89: 733-739, 2004. [PubMed: 14764789, related citations] [Full Text]

  101. Stoica, A., Katzenellenbogen, B. S., Martin, M. B. Activation of estrogen receptor-alpha by the heavy metal cadmium. Molec. Endocr. 14: 545-553, 2000. [PubMed: 10770491, related citations] [Full Text]

  102. 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., Bagger, Y., Gulcher, J. R., Eisman, J. A., Christiansen, C., Sigurdsson, G., Kong, A., Thorsteinsdottir, U., Stefansson, K. Multiple genetic loci for bone mineral density and fractures. New Eng. J. Med. 358: 2355-2365, 2008. [PubMed: 18445777, related citations] [Full Text]

  103. Taguchi, A., Kobayashi, J., Suei, Y., Ohtsuka, M., Tanimoto, K., Sanada, M., Ohama, K. Relationship between estrogen-receptor genotype and tooth loss in postmenopausal women. (Letter) JAMA 286: 2234-2235, 2001. [PubMed: 11710886, related citations] [Full Text]

  104. Taguchi, A. Estrogen receptor alpha polymorphisms and the risk of myocardial infarction. (Letter) JAMA 292: 1683 only, 2004. [PubMed: 15479929, related citations] [Full Text]

  105. Takeyama, J., Suzuki, T., Inoue, S., Kaneko, C., Nagura, H., Harada, N., Sasano, H. Expression and cellular localization of estrogen receptors alpha and beta in the human fetus. J. Clin. Endocr. Metab. 86: 2258-2262, 2001. [PubMed: 11344236, related citations] [Full Text]

  106. Tobias, J. H., Steer, C. D., Vilarino-Guell, C., Brown, M. A. Estrogen receptor alpha regulates area-adjusted bone mineral content in late pubertal girls. J. Clin. Endocr. Metab. 92: 641-647, 2007. [PubMed: 17105837, related citations] [Full Text]

  107. Toy, W., Shen, Y., Won, H., Green, B., Sakr, R. A., Will, M., Li, Z., Gala, K., Fanning, S., King, T. A., Hudis, C., Chen, D., Taran, T., Hortobagyi, G., Greene, G., Berger, M., Baselga, J., Chandarlapaty, S. ESR1 ligand-binding domain mutations in hormone-resistant breast cancer. Nature Genet. 45: 1439-1445, 2013. [PubMed: 24185512, images, related citations] [Full Text]

  108. van Meurs, J. B. J., Schuit, S. C. E., Weel, A. E. A. M., van der Klift, M., Bergink, A. P., Arp, P. P., Colin, E. M., Fang, Y., Hofman, A., van Duijn, C. M., van Leeuwen, J. P. T. M., Pols, H. A. P., Uitterlinden, A. G. Association of 5-prime estrogen receptor alpha gene polymorphisms with bone mineral density, vertebral bone area and fracture risk. Hum. Molec. Genet. 12: 1745-1754, 2003. [PubMed: 12837697, related citations] [Full Text]

  109. Vincent-Salomon, A., Raynal, V., Lucchesi, C., Gruel, N., Delattre, O. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 809 only, 2008. [PubMed: 18583967, related citations] [Full Text]

  110. Wada-Hiraike, O., Yano, T., Nei, T., Matsumoto, Y., Nagasaka, K., Takizawa, S., Oishi, H., Arimoto, T., Nakagawa, S., Yasugi, T., Kato, S., Taketani, Y. The DNA mismatch repair gene hMSH2 is a potent coactivator of oestrogen receptor-alpha. Brit. J. Cancer 92: 2286-2291, 2005. [PubMed: 15886699, images, related citations] [Full Text]

  111. Walter, P., Green, S., Greene, G., Krust, A., Bornert, J.-M., Jeltsch, J.-M., Staub, A., Jensen, E., Scrace, G., Waterfield, M., Chambon, P. Cloning of the human estrogen receptor cDNA. Proc. Nat. Acad. Sci. 82: 7889-7893, 1985. [PubMed: 3865204, related citations] [Full Text]

  112. Wang, X., Yang, Z., Zhang, H., Ding, L., Li, X., Zhu, C., Zheng, Y., Ye, Q. The estrogen receptor-interacting protein HPIP increases estrogen-responsive gene expression through activation of MAPK and AKT. Biochim. Biophys. Acta 1783: 1220-1228, 2008. [PubMed: 18302941, related citations] [Full Text]

  113. Wei, X., Xu, H., Kufe, D. MUC1 oncoprotein stabilizes and activates estrogen receptor alpha. Molec. Cell 21: 295-305, 2006. [PubMed: 16427018, related citations] [Full Text]

  114. Weis, K. E., Ekena, K., Thomas, J. A., Lazennec, G., Katzenellenbogen, B. S. Constitutively active human estrogen receptors containing amino acid substitutions for tyrosine 537 in the receptor protein. Molec. Endocr. 10: 1388-1398, 1996. [PubMed: 8923465, related citations] [Full Text]

  115. Ye, Q., Chung, L. W. K., Li, S., Zhau, H. E. Identification of a novel FAS/ER-alpha fusion transcript expressed in human cancer cells. Biochim. Biophys. Acta 1493: 373-377, 2000. [PubMed: 11018265, related citations] [Full Text]

  116. Zhao, C., Koide, A., Abrams, J., Deighton-Collins, S., Martinez, A., Schwartz, J. A., Koide, S., Skafar, D. F. Mutation of Leu-536 in human estrogen receptor-alpha alters the coupling between ligand binding, transcription activation, and receptor conformation. J. Biol. Chem. 278: 27278-27286, 2003. [PubMed: 12736255, related citations] [Full Text]

  117. Zuppan, P., Hall, J. M., Lee, M. K., Ponglikitmongkol, M., King, M.-C. Possible linkage of the estrogen receptor gene to breast cancer in a family with late-onset disease. Am. J. Hum. Genet. 48: 1065-1068, 1991. [PubMed: 2035527, related citations]

  118. Zuppan, P. J., Hall, J. M., Ponglikitmongkol, M., Spielman, R., King, M. C. Polymorphisms at the estrogen receptor (ESR) locus and linkage relationships on chromosome 6q. (Abstract) Cytogenet. Cell Genet. 51: 1116 only, 1989.


Bao Lige - updated : 03/18/2020
Bao Lige - updated : 01/31/2020
Marla J. F. O'Neill - updated : 12/03/2018
Ada Hamosh - updated : 11/30/2015
Matthew B. Gross - updated : 11/3/2015
Ada Hamosh - updated : 8/26/2014
Ada Hamosh - updated : 1/9/2014
Ada Hamosh - updated : 12/19/2013
Marla J. F. O'Neill - updated : 11/7/2013
Marla J. F. O'Neill - updated : 8/8/2013
Patricia A. Hartz - updated : 4/19/2013
Patricia A. Hartz - updated : 4/27/2012
Patricia A. Hartz - updated : 3/9/2011
Ada Hamosh - updated : 12/29/2009
John A. Phillips, III - updated : 5/7/2009
Patricia A. Hartz - updated : 3/26/2009
Ada Hamosh - updated : 1/6/2009
Ada Hamosh - updated : 8/6/2008
Ada Hamosh - updated : 6/10/2008
Patricia A. Hartz - updated : 5/29/2008
Ada Hamosh - updated : 3/25/2008
Patricia A. Hartz - updated : 2/7/2008
Marla J. F. O'Neill - updated : 1/21/2008
John A. Phillips, III - updated : 12/19/2007
Jane Kelly - updated : 11/27/2007
Victor A. McKusick - updated : 5/24/2007
Marla J. F. O'Neill - updated : 4/19/2007
Patricia A. Hartz - updated : 6/2/2006
Patricia A. Hartz - updated : 2/9/2006
Marla J. F. O'Neill - updated : 1/20/2006
George E. Tiller - updated : 1/10/2006
Patricia A. Hartz - updated : 11/9/2005
Marla J. F. O'Neill - updated : 8/30/2005
John A. Phillips, III - updated : 8/1/2005
Marla J. F. O'Neill - updated : 5/19/2005
Marla J. F. O'Neill - updated : 5/10/2005
George E. Tiller - updated : 5/6/2005
John A. Phillips, III - updated : 4/11/2005
Victor A. McKusick - updated : 1/25/2005
Victor A. McKusick - updated : 12/29/2004
John A. Phillips, III - updated : 10/14/2004
Victor A. McKusick - updated : 5/12/2004
Victor A. McKusick - updated : 12/8/2003
Patricia A. Hartz - updated : 10/10/2003
Ada Hamosh - updated : 8/4/2003
Ada Hamosh - updated : 7/24/2003
Victor A. McKusick - updated : 6/6/2003
Ada Hamosh - updated : 5/29/2003
Stylianos E. Antonarakis - updated : 4/15/2003
Patricia A. Hartz - updated : 1/30/2003
Ada Hamosh - updated : 10/18/2002
Ada Hamosh - updated : 9/20/2002
John A. Phillips, III - updated : 7/29/2002
John A. Phillips, III - updated : 7/26/2002
Victor A. McKusick - updated : 4/5/2002
Ada Hamosh - updated : 4/2/2002
John A. Phillips, III - updated : 10/9/2001
John A. Phillips, III - updated : 9/21/2001
John A. Phillips, III - updated : 7/9/2001
Paul J. Converse - updated : 2/1/2001
Victor A. McKusick - updated : 11/30/2000
John A. Phillips, III - updated : 11/13/2000
George E. Tiller - updated : 10/27/2000
Ada Hamosh - updated : 10/18/2000
John A. Phillips, III - updated : 9/29/2000
Stylianos E. Antonarakis - updated : 8/2/2000
Jane Kelly - updated : 6/22/2000
Wilson H. Y. Lo - updated : 3/21/2000
Ada Hamosh - updated : 2/8/2000
John A. Phillips, III - updated : 9/29/1999
Ada Hamosh - updated : 5/20/1999
John A. Phillips, III - updated : 4/15/1999
John A. Phillips, III - updated : 3/18/1999
Stylianos E. Antonarakis - updated : 2/4/1999
Paul Brennan - updated : 1/16/1998
Victor A. McKusick - updated : 6/23/1997
John A. Phillips, III - updated : 3/20/1997
John A. Phillips, III - updated : 2/25/1997
John A. Phillips, III - updated : 1/8/1997
Creation Date:
Victor A. McKusick : 6/4/1986
alopez : 02/12/2021
mgross : 03/18/2020
mgross : 01/31/2020
carol : 10/01/2019
carol : 09/27/2019
carol : 09/18/2019
carol : 12/03/2018
carol : 11/14/2017
carol : 11/13/2017
alopez : 11/30/2015
alopez : 11/30/2015
mgross : 11/3/2015
mgross : 11/3/2015
carol : 8/14/2015
alopez : 8/26/2014
alopez : 1/9/2014
alopez : 12/19/2013
alopez : 11/8/2013
alopez : 11/8/2013
mcolton : 11/7/2013
carol : 8/9/2013
carol : 8/9/2013
tpirozzi : 8/8/2013
tpirozzi : 8/8/2013
tpirozzi : 8/8/2013
mgross : 4/19/2013
carol : 2/26/2013
mgross : 2/5/2013
carol : 1/2/2013
terry : 10/10/2012
mgross : 4/27/2012
terry : 2/15/2012
terry : 1/13/2012
carol : 11/2/2011
ckniffin : 4/8/2011
mgross : 3/9/2011
terry : 3/9/2011
alopez : 3/1/2011
alopez : 3/1/2011
wwang : 7/1/2010
terry : 6/23/2010
alopez : 6/22/2010
alopez : 1/6/2010
terry : 12/29/2009
mgross : 12/10/2009
terry : 12/3/2009
terry : 6/3/2009
alopez : 5/7/2009
mgross : 3/26/2009
mgross : 3/26/2009
alopez : 3/11/2009
mgross : 2/4/2009
carol : 2/2/2009
wwang : 1/12/2009
alopez : 1/7/2009
terry : 1/6/2009
terry : 9/25/2008
alopez : 8/28/2008
terry : 8/6/2008
alopez : 6/16/2008
terry : 6/10/2008
mgross : 6/2/2008
terry : 5/29/2008
alopez : 3/25/2008
mgross : 2/8/2008
mgross : 2/8/2008
terry : 2/7/2008
carol : 1/21/2008
carol : 12/26/2007
carol : 12/19/2007
carol : 11/27/2007
carol : 11/27/2007
alopez : 6/6/2007
terry : 5/24/2007
wwang : 4/25/2007
wwang : 4/19/2007
wwang : 11/8/2006
mgross : 7/5/2006
terry : 6/27/2006
mgross : 6/8/2006
terry : 6/2/2006
mgross : 2/24/2006
terry : 2/9/2006
terry : 2/3/2006
wwang : 1/27/2006
wwang : 1/20/2006
wwang : 1/20/2006
terry : 1/10/2006
wwang : 11/28/2005
terry : 11/9/2005
wwang : 8/30/2005
alopez : 8/1/2005
wwang : 5/19/2005
wwang : 5/10/2005
tkritzer : 5/6/2005
alopez : 4/11/2005
tkritzer : 1/31/2005
terry : 1/25/2005
tkritzer : 12/30/2004
terry : 12/29/2004
alopez : 10/14/2004
alopez : 7/27/2004
tkritzer : 5/18/2004
terry : 5/12/2004
tkritzer : 12/9/2003
terry : 12/8/2003
ckniffin : 10/15/2003
mgross : 10/10/2003
mgross : 10/10/2003
carol : 8/28/2003
alopez : 8/5/2003
terry : 8/4/2003
tkritzer : 7/25/2003
terry : 7/24/2003
carol : 6/19/2003
tkritzer : 6/17/2003
terry : 6/6/2003
mgross : 5/30/2003
terry : 5/29/2003
mgross : 4/15/2003
mgross : 2/4/2003
terry : 1/30/2003
mgross : 10/18/2002
mgross : 10/18/2002
terry : 10/18/2002
alopez : 9/20/2002
carol : 9/18/2002
tkritzer : 9/13/2002
tkritzer : 9/6/2002
tkritzer : 7/29/2002
tkritzer : 7/26/2002
cwells : 4/15/2002
cwells : 4/5/2002
cwells : 4/5/2002
cwells : 4/3/2002
terry : 4/2/2002
cwells : 10/9/2001
cwells : 10/9/2001
cwells : 10/5/2001
cwells : 9/26/2001
cwells : 9/21/2001
alopez : 9/21/2001
carol : 9/13/2001
alopez : 8/13/2001
alopez : 8/13/2001
alopez : 8/13/2001
alopez : 7/9/2001
cwells : 2/6/2001
cwells : 2/1/2001
mgross : 12/4/2000
terry : 11/30/2000
terry : 11/13/2000
mcapotos : 11/1/2000
mcapotos : 10/27/2000
alopez : 10/18/2000
mgross : 10/3/2000
terry : 9/29/2000
mcapotos : 8/10/2000
mgross : 8/2/2000
alopez : 6/22/2000
carol : 3/21/2000
carol : 3/21/2000
alopez : 2/8/2000
mgross : 9/29/1999
alopez : 5/20/1999
terry : 5/20/1999
mgross : 4/15/1999
mgross : 3/23/1999
mgross : 3/18/1999
mgross : 3/17/1999
carol : 2/4/1999
dkim : 9/11/1998
alopez : 7/6/1998
terry : 6/30/1998
carol : 1/16/1998
terry : 8/7/1997
terry : 8/5/1997
mark : 6/26/1997
jenny : 6/23/1997
jenny : 5/28/1997
jenny : 5/27/1997
jenny : 3/20/1997
mark : 3/18/1997
jenny : 2/25/1997
carol : 11/12/1996
mark : 10/16/1996
terry : 10/31/1995
mark : 10/2/1995
carol : 1/5/1995
pfoster : 4/21/1994
warfield : 4/20/1994
carol : 10/26/1993

* 133430

ESTROGEN RECEPTOR 1; ESR1


Alternative titles; symbols

ESTROGEN RECEPTOR; ESR; ER
ESTROGEN RECEPTOR, ALPHA; ESRA


HGNC Approved Gene Symbol: ESR1

Cytogenetic location: 6q25.1-q25.2     Genomic coordinates (GRCh38): 6:151,656,672-152,129,619 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
6q25.1-q25.2 {Migraine, susceptibility to} 157300 Autosomal dominant 3
{Myocardial infarction, susceptibility to} 608446 3
Breast cancer, somatic 114480 3
Estrogen resistance 615363 Autosomal recessive 3

TEXT

Description

The estrogen receptor (ESR) is a ligand-activated transcription factor composed of several domains important for hormone binding, DNA binding, and activation of transcription. Alternative splicing results in several ESR1 mRNA transcripts, which differ primarily in their 5-prime untranslated regions. The translated receptors show less variability (Kos et al., 2001; Flouriot et al., 2000).


Cloning and Expression

Walter et al. (1985) cloned and Greene et al. (1986) sequenced a cDNA for the entire translated portion of the messenger RNA for the estrogen receptor of MCF-7 human breast cancer cells. Expression with production of a functional protein was accomplished in Chinese hamster ovary cells. The 1,785 nucleotides of the cDNA correspond to a polypeptide of 595 amino acids and a molecular weight of 66,200 (about that estimated from other studies of the estrogen receptor). Amino acid sequence comparisons showed considerable similarities between human ESR, human GCR (138040), and the putative v-erbA (190120) oncogene product. Both ESR and GCR exert their effects by binding directly to an intranuclear receptor molecule that is weakly associated with nuclear components in the absence of ligand. Binding of hormone to its receptor results in conversion of the receptor-steroid complex to a form that binds with high affinity to nuclear components. Green et al. (1986) also cloned and sequenced human estrogen receptor cDNA, using the breast cancer cell line MCF-7. They found extensive homology between ESR cDNA and the ERBA oncogene.

Flouriot et al. (2000) cloned a 46-kD isoform of ESR1 that lacks 173 N-terminal amino acids present in the major 66-kD isoform. These amino acids form a ligand-independent transactivation domain. The 46-kD isoform retains the DNA-binding domains and the hormone-binding domain, which includes the functional domain for hormone-inducible transcription activation.

Li et al. (2003) confirmed that the 46-kD endothelial cell protein (ESR46) is an N-terminal truncated product of full-length ESR-alpha (ESR66) that results from alternative splicing. ESR46 is expressed in the plasma membrane, cytosol, and nucleus of resting, estrogen-deprived cells. They found that ESR46 is localized and further dynamically targeted to the plasma membrane in a palmitoylation-dependent manner.


Biochemical Features

Crystal Structure

Shiau et al. (1998) reported the crystal structure of the human ESRA ligand-binding domain (LBD) bound to both the agonist diethylstilbestrol (DES) and a peptide derived from the nuclear receptor (NR) box II region of the coactivator GRIP1 (601993) as well as the crystal structure of the ESRA-LBD bound to the selective antagonist 4-hydroxytamoxifen (OHT). In the DES-LBD-peptide complex, the peptide binds as a short alpha helix to a hydrophobic groove on the surface of the LBD. In the OHT-LBD complex, helix 12 occludes the coactivator recognition groove by mimicking the interactions of the NR box peptide with the LBD.


Gene Structure

Ponglikitmongkol et al. (1988) showed that the human ESR gene is more than 140 kb long. It contains 8 exons, and the position of its introns has been highly conserved, being, for example, remarkably similar to those of one of the chicken thyroid hormone receptor genes.

Kos et al. (2001) reviewed the organization of the ESR1 gene. They described promoters used in the generation of ESR1 transcripts in human and other species and suggested a consistent nomenclature. The possible role of multiple promoters in the differential expression of ESR1 in tissues and during development was also discussed.


Mapping

Walter et al. (1985) determined that the human ESR gene maps to chromosome 6. By in situ hybridization, using a cDNA probe containing the coding sequence for the estrogen receptor, Gosden et al. (1986) assigned the gene to 6q24-q27. Zuppan et al. (1989) reported linkage data between the ESR locus and MYB (189990) and TCP10 (187020) on chromosome 6q. The estimated recombination fractions were 0.19 between ESR and MYB and 0.14 between ESR and TCP10, assuming equal male and female recombination fractions. To localize ESR more precisely, Menasce et al. (1993) isolated a YAC containing the gene and mapped it to 6q25.1 by fluorescence in situ hybridization (FISH) and a new simple method of post-FISH chromosome banding. Using a single interspecific backcross, Justice et al. (1990) demonstrated the genetic location of the Esr gene in relation to other loci on mouse chromosome 10.


Gene Function

Using a construct containing the human estrogen receptor cDNA with a yeast PGK promoter, Metzger et al. (1988) demonstrated that the human estrogen receptor can be expressed in yeast. Initiation of transcription in the system operated in a strictly hormone-dependent manner, documenting an extraordinarily high level of conservation of this activation mechanism.

Issa et al. (1994) reported that CpG island methylation, an epigenetic modification of DNA known to correlate closely with silencing of gene transcription, appears in the ESR gene in a subpopulation of cells that increases as a direct function of age in human colonic mucosa. They found that the same methylation change characterized almost all cells in the 45 colorectal tumors examined, including the earliest stages of tumor formation. ESR gene expression is diminished or absent in colorectal tumors, and introduction of an exogenous ESR gene in cultured colon carcinoma cells was found to result in marked growth suppression. Issa et al. (1994) concluded that methylation-associated inactivation of the ESR gene in aging colorectal mucosa may be one of the earliest events that predispose to sporadic colorectal tumorigenesis.

One mechanism suggested to play a role in the progression of human breast cancer from hormone dependence to independence is the expression or altered expression of mutant and/or variant forms of the estrogen receptor. Murphy et al. (1996) stated that 2 major types of variant ESR mRNA had been reported in human breast biopsy samples so far: truncated transcripts and exon-deleted transcripts. Murphy et al. (1996) provided data on a novel type of abnormal ESR mRNA. They found larger-than-wildtype ESR mRNA RT-PCR products in 9.4% of 212 human breast tumors analyzed. Cloning and sequencing of these larger RT-PCR products showed 3 different types: complete duplication of exon 6 in 7.5%; complete duplication of exons 3 and 4 in 1 tumor; and a 69-bp insertion between exons 5 and 6 in 3 tumors. While it is unknown if these novel ESR-like mRNAs are stably translated in vivo, any resulting protein would be structurally altered, possibly resulting in altered function.

Two isoforms of the human ESR, ESRA and ESR-beta (ESR2; 601663), occur, each with distinct tissue and cell patterns of expression. Additional ESR isoforms, generated by alternative mRNA splicing, have been defined in several tissues and are postulated to play a role in tumorigenesis or in modulating the estrogen response. By RT-PCR and hybridization blotting analysis, Shupnik et al. (1998) examined 71 human pituitary adenomas of varying phenotypes and 6 normal pituitary specimens for ESR mRNA isoforms. All 14 prolactinomas contained ESRA, and 5 of 14 contained ESRB mRNA. In comparison, 6 tumors that expressed prolactin (176760) and growth hormone (GH; 139250) expressed ESRB (4 of 6) more frequently than ESRA (3 of 6). ESRB mRNA was also found more frequently in null cell (8 of 24 ESRA and 14 of 24 ESRB) and gonadotrope (13 of 21 ESRA and 18 of 21 ESRB) tumors. Additionally, ESRB was found in 4 of 6 tumors that contained only GH, although ESRA was not observed in this tumor type. Expression of the 2 ESR isoforms within a tumor type was overlapping, but some tumors contained only 1 isoform. Expression of ESRA mRNA splice variants also varied with cell type. All normal pituitaries contained ESRA deletions of exon 4, 5, and 7, whereas only 2 of 6 samples contained the exon 2 deletion variant. Although the same ESRA mRNA variants were observed among the various tumor types, the proportion of specific splice variants expressed varied. The authors concluded that expression of the ESR isoforms, as well as of the mRNA splice variants, may influence the biologic properties of these tumors and affect their ability to respond to estrogen and antiestrogen therapies.

Vaults are large ribonucleoprotein particles composed largely of MVP (605088). Abbondanza et al. (1998) found that MVP coprecipitated with ER from nuclear extracts of MCF-7 human breast cancer cells and that ER associated with intact vaults. Mutation analysis showed that a central region of ER containing nuclear localization signals was involved in the interaction. A limited amount of ER molecules in the nuclear extract appeared to be associated with MVP. Physiologic concentrations of estradiol increased the amount of MVP present in MCF-7 nuclear extracts and coimmunoprecipitated with ER. The hormone-dependent interaction of vaults with ER was reproduced in vitro.

Lawson et al. (1999) studied ESR1 expression in normal breast tissue biopsies obtained by open surgical and core biopsy of benign lesions of the breast from 74 white Australian women and 92 Japanese women. Among Australians, the median of the percentage of ESR1-positive cells is about 12% for women younger than 50 years and 17% among women 50 years or older. Among Japanese, the overall median is about 9% and does not vary substantially with age. The authors concluded that their results are compatible with the hypothesis that the expression of estrogen receptors in normal breast tissue increases the risk of breast cancer and explains the 5-fold contrast of breast cancer incidence between white and Asian women. The difference in estrogen receptor expression as well as in breast cancer incidence between white and Japanese is more evident among older women.

Primary transcripts of the human ESR and progesterone receptor (PGR; 607311) genes undergo a number of alternative-splicing events that result in a range of variant mRNA isoforms in receptor-positive tissues. Despite in vitro demonstrations of a possible role for some of these isoforms in hormonal sensitivity, the clinical significance of this process is uncertain. By RT-PCR and Southern blot analysis, Balleine et al. (1999) documented the coexpression of variant ESR and PGR transcripts in a series of receptor-positive breast tumors. In 35 ESR-positive tumors, a common profile of variant ESR transcripts was present, with all tumors containing the exon 2-deleted and exon 7-deleted ESR variants, 94% containing the exon 4-deleted ESR variant, and 83% containing the exon 5-deleted ESR variant. In 25 of these cases, which were also PGR positive, the most highly expressed PGR variants, the exon 4-deleted PGR, exon 6-deleted PGR, and delta-4/2PGR, a transcript from which a 126-bp portion of PGR exon 4 was deleted, were detected in more than 90% of the cases. The alternatively spliced ESR isoforms were expressed at higher relative levels than the PGR isoforms, which had mean levels of expression less than 10% that of wildtype PGR. The most abundant isoform was the exon 7-deleted ESR, which was present at levels ranging from 29 to 83% of the wildtype. Balleine et al. (1999) concluded that the common profile of alternatively spliced ESR and PGR transcripts seen in breast tumors precludes its use as a discriminator of hormone responsiveness or other clinical end points. Furthermore, the low level of expression of the majority of variant isoforms called into question their potential for impacting significantly on receptor function.

Using transient transfection assays, Fan et al. (1999) demonstrated that BRCA1 inhibits signaling by the ligand-activated estrogen receptor ESR-alpha through the estrogen-responsive enhancer element and blocks the C-terminal transcriptional activation function AF2 of ESR-alpha. These results suggested that wildtype BRCA1 protein may function, in part, to suppress estrogen-dependent mammary epithelial proliferation by inhibiting ESR-alpha-mediated transcriptional pathways related to cell proliferation, and that loss of this ability may contribute to tumorigenesis.

Chaidarun and Alexander (1998) examined the effects of a human ESR1 isoform on estrogen-mediated gene activation in U2-OS osteosarcoma cells. ESR1-5, an ESR1 variant generated by an alternate splice event that omits exon 5 and alters the reading frame of the resulting mRNA, was coexpressed with normal ESR1 in several estrogen-responsive neoplastic tissues. ESR1-5 encodes the hormone-independent trans-activating function (AF-1), as well as the constitutive receptor dimerization and DNA-binding domains. The ESR1-5 protein is prematurely terminated and lacks the majority of the hormone-binding and activating function-2 (AF-2) domains. When ESR1-5 was cotransfected with normal ESR1, both basal and estrogen-stimulated reporter activation were increased approximately 500% over the levels observed when cells were transfected with ESR1 alone. Electromobility shift/supershift assays using nuclear extracts of U2-OS cells stably transfected with ESR1 and ESR1-5 confirmed the constitutive binding of ESR1-5 and ESR1 protein to the estrogen-response element (ERE) sequence independent of estrogen, and also showed an increase in ESR1-5/ESR1-ERE complexes with estrogen treatment. These data were considered to be consistent with the interactive effects of normal ESR1 and ESR1-5 on transcription from classic ERE gene promoters. Chaidarun and Alexander (1998) concluded that ESR1-5 acts as a dominant-positive receptor that increases both basal and estrogen-stimulated gene transactivation of normal ESR1.

Ye et al. (2000) investigated the expression of ESR1 in prostate cancer cell lines and unexpectedly found a FASN/ESR1 fusion transcript. Using semi-nested RT-PCR analysis of ESR1 and its variants, Ye et al. (2000) found that the N-terminal coding region of FASN containing domain 1 was fused to the C-terminal coding region of the ESR1 ligand binding domain. Nested RT-PCR also detected the fusion transcript in breast, cervical, and bladder cancer cell lines.

Esmaeli et al. (2000) found immunohistochemical evidence that estrogen receptors are present in the meibomian glands of the upper eyelid. Unlike sebaceous glands elsewhere on the skin, the meibomian glands lack androgen receptors. Esmaeli et al. (2000) suggested that these eyelid estrogen receptors may play a role in modulation of the tear film lipid layer. They concluded that estrogen receptor activity may be linked to meibomian gland dysfunction and dry eye syndrome.

ESR1 is downregulated in the presence of its cognate ligand, estradiol, through the ubiquitin proteasome pathway. Lonard et al. (2000) showed that ubiquitin proteasome function is required for ESR1 to serve as a transcriptional activator. Deletion of the last 61 amino acids of ESR1, including residues that form helix 12, abolished ligand-mediated downregulation of the receptor, as did point mutations in the ligand-binding domain that impaired coactivator binding. In addition, the authors found that coactivators also are subject to degradation by the 26S proteasome, but their intrinsic transcriptional activity is not affected. These data provided evidence that protein interactions with ESR1 coactivator binding surfaces are important for ligand-mediated receptor downregulation and suggested that receptor and coactivator turnover contributes to ESR1 transcriptional activity.

Simoncini et al. (2000) showed that the estrogen receptor isoform ESR-alpha binds in a ligand-dependent manner to the p85-alpha (171833) regulatory subunit of the phosphatidylinositol-3-hydroxy kinase (PI3K). Stimulation with estrogen increases ESR-alpha-associated PI3K activity, leading to the activation of protein kinase B/AKT (164730) and endothelial nitric oxide synthase (eNOS; 163729). Recruitment and activation of PI3K by ligand-bound ESR-alpha are independent of gene transcription, do not involve phosphotyrosine adaptor molecules or src homology domains of p85-alpha, and extend to other steroid hormone receptors. Mice treated with estrogen showed increased eNOS activity and decreased vascular leukocyte accumulation after ischemia and reperfusion injury. This vascular protective effect of estrogen was abolished in the presence of PI3K or eNOS inhibitors. Simoncini et al. (2000) concluded that their findings defined a physiologically important nonnuclear estrogen-signaling pathway involving the direct interaction of ESR-alpha with PI3K.

Pelletier and El-Alfy (2000) studied the immunocytochemical localization of ESRA and ESRB in human reproductive tissues. In the ovary, ESRB immunoreactivity was found in nuclei of granulosa cells of growing follicles at all stages from primary to mature follicles, interstitial gland, and germinal epithelium cells. Nuclear staining for ESRA occurred in thecal, interstitial gland, and germinal epithelium cells. In the uterus, strong ESRA immunoreactivity was detected in nuclei of epithelial, stromal, and muscle cells. Similar localization was obtained for ESRB, although the staining was much weaker. In the vagina, only ESRA could be detected; a nuclear reaction was observed in deep layers of the stratified epithelium as well as in stromal and muscle cells. In the mammary gland, both ESR subtypes were observed in epithelial and stromal cells. In the testis, ESRB was detected in the nuclei of Sertoli and Leydig cells, whereas ESRA immunoreactivity was only observed in Leydig cells, with no tubular labeling. In the efferent ducts, only ESRB could be detected, whereas neither ESRB nor ESRA could be found in the epididymis. In the prostate, ESRB nuclear immunolabeling was observed in both basal and secretory cells in alveoli as well as in stromal cells, whereas ESRA could not be detected. The authors concluded that there is a cell-specific localization for each of the ESR subtypes in the majority of the reproductive organs studied.

Chiang et al. (2000) examined the regulation of ESRA and ESRB expression by human chorionic gonadotropin (CG; see 118850) in human granulosa-luteal cells. CG treatment significantly attenuated the ESRA (45%) and ESRB (40%) mRNA levels. The CG-induced decrease in ESRA and ESRB expression was mimicked by 8-bromo-cAMP and forskolin treatment. Next, the effect of gonadotropin-releasing hormone (GNRH; 152760) on estrogen receptor expression was studied. Sixty-eight percent and 60% decreases in ESRA and ESRB mRNA levels, respectively, were observed after treatment with GNRH agonist. Pretreatment of the cells with a protein kinase C (PKC; see 176960) inhibitor completely reversed the GNRH agonist-induced downregulation of ESRA and ESRB expression, suggesting the involvement of PKC in GNRH signal transduction in granulosa-luteal cells. Chiang et al. (2000) observed a differential expression of ESRA and ESRB mRNA in granulosa-luteal cells in vitro. The authors concluded that the demonstration of CG- and GNRH agonist-induced downregulation of ESRA and ESRB gene expression suggests that CG and GNRH may contribute to the control of granulosa-luteal cell function. Furthermore, they inferred that the effects of CG and GNRH on ESRA and ESRB expression in granulosa-luteal cells are mediated in part by activation of PKA (see 176911) and PKC signaling pathways, respectively.

Bord et al. (2001) examined the expression of ESR1 and ESR2 in neonatal human rib bone. ESR1 and ESR2 immunoreactivity was seen in proliferative and prehypertrophic chondrocytes in the growth plate, with lower levels of expression in the late hypertrophic zone. Different patterns of expression of the 2 estrogen receptors were seen in bone. In cortical bone, intense staining for ESR1 was observed in osteoblasts and osteocytes adjacent to the periosteal-forming surface and in osteoclasts on the opposing resorbing surface. In cancellous bone, ESR2 was strongly expressed in both osteoblasts and osteocytes, whereas only low expression of ESR1 was seen in these areas. Nuclear and cytoplasmic staining for ESR2 was apparent in osteoclasts. The authors concluded that these observations demonstrate distinct patterns of expression for the 2 ER subtypes in developing human bone and indicate functions in both the growth plate and mineralized bone. In the latter, ESR1 is predominantly expressed in cortical bone, whereas ESR2 shows higher levels of expression in cancellous bone.

Takeyama et al. (2001) examined the expression and cellular localization of ESR1 and ESR2 in various human fetal tissues by semiquantitative RT-PCR (13 and 20 gestational weeks) and immunohistochemistry (13, 20, and 38 gestational weeks), respectively, to study the possible effects of estrogens on human fetal tissues during development. Relatively high levels of ESR2 expression were detected in various human fetal tissues, whereas those tissues expressing ESR2 had markedly lower levels of ESR1 expression. ESR2 mRNA expression was especially high in the adrenal gland. ESR2-immunoreactive protein was localized to the definitive zone, but not in the fetal zone, of the adrenal cortex. Although low levels of ESR2 mRNA were present in the brain, heart, lung, and kidney, ESR2 immunoreactivity was not detected in these tissues. These results suggested that the effects of estrogens in these tissues are predominantly mediated through ESR2. ESR2 immunoreactivity was detected in Sertoli cells and spermatogonia in the male reproductive tract and in germ cells in the fetal testis and epididymis. In the female reproductive tract, both ESR1 and ESR2 were immunopositive in the epithelium of the oviduct. These results demonstrated the possible sites for estrogenic action in the human fetus and suggested that the effects of estrogen via ESR2 may play important roles in human fetal development, especially in the definitive zone of the adrenal cortex and in the reproductive tissues of the developing fetus.

The therapeutic effectiveness of selective estrogen receptor modulators such as tamoxifen and raloxifene in breast cancer depends on their antiestrogenic activity. In the uterus, however, tamoxifen is estrogenic. Shang and Brown (2002) showed that both tamoxifen and raloxifene induce the recruitment of corepressors to target gene promoters in mammary cells. In endometrial cells, tamoxifen, but not raloxifene, acts like estrogen by stimulating the recruitment of coactivators to a subset of genes. The estrogen-like activity of tamoxifen in the uterus requires a high level of steroid receptor coactivator-1 (SRC1; 602691) expression. Thus, Shang and Brown (2002) concluded that cell type- and promoter-specific differences in coregulator recruitment determine the cellular response to selective estrogen receptor modulators.

Kumar et al. (2002) identified a naturally occurring short form of MTA1 (603526), which they called MTA1s, that contains a theretofore unknown sequence of 33 amino acids with an estrogen receptor-binding motif, leu-arg-ile-leu-leu (LRILL). MTA1s localizes in the cytoplasm, sequesters estrogen receptor in the cytoplasm, and enhances nongenomic responses of estrogen receptor. Deletion of the LRILL motif in MTA1s abolished its corepressor function and its interaction with estrogen receptor, and restored nuclear localization of estrogen receptor. Dysregulation of human epidermal growth factor receptor-2 (ERBB2; 164870) in breast cancer cells enhanced the expression of MTA1s and the cytoplasmic sequestration of estrogen receptor. Expression of MTA1s in breast cancer cells prevented ligand-induced nuclear translocation of estrogen receptor and stimulated malignant phenotypes. MTA1s expression is increased in human breast tumors with no or low nuclear estrogen receptor. The regulation of the cellular localization of estrogen receptor by MTA1s represents a mechanism for redirecting nuclear receptor signaling by nuclear exclusion.

Auboeuf et al. (2002) examined the impact of transcription mediated by steroid receptors, including progesterone and estrogen receptors, on RNA processing using reporter genes subject to alternative splicing driven by steroid-sensitive promoters. Steroid hormones affected the processing of pre-mRNA synthesized from steroid-sensitive promoters, but not from steroid unresponsive promoters, in a steroid receptor-dependent and receptor-selective manner. Several nuclear receptor coregulators showed differential splicing effects, suggesting that steroid hormone receptors may simultaneously control gene transcription activity and exon content of the product mRNA by recruiting coregulators involved in both processes.

Flouriot et al. (2000) showed that the 46-kD isoform of ESR1 can heterodimerize with the 66-kD isoform and can competitively inhibit the ligand-independent transactivation function of the larger receptor.

Kos et al. (2002) noted that the various ESR1 promoters are utilized in a tissue-specific manner to control the levels of mRNA variants in individual tissues. They looked specifically at the effect of upstream AUGs on the translation of major ESR1 variants. The 5-prime untranslated region of 1 mRNA species significantly suppressed the translation of a reporter gene in transient transfection assays. The 5-prime untranslated region of 3 others showed moderate negative effects. Toeprinting analysis revealed that leaky scanning occurs with mRNAs bearing the most inhibitory 5-prime untranslated regions. Kos et al. (2002) concluded that short upstream open reading frames in the 5-prime untranslated regions of ESR1 mRNAs can regulate receptor expression.

Reid et al. (2003) established the ubiquitination status and subnuclear distribution of ESR, its mobility, the kinetics of transcriptional activation, and the cyclic recruitment of E3 ligases and the 19S regulatory component of the proteasome. They demonstrated that proteasome-mediated degradation and ESR-mediated transactivation are linked and act to turn over ESR continuously on responsive promoters, and concluded that the cyclic turnover of ESR permits continuous responses to changes in the concentration of estradiol.

Zhao et al. (2003) noted that the crystal structure of the ER-alpha ligand-binding domain suggests that leu536 may be involved in hydrophobic interactions at the start of helix 12. They found that certain mutations of leu536 increased the ligand-independent activity of ER-alpha and reduced or eliminated the agonist activity of 17-beta-estradiol and 4-hydroxytamoxifen in a reporter assay. Zhao et al. (2003) concluded that leu536 is critical in coupling the binding of ligand to changes in the conformation and transcriptional activity of ER-alpha. A heterodimer of the dioxin receptor (AHR; 600253) and ARNT (126110), which are basic helix-loop-helix/PAS family transcription factors, mediates most of the toxic effects of dioxins.

Li et al. (2003) showed that the ESR46 isoform more efficiently modulates membrane-initiated estrogen actions, including activation of eNOS, than full-length ESR66. Conversely, ESR66 more efficiently mediates estrogen response element reporter-gene transactivation than ESR46.

Ohtake et al. (2003) demonstrated that the agonist-activated AHR/ARNT heterodimer directly associates with the estrogen receptors ESR-alpha and ESR-beta. They showed that this association results in the recruitment of unliganded estrogen receptor and the coactivator p300 (602700) to estrogen-responsive gene promoters, leading to activation of transcription and estrogenic effects. The function of liganded estrogen receptor was found to be attenuated. Estrogenic actions of AHR agonists were detected in wildtype ovariectomized mouse uteri, but were absent in Ahr -/- or Er-alpha -/- ovariectomized mice. Ohtake et al. (2003) concluded that their findings suggest a novel mechanism by which estrogen receptor-mediated estrogen signaling is modulated by a coregulatory-like function of activated AHR/ARNT, giving rise to adverse estrogen-related actions of dioxin-type environmental contaminants.

Garcia-Morales et al. (1994) found that cadmium is a potent stimulator of the estrogen receptor independent of estradiol. Stoica et al. (2000) found that cadmium activates ESR1 at concentrations as low as 10(-11) M. Cadmium was found to block the binding of estradiol to ESR1 in a noncompetitive manner, suggesting that the heavy metal interacts with the hormone-binding domain of the receptor. Stoica et al. (2000) showed that cadmium activates ESR1 through an interaction with the hormone-binding domain of the receptor. Transfection and binding assays with ESR1 mutants identified cys381, cys447, glu523, his524, and asp538 as possible interaction sites of cadmium with the hormone-binding domain of ER-alpha. Johnson et al. (2003) showed that cadmium has a potent estrogen-like activity in vivo. Female rats were ovariectomized on postnatal day 28 and allowed to recover for 3 weeks. Rats were then given a single intraperitoneal dose of cadmium (5 microgram/per kg body weight or approximately 27 nmol/kg). This exposure to cadmium increased uterine weight, promoted growth and development of mammary glands, and induced hormone-regulated genes in ovariectomized animals. In the uterus, the increase in wet weight was accompanied by proliferation of the endometrium and induction of progesterone receptor (Pgr; 607311), and complement component C3 (120700). In the mammary gland, cadmium promoted an increase in the formation of side branches and alveolar buds and the induction of casein, whey acidic protein, Pgr, and C3. In utero exposure to the metal also mimicked the effects of estrogens. Female offspring experienced an earlier onset of puberty and an increase in the epithelial area and the number of terminal end buds in the mammary gland.

Using chromatin immunoprecipitation assays, Metivier et al. (2003) identified protein complexes recruited by ESR1 to the pS2 (TFF1; 113710) promoter in a breast carcinoma cell line following estradiol treatment, and they determined the order in which the complexes were recruited.

By immunoprecipitation of human breast cancer cell lines and protein pull-down assays with in vitro translated proteins, Wada-Hiraike et al. (2005) demonstrated that ESR1 interacted with MSH2 (609309) in a ligand-dependent manner, whereas ESR2 (601663) and MSH2 interacted in a ligand-independent manner. Both receptors bound MSH2 through its MSH3 (600887)/MSH6 (600678)-interaction domain. In a transient expression assay, MSH2 potentiated the transactivation function of ligand-activated ESR1 but not ESR2. Wada-Hiraike et al. (2005) concluded that MSH2 may be a coactivator of ESR1-dependent gene expression.

Wei et al. (2006) found that the MUC1 (158340) C-terminal subunit associated with ESR1 and that the interaction was stimulated by 17-beta-estradiol in human breast carcinoma cell lines. MUC1 bound directly to the ESR1 DNA-binding domain and stabilized ESR1 by blocking its ubiquitination and degradation. Chromatin immunoprecipitation assays demonstrated that MUC1 associated with ESR1 complexes on estrogen-responsive promoters, enhanced ESR1 promoter occupancy, and increased recruitment of p160 (PELP1; 609455) coactivators SRC1 and GRIP1 (604597). MUC1 stimulated ESR1-mediated transcription and contributed to estradiol-mediate d growth and survival of breast cancer cells.

Aguirre et al. (2007) demonstrated that extracellular signal-regulated kinases (see ERKs, 176948) are not activated by stretching in osteocytic and osteoblastic cells in which both ESR1 and ESR2 have been knocked out or knocked down; this effect was partially reversed by transfection of either of the 2 human ESRs, and fully by transfection of both receptors. ERK activation in response to stretching was also recovered by transfecting the ligand-binding domain of either receptor or an ESR1 mutant that does not bind estrogens. Mechanoresponsiveness was restored by transfecting ESR1 targeted to the plasma membrane but not to the nucleus, and ESR1 mutants with impaired plasma membrane localization or binding to caveolin-1 (601047) failed to confer ERK activation in response to stretching. An ESR antagonist abrogated ERK activation as well as the antiapoptotic effect of mechanical stimulation. Aguirre et al. (2007) concluded that in addition to their role as ligand-dependent mediators of the effects of estrogens, the ESRs participate in the transduction of mechanical forces into prosurvival signaling in bone cells in a ligand-independent manner.

Perillo et al. (2008) analyzed how H3 histone methylation and demethylation control expression of estrogen-responsive genes and showed that a DNA-bound estrogen receptor directs transcription by participating in bending chromatin to contact the RNA polymerase II (see 180660) recruited to the promoter. This process is driven by receptor-targeted demethylation of H3K9 (see 602810) at both enhancer and promoter sites and is achieved by activation of resident LSD1 (AOF2; 609132) demethylase. Localized demethylation produces hydrogen peroxide, which modifies the surrounding DNA and recruits 8-oxoguanine-DNA glycosylase-1 (601982) and topoisomerase II-beta (126431), triggering chromatin and DNA conformational changes that are essential for estrogen-induced transcription. Perillo et al. (2008) concluded that their data showed a strategy that uses controlled DNA damage and repair to guide productive transcription.

Hurtado et al. (2008) showed that estrogen-estrogen receptor (ER) and tamoxifen-ER complexes directly repress ERBB2 (164870) transcription by means of a cis-regulatory element within the ERBB2 gene in human cell lines. Hurtado et al. (2008) implicated the paired box-2 gene product (PAX2; 167409) in a previously unrecognized role, as a crucial mediator of ER repression of ERBB2 by the anticancer drug tamoxifen. Hurtado et al. (2008) showed that PAX2 and the ER coactivator AIB1/SRC3 (601937) compete for binding and regulation of ERBB2 transcription, the outcome of which determines tamoxifen response in breast cancer cells. The repression of ERBB2 by ER-PAX2 links these 2 breast cancer subtypes and suggests that aggressive ERBB2-positive tumors can originate from ER-positive luminal tumors by circumventing this repressive mechanism. Hurtado et al. (2008) concluded that their data provided mechanistic insight into the molecular basis of endocrine resistance in breast cancer.

Wang et al. (2008) found that human HPIP (PBXIP1; 618819) interacted with both ER-alpha and ER-beta in mammalian cells. Overexpression and knockdown analyses revealed that HPIP interaction increased expression of ER-alpha target genes by enhancing phosphorylation of ER-alpha at ser167 by MAPK and AKT. Immunoprecipitation experiments demonstrated that ER-beta also interacted with ER-alpha, thereby decreasing binding of ER-alpha to HPIP and inhibiting expression of ER-alpha target genes.

Fullwood et al. (2009) described the development of a new strategy, which they called chromatin interaction analysis by paired-end tag sequencing (ChIA-PET), for the de novo detection of global chromatin interactions, with which they comprehensively mapped the chromatin interaction network bound by estrogen receptor-alpha (ER-alpha) in the human genome. Fullwood et al. (2009) found that most high-confidence remote ER-alpha-binding sites are anchored at gene promoters through long-range chromatin interactions, suggesting that ER-alpha functions by extensive chromatin looping to bring genes together for coordinated transcriptional regulation. Fullwood et al. (2009) proposed that chromatin interactions constitute a primary mechanism for regulating transcription in mammalian genomes.

Using breast cancer and other cancer cell lines, Hurtado et al. (2011) showed that FOXA1 (602294) mediated ER binding and function. Almost all ER-chromatin interactions and gene expression changes depended on FOXA1, and FOXA1 influenced genomewide chromatin accessibility. FOXA1 was also required for the inhibitory activity of tamoxifen against ER.

The classical sex steroid receptors, including ER-alpha, translocate from the plasma membrane to the nucleus following ligand binding and influence gene transcription. Sex steroid receptors also engage in cell signaling at the plasma membrane, and this function requires both ligand binding and receptor palmitoylation at a conserved cysteine residue. Palmitoylacyltransferases contain an asp-his-his-cys (DHHC) signature sequence. By expressing plasmids encoding each mouse DHHC protein in MCF-7 human breast cancer cells, Pedram et al. (2012) found that only Dhhc7 (614604) and Dhhc21 (614605) functioned in ER-alpha palmitoylation. Knockdown of endogenous DHHC7 and DHHC21 via small interfering RNA reduced ER-alpha localization at the plasma membrane and impaired intracellular signaling by 17-beta-estradiol via ERK (see 176948) and AKT kinases and cAMP generation. Nuclear ER-alpha localization and transcriptional activation were unaffected. Mutation of a cysteine in the C-terminal region of mouse ER-alpha abrogated palmitoylation by Dhhc7 and Dhhc21 and inhibited cell signaling from the plasma membrane.

Using a constitutively active mutant of the mouse nonreceptor tyrosine phosphatase Shp2 (PTPN11; 176876), He et al. (2012) found that Shp2 integrated leptin (LEP; 164160) and estrogen signaling in transgenic female mice. Transgenic females, but not males, were resistant to high-fat diet-induced obesity and liver steatosis via enhanced leptin and insulin sensitivity and downstream ERK activation. SHP2 and ESR1 interacted directly in MCF-7 cells and female mouse tissues, and the interaction was enhanced by estrogen stimulation. Ovariectomy of transgenic mice reversed their resistance to high-fat diet-induced obesity.

Hypercholesterolemia is a risk factor for ER-positive breast cancers and is associated with a decreased response of tumors to endocrine therapies. Nelson et al. (2013) showed that 27-hydroxycholesterol (27HC), a primary metabolite of cholesterol and an ER and liver X receptor (see LXRA, 602423) ligand, increases ER-dependent growth and LXR-dependent metastasis in mouse models of breast cancer. The effects of cholesterol on tumor pathology required its conversion to 27HC by the cytochrome P450 oxidase CYP27A1 (606530) and were attenuated by treatment with CYP27A1 inhibitors. In human breast cancer specimens, CYP27A1 expression levels correlated with tumor grade. In high-grade tumors, both tumor cells and tumor-associated macrophages exhibited high expression levels of the enzyme. Thus, Nelson et al. (2013) concluded that lowering circulating cholesterol levels or interfering with its conversion to 27HC may be a useful strategy to prevent and/or treat breast cancer.

Kim et al. (2013) showed that CAC1 (CACUL1; 618764) and ER-alpha interacted in a ligand-independent manner and colocalized to nucleus in transacted H1299 cells. Interaction with CAC1 repressed ER-alpha transcriptional activity, and CoRNR box-2 of CAC1 was required for both ER-alpha binding and repression. CAC1 interacted with ER-alpha coactivator LSD1 and inhibited its binding to the ER-alpha target promoter, thereby repressing LSD1-enhanced ER-alpha activity. In addition, CAC1 increased paclitaxel-induced cell death, as overexpression of CAC1 in paclitaxel-resistant MCF7 cells increased their sensitivity to paclitaxel by suppressing ER-alpha activation.

The ventrolateral subdivision of the murine ventromedial hypothalamus (VMHvl) contains neurons whose activity increases during male-male and male-female social encounters. Non-cell-type-specific optogenetic activation of this region elicited attack behavior, but not mounting. Lee et al. (2014) identified a subset of VMHvl neurons marked by Esr1, and investigated their role in male social behavior. Optogenetic manipulations indicated that Esr1-positive (but not Esr1-negative) neurons are sufficient to initiate attack, and that their activity is continuously required during ongoing agonistic behavior. Surprisingly, weaker optogenetic activation of these neurons promoted mounting behavior, rather than attack, toward both males and females, as well as sniffing and close investigation. Increasing photostimulation intensity could promote a transition from close investigation and mounting to attack, within a single social encounter. Importantly, time-resolved optogenetic inhibition experiments revealed requirements for Esr1-positive neurons in both the appetitive (investigative) and the consummatory phases of social interactions. Combined optogenetic activation and calcium imaging experiments in vitro, as well as c-Fos analysis in vivo, indicated that increasing photostimulation intensity increases both the number of active neurons and the average level of activity per neuron. Lee et al. (2014) concluded that these data suggested that Esr1-positive neurons in VMHvl control the progression of a social encounter from its appetitive through its consummatory phases, in a scalable manner that reflects the number or type of active neurons in the population.

Cho et al. (2015) used ribosome profiling and RNA sequencing to quantify changes in gene and protein expression in mouse hippocampus following contextual fear conditioning. They identified 3 phases of alterations: an initial wave of translational regulation at 5 to 10 minutes, a second wave of induction of immediate early genes at 10 to 30 minutes, and suppression of genes through decreased mRNA levels after 30 minutes, which continued through 4 hours. Ingenuity pathway analysis identified Esr1 as the most prominent upstream regulator of differentially expressed genes (DEGs) at 4 hours, and Esr1 was predicted to be inhibited. Half of the decreased DEGs at 4 hours were putative downstream targets of Esr1, and several other DEGs were downstream targets of Otx1 (600036), an Esr1 downstream target. Quantitative RT-PCR on hippocampal RNAs following administration of an Esr1 antagonist showed marked decreases in Otx1 and other putative Esr1 targets. Administration of an Esr1 agonist into mouse hippocampus after learning significantly impaired memory formation compared with controls in 2 hippocampus-dependent tasks. Cho et al. (2015) concluded that downregulation of ESR1 signaling is important for memory formation.

Mohammed et al. (2015) showed that the progesterone receptor (PR; 607311) is not merely an ER-alpha-induced gene target, but is also an ER-alpha-associated protein that modulates its behavior. In the presence of agonist ligands, PR associates with ER-alpha to direct ER-alpha chromatin binding events within breast cancer cells, resulting in a unique gene expression program that is associated with good clinical outcome. Progesterone inhibited estrogen-mediated growth of ER-alpha(+) cell line xenografts and primary ER-alpha(+) breast tumor explants, and had increased antiproliferative effects when coupled with an ER-alpha antagonist. Copy number loss of the PGR gene is a common feature in ER-alpha(+) breast cancers, explaining lower PR levels in a subset of cases. Mohammed et al. (2015) concluded that their findings indicated that PR functions as a molecular rheostat to control ER-alpha chromatin binding and transcriptional activity.


Molecular Genetics

Estrogen Resistance

In a 28-year-old man with estrogen resistance (ESTRR; 615363), Smith et al. (1994) performed single-strand conformation polymorphism analysis of the ESR1 gene and observed a variant banding pattern in exon 2. Direct sequencing revealed a homozygous nonsense mutation (R157X; 133430.0002).

McInerney et al. (1996) characterized a human ESR1 variant, V364E (133430.0003), which they demonstrated to be a strong dominant-negative inhibitor of wildtype ESR1.

In an 18-year-old woman with estrogen resistance, Quaynor et al. (2013) identified homozygosity for a missense mutation in the ESR1 gene (Q375H; 133430.0006). Clegg and Palmer (2013) noted that the Esr1 -/AA knockin mouse model (see ANIMAL MODEL) recapitulates the phenotype of the female patient studied by Quaynor et al. (2013), in that they have hypoplastic mammary glands, anovulation, and altered steroidogenesis, but have normal body weight, adiposity, locomotor activity, and glucose homeostasis. These findings supported the proposal by Quaynor et al. (2013) that nonclassical regulation of ESR1 is sufficient to protect against the obesity-metabolic syndrome phenotype associated with total loss of activity of ESR1, but is not sufficient to rescue the infertility phenotype.

In 3 sibs from a consanguineous Algerian family with estrogen resistance, Bernard et al. (2017) identified homozygosity for a missense mutation in the ESR1 gene (R394H; 133430.0007) that segregated with disease.

Association with Breast Cancer

Zuppan et al. (1991) reported a lod score of 1.85 for linkage of ESR at zero recombination with late-onset breast cancer in 1 extended family with 8 affected members. Simulation of this pedigree assuming independent inheritance of breast cancer and ESR genotypes led to a lod score equal to or greater than 1.85 only once in 2,000 replicates. Zuppan et al. (1991) suggested testing linkage in other breast cancer families of the late-onset type.

McGuire et al. (1991) used the screening techniques of chemical mismatch cleavage, single-strand conformation polymorphism (SSCP), and gel retardation to discover a number of estrogen receptor mRNA variants in tissues from breast cancers. They identified basepair insertions, transitions, and deletions as well as alternative splicing, yielding deletions of exon 3, 5, or 7. Using a yeast transactivation assay, they discovered receptors with 'outlaw' function, including both a dominant-positive receptor which was transcriptionally active in the absence of estrogen, and a dominant-negative receptor, which was itself transcriptionally inactive but prevented the action of normal estrogen receptor. McGuire et al. (1991) concluded that these variants could have clinical significance, helping to explain differences in the behavior of breast tumors and patient outcome.

Ponglikitmongkol et al. (1988) found that the ESR isolated from a human breast cancer cell line contained a gly400-to-val mutation in the hormone-binding domain. A highly conserved 66-amino acid region of the estrogen and glucocorticoid receptors, which corresponds to part of the receptor DNA-binding domain (region C), determines the specificity of target gene recognition. This region contains 2 subregions (CI and CII), encoded by 2 separate exons that are analogous to 'zinc fingers.' By the study of chimeric estrogen receptor, Mader et al. (1989) showed that 3 amino acids located on the C-terminal side of the CI finger play a key role in the specificity.

It is accepted that the presence of estrogen receptor identifies those breast cancer patients with a lower risk of relapse and better overall survival (Clark and McGuire, 1988), and the measurement of ESR has become a standard assay in the clinical management of breast cancer. Receptor status also provides a guideline for those tumors that may be responsive to hormonal intervention. But only about half of ESR-positive patients respond to various hormonal therapies and of those who do respond initially, most will eventually develop hormonally unresponsive disease following a period of treatment even though ESR is often still present. Sluyser and Mester (1985) hypothesized that the loss of hormone dependence of certain breast tumors may be due to the presence of mutated or truncated steroid receptors that activate transcription even in the absence of hormone. Fuqua et al. (1993) reviewed ESR mutations that may be important in breast cancer progression. Scott et al. (1991), for example, had found truncated forms of DNA-binding ESR in human breast cancer. To better understand structure-activity relationships of the human estrogen receptor, Weis et al. (1996) examined the role of tyrosine-537 in transcriptional response of the receptor. This residue is close to a region of the hormone-binding domain shown previously to be important in hormone-dependent transcriptional activity; it also has been proposed to be a tyrosine kinase phosphorylation site important in ESR activity. Weis et al. (1996) substituted 5 amino acids at this position (alanine, phenylalanine, glutamic acid, lysine, or serine) and screened these mutants for their biologic activities in the presence and absence of estradiol. Two of the ESR mutants, tyr537 to ala (Y537A) and tyr537 to ser (Y537S), displayed estrogen-independent constitutive activity that was approximately 20% or 100%, respectively, of the activity of the wildtype receptor with estradiol. In some circumstances, the tyr537-to-glu (Y537E) and tyr537-to-lys (Y537K) proteins also exhibited some low level of constitutive activity. Their findings indicated that tyrosine-537 is in a region important in the ligand regulation of ESR transcriptional activity and that certain amino acid substitutions at this position can shift ESR into a conformation that is active even without ligand.

Sluyser (1995) reviewed the somatically generated mutations in ESR that had been found at the mRNA/cDNA level in human breast cancer biopsies and in established breast cancer cell lines. Aberrantly spliced ESR mRNA causes the appearance of truncated or internally deleted ESR protein forms. Studies on the functional activity of ESR variants in expression systems demonstrated dominant-positive receptors that are transcriptionally active in the absence of estrogen, and dominant-negative receptors that are themselves transcriptionally inactive but that prevent the action of the normal receptor. The ESR variants are believed to confer resistance to endocrine therapy in breast cancer patients. Abnormally spliced forms of ESR, similar to those in breast cancer, were reported by McGuire et al. (1992) and by others. In all, 19 somatic mutations were tabulated and mapped on a diagram of the structural organization of the ESR gene.

Andersen et al. (1997) studied leukocyte DNA from 143 patients with familial clustering of breast and/or ovarian cancer and tumor DNA from 96 breast carcinomas for base mutations in the ESR gene. Three patients with a family history of cancer were carrying a gly160-to-cys germline substitution, which they concluded represents a polymorphism because it was detected in 4 females and 4 males of 729 controls, split about equally between males and females. However, in the 229 female controls in whom family history of cancer was known, 1 of 2 who had a sister with breast cancer was carrying the variant allele; hence, a possible clinical significance of the gly160-to-cys change should be further investigated. Somatic mutations were not detected in any of the tumors studied, and the data did not provide support for somatic ESR base mutations as an important mechanism for hormonal therapy resistance in estrogen receptor-positive breast carcinomas.

Using an Affymetrix 10K SNP array to screen for gene copy number changes in breast cancer, Holst et al. (2007) detected a single-gene amplification of the ESR1 gene. A subsequent tissue microarray analysis of more than 2,000 clinical breast cancer samples showed ESR1 amplification in 20.6% of breast cancers. In 99% of tumors with ESR1 amplification, overexpression of estrogen receptor protein was demonstrated, compared with 66.6% of cancers without ESR1 amplification. In 175 women who had received adjuvant tamoxifen monotherapy, survival was significantly longer for women with cancer with ESR1 amplification than for women with estrogen receptor-expressing cancers without ESR1 amplification (P = 0.023). Notably, they also found ESR1 amplification in benign and precancerous breast diseases, suggesting that ESR1 amplification may be a common mechanism in proliferative breast disease and a very early genetic alteration in a large subset of breast cancers.

In correspondences, Brown et al. (2008), Horlings et al. (2008), Vincent-Salomon et al. (2008), and Reis-Filho et al. (2008) reported attempts to replicate the finding of Holst et al. (2007) of a high frequency of ESR1 amplification in breast cancer. No group was able to replicate the results of Holst et al. (2007), using a variety of methods including array comparative genomic hybridization (CGH), FISH, and quantitative PCR. Amplification was found at a frequency of approximately 10% or less (Albertson, 2008). In a discussion of the findings of all of these groups, Albertson (2008) noted that although Holst et al. (2007) reported to have followed the standard procedure for scoring FISH, i.e., to count closely spaced signals as 1 signal, in their reply to the contesting groups Holst et al. (2008) emphasized the importance of scoring clusters of signals. Holst et al. (2008) stated, 'In our laboratory, most ESR1-amplified tumors have small gene clusters that could be considered as one signal if 'ERBB criteria' were applied....We therefore feel that estimating the ESR1 gene copy number may--given the currently available reagents--enable a more reliable identification of amplified cancers than classical counting.' Albertson (2008) concluded that this and other discrepancies, including that involving the clinical significance concerning prognosis, indicated that 'the jury is still out on the question of ESR1 amplification and its clinical significance.'

Certain malignant breast tumors (see 114480) are characterized by a high prediction uncertainty ('low-confidence') with respect to ESR status. Kun et al. (2003) analyzed these 'low-confidence' tumors and determined that their uncertain prediction status arose as a result of widespread perturbations in multiple genes whose expression is important for ESR-subtype discrimination. Patients with 'low-confidence' ESR-positive tumors exhibited a significantly worse overall survival (p = 0.03) and shorter time to distant metastasis (p = 0.004) compared with their 'high-confidence' ESR-positive counterparts. Elevated expression of ERBB2 (164870) was significantly correlated with a breast tumor exhibiting a 'low-confidence' prediction. Although ERBB2 signaling has been proposed to inhibit the transcriptional activity of ESR1, a large proportion of the perturbed genes in the 'low-confidence'/ERBB2-positive samples are not known to be estrogen responsive. Kun et al. (2003) proposed that a significant portion of the effect of ERBB2 on ESR-positive breast tumors may involve ESR-independent mechanisms of gene activation, which may contribute to the clinically aggressive behavior of the 'low-confidence' breast tumor subtype.

Toy et al. (2013) conducted a comprehensive genetic analysis of 2 independent cohorts of metastatic ER-positive breast tumors and identified mutations in ESR1 affecting the ligand-binding domain (LBD) in 14 of 80 cases. These included highly recurrent mutations encoding Y537S, Y537N, and asp538 to gly (D538G) alterations. Molecular dynamics simulations suggested that the structures of the Y537S and D538G mutants involve hydrogen bonding of the mutant amino acids with asp351, thus favoring the agonist conformation of the receptor. Consistent with this model, mutant receptors drive ER-dependent transcription and proliferation in the absence of hormone and reduce the efficacy of ER antagonists.

Robinson et al. (2013) enrolled 11 patients with ER-positive metastatic breast cancer in a prospective clinical sequencing program for advanced cancers. Whole-exome and transcriptome analysis identified 6 cases that harbored mutations of ESR1 affecting its LBD, all of whom had been treated with antiestrogens and estrogen deprivation therapies. A survey of The Cancer Genome Atlas identified 4 endometrial cancers with similar mutations of ESR1. The 5 LBD-localized ESR1 mutations identified, encoding L536Q, Y537S, Y537C, Y537, and D538G, were shown to result in constitutive activity and continued responsiveness to antiestrogen therapies in vitro.

Association with Bone Mineral Density Variation

Lorentzon et al. (1999) investigated the influence of ESRA gene polymorphism and estradiol on height and bone density during and after puberty in males. Using the restriction enzymes XbaI and PvuII, the allelic variants XX, Xx, xx, PP, Pp, and pp were identified in 90 Caucasian boys. In a multivariate analysis including pubertal development, physical activity, and body weight, the XbaI genotype independently predicted total body BMD, head BMD, and spine volumetric BMD (P less than 0.05). The PvuII genotype independently predicted spine volumetric BMD (pp greater than PP; P of 0.01). The 20 boys with the PP allelic variant were found to have a greater body height than the other 70 boys (182 cm vs 179 cm; P of 0.03). At a 2-year follow-up, the XbaI genotype was still independently related to total body BMD, head BMD, and spine volumetric BMD. The authors concluded that ESRA polymorphism is related to bone density and height during late puberty and at attainment of peak bone density in young men.

Bone mineral density, the major determinant of osteoporotic fracture risk, has a strong genetic component. The discovery that inactivation of ESR1 is associated with low BMD indicated ESR1 as a candidate gene for osteoporosis (166710). Becherini et al. (2000) genotyped 610 postmenopausal women for 3 ESR1 gene polymorphisms (intron 1 RFLPs PvuII and XbaI, and a (TA)n repeat 5-prime upstream of exon 1). Although no significant relationship between intron 1 RFLPs and BMD was observed, a statistically significant correlation between (TA)n-repeat allelic variants and lumbar BMD was observed (P = 0.04, ANCOVA), with subjects having a low number of repeats (TA less than 15) showing the lowest BMD values. The authors observed a statistically significant difference in the mean +/- SD number of (TA)n repeats between 73 analyzed women with a vertebral fracture and the nonfracture group, equivalent to a 2.9-fold increased fracture risk in women with a low number of repeats. The authors concluded that in their large sample the (TA)n polymorphism in ESR1 accounts for part of the heritable component of BMD and may prove useful in the prediction of vertebral fracture risk in postmenopausal osteoporosis.

Colin et al. (2003) studied the combined influence of polymorphisms in the ESR1 and the VDR (601769) genes on the susceptibility to osteoporotic vertebral fractures in 634 women aged 55 years and older. Three VDR haplotypes (1, 2, and 3) of the BsmI, ApaI, and TaqI restriction fragment length polymorphisms and 3 ESR1 haplotypes (1, 2, and 3) of the PvuII and XbaI RFLPs were identified. ESR1 haplotype 1 was dose-dependently associated with increased vertebral fracture risk corresponding to an odds ratio of 1.9 (95% confidence interval, 0.9-4.1) per copy of the risk allele. VDR haplotype 1 was overrepresented in vertebral fracture cases. There was a significant interaction (P = 0.01) between ESR1 haplotype 1 and VDR haplotype 1 in determining vertebral fracture risk. The association of ESR1 haplotype 1 with vertebral fracture risk was present only in homozygous carriers of VDR haplotype 1. The risk of fracture was 2.5 for heterozygous and 10.3 for homozygous carriers of ESR1 haplotype 1. These associations were independent of BMD. The authors concluded that interaction between ESR1 and VDR gene polymorphisms leads to increased risk of osteoporotic vertebral fractures in women, largely independent of BMD.

Van Meurs et al. (2003) investigated the influence of genetic variation in ESR1 on several bone parameters in 2,042 individuals of the Rotterdam Study, a prospective population-based cohort study of elderly individuals. They analyzed 3 polymorphic sites in the 5-prime region of the ESR1 gene: a (TA)n repeat in the promoter region, molecular haplotypes of the PvuII and XbaI RFLPs in intron 1, and inferred long-range haplotypes thereof. Linkage disequilibrium (LD) analysis between the PvuII-XbaI haplotype and the (TA)n repeat showed strong LD between the 2 sites. Reconstruction of long-range haplotypes over the entire 5-prime region revealed 6 frequent long-range haplotypes. In women only, there was an allele dosage effect of haplotype 'px' (p = 0.003) and a low number of (TA)n repeats (p = 0.008) with decreased lumbar spine BMD and decreased vertebral bone area (p = 0.016). There was also an increased vertebral fracture risk with evidence for an allele dosage effect with an odds ratio of 2.2 (95% CI, 1.3-3.5) for haplotype 'px' and an odds ratio of 2.0 (95% CI, 1.5-3.2) for a low number of (TA)n repeats. The ESR1 genotype-dependent fracture risk was largely independent of BMD and bone area. Van Meurs et al. (2003) concluded that ESR1 polymorphisms in the 5-prime promoter region are associated with vertebral fracture risk, lumbar spine BMD, and vertebral bone area in postmenopausal women, but not in men.

Khosla et al. (2004) studied relationships between polymorphisms of the ESRA and ESRB genes, BMD, and rates of bone loss in an age-stratified random sample of 283 Rochester, Minnesota, men aged 22 to 90 years. DNA was analyzed for the XbaI and PvuII ESRA and AluI ESRB polymorphisms. The X/P and x/p alleles of the ESRA gene were in strong linkage disequilibrium. BMD at multiple sites did not differ as a function of either the ESRA or ESRB genotype. However, the ESRA (but not ESRB) genotypes did modulate the relationship between BMD or rates of bone loss and bioavailable estradiol levels in these men. The authors concluded that the ESRA genotype may modulate the relationship between BMD or rates of bone loss and estrogen levels in men and that bone mass in men with the X or P alleles may be more susceptible to the consequences of estrogen deficiency (and conversely, benefit most from estrogen sufficiency) than in men with the xx or pp genotypes.

Sowers et al. (2004) conducted a 10-year prospective study of peak bone mass and its change in 604 women, aged 24 to 44 years at study initiation, and related changes in bone mineral density (BMD) and osteocalcin (OCN; 112260) concentrations to ESR1 gene polymorphisms (the XbaI and PvuII RFLPs) in 442 of these women. The authors concluded that while ESR1 genotype associations were statistically significant in explaining the rate of perimenopausal bone loss and its turnover, baseline BMI or becoming postmenopausal contributed more to the magnitude of the difference in bone change.

In 945 postmenopausal Scottish women who had not received hormone replacement therapy (non-HRT), Albagha et al. (2005) found that annual rates of femoral neck bone loss were 14% higher in subjects who carried 1 copy of the 'px' allele and 22% higher in those who carried 2 copies compared to those who did not carry the px haplotype (p = 0.009). The px haplotype was associated with lower femoral neck BMD in non-HRT postmenopausal women (p = 0.02), and with reduced calcaneal broadband ultrasound attenuation in the whole study population of 3,054 Scottish women (p = 0.005). Albagha et al. (2005) concluded that the ESR1 px haplotype is associated with reduced femoral neck BMD and increased rates of femoral neck bone loss in non-HRT postmenopausal women, and suggested that the association with broadband ultrasound attenuation may explain the fact that ESR1 intron 1 alleles predict osteoporotic fractures by a mechanism partly independent of differences in BMD.

Tobias et al. (2007) investigated whether the gain in area-adjusted bone mineral content (ABMC) in girls occurs in late puberty and examined whether the magnitude of this gain is related to ESR1 polymorphisms. For rs2234693 (PvuII) and rs9340799 (XbaI) polymorphisms, differences in spinal ABMC in late puberty were 2-fold greater in girls who were homozygous for the C and G alleles, respectively (P = 0.001). For rs7757956, the difference in total body less head ABMC in late puberty was 50% less in individuals homozygous or heterozygous for the A allele (P = 0.006). Tobias et al. (2007) concluded that gains in ABMC in late pubertal girls are strongly associated with ESR1 polymorphisms, suggesting that estrogen contributes to this process via an estrogen receptor-alpha-dependent pathway.

In a case-control study of 70 osteoporotic Mexican women and 70 nonosteoporotic female controls, Gomez et al. (2007) analyzed the (TA)n repeat and 2014G-A polymorphisms of the ESR1 gene and found that, with correction for population stratification, the 2014G allele was associated with osteoporosis (OR, 4.34; p = 0.006) in the Mexican population, whereas the TA repeat polymorphism was not.

In a genomewide association study to find common sequence variants that influence bone mineral density and low-trauma fractures in 3 populations of European descent, Styrkarsdottir et al. (2008) identified a complex pattern of association in the 6q25 region (see BMND11, 612114). SNPs in this region showed an association with BMD of both hip and spine, although no single SNP could fully explain the association. At least 3 SNPs were required to account for the overall association; one of these was in an intron of an ESR1 splice variant, and the other 2 were in the nearby C6ORF97 gene.

Association with Myocardial Infarction or Cardiovascular Risk Factors

In a study of 309 postmenopausal women with coronary artery disease, Herrington et al. (2002) found that women who were of a particular genotype at the ESR1 locus had an augmented response of HDL cholesterol to hormone replacement therapy (see 133430.0004).

In a study of 2,617 men and 3,791 postmenopausal women, aged 55 years and older, followed up over a period of 7 years or more, Schuit et al. (2004) found that postmenopausal women who carried the ESR1 haplotype 1 had an increased risk of myocardial infarction (MI; 608446) and ischemic heart disease, independent of known cardiovascular risk factors. In men, no association was observed. Haplotype 1 is comprise of 2 polymorphisms located in the first intron of the ESR1 gene, 397 bp (PvuII; rs2234693) and 351 bp (XbaI; rs9340799) upstream of exon 2.

Taguchi (2004) raised the question of whether the connection between ESR1 and MI might not be direct. Since Mattila et al. (1989) reported an association between poor dental health and acute MI in men, several studies had shown such an association in women as well (Emingil et al., 2000); Taguchi et al. (2001) reported a significant association between ESR1 polymorphisms and tooth loss in postmenopausal women. Thus, the association between ESR1 polymorphisms and MI in postmenopausal women could actually be due to an association between ESR1 polymorphisms and tooth loss. Another possibility that Taguchi (2004) raised was that endothelial dysfunction that may lead to MI contributes to periodontitis and subsequent tooth loss in postmenopausal women. A third possibility was that the association between ESR1 polymorphisms and tooth loss and between ESR1 polymorphisms and MI are both real, so that the association between tooth loss and MI in postmenopausal women in previous studies is spurious and actually due to the ESR1 polymorphisms.

Shearman et al. (2005) tested for interaction between smoking and ESR1 variation in association with plasma concentration of atherogenic small, low density lipoprotein (LDL) particles and LDL particle size. Among 1,727 unrelated subjects from the population-based Framingham Heart Study, women who smoked and had the common ESR1 rs2234693 TT genotype had more than 1.7-fold higher levels of small LDL particles than women with the alternative genotypes (P-value for smoking-genotype interaction = 0.001). Similar results were obtained for 3 other ESR1 variants including rs9340799, in the same linkage disequilibrium block. A similar substantial gender-specific result was also evident with a 975C-G variant, in a separate linkage disequilibrium block, in exon 4 (P = 0.003). Women who smoked and had specific, common ESR1 genotypes had a substantially higher plasma concentration of atherogenic small LDL particles. Significant results revealed a dose-dependent effect of smoking and were evident in both pre- and postmenopausal women. The reported association has the potential to explain the risks associated with estrogen use in certain women and an association between the ESR1 haplotype 1 (rs2234693 T allele and rs9340799 A allele) with increased MI and ischemic heart disease, independent of the standard, established cardiovascular risk factors.

Associations Pending Confirmation

For discussion of a possible association between variation in the ESR1 gene and osteoarthritis of the knee, see 165720.

For discussion of a possible association between variation in the ESR1 gene and the waist-to-hip ratio in women, see 605552.

For discussion of a possible association between variation in the ESR1 gene and age-related macular degeneration, see ARMD1 (603075).

For discussion of a possible association between variation in the ESR1 gene and age of onset of menopause, see 300488.


Animal Model

Korach (1994) investigated hormone responsiveness in genetic mutant mice without a functional estrogen receptor, created through gene knockout techniques. Both sexes of these mutant animals were infertile and showed a variety of phenotypic changes associated with the gonads, mammary glands, reproductive tracts, and skeletal tissues.

To clarify the role of estrogen signaling in reproductive tract development and function, Couse et al. (1999) generated mice lacking ESRA and ESRB by targeted disruption. ESRA/ESRB knockout males were infertile but possessed a grossly normal reproductive tract. They exhibited various stages of spermatogenesis, but the numbers and motility of epididymal sperm were reduced significantly. ESRA/ESRB knockout females exhibited proper differentiation of the mullerian-derived structures of the uterus, cervix, and upper vagina, but these structures were severely hypoplastic in adults. Similar uterine hypoplasia was observed in ESRA, but not in ESRB, knockout mice. The ovaries of adult ESRA/ESRB knockout females exhibited morphologic phenotypes that were clearly distinct from those of the prepubertal ESRA/ESRB knockout females and the individual estrogen receptor knockout mice. The double-knockout female ovaries had structures resembling seminiferous tubules of the testis. Within the lumen of the tubule-like structures were degenerating granulosa cells and cells resembling Sertoli cells of the testis. Couse et al. (1999) argued that certain characteristics of the adult ESRA/ESRB knockout ovary indicated redifferentiation of varying components rather than a developmental phenomenon: the absence of similar structures in prepubertal ESRA/ESRB knockout ovaries; the consistent spherical shape of the tubules, suggesting origination from a once healthy follicle; and age-related increases in the area of transdifferentiation. The ovaries of adult ESRA/ESRB knockout females expressed mullerian-inhibiting substance (600957), sulfated glycoprotein-2 (185430), and Sox9 (608160). Couse et al. (1999) concluded that the loss of both receptors leads to an ovarian phenotype that is distinct from that of the individual estrogen receptor knockout mutants, which indicates that both receptors are required for the maintenance of germ and somatic cells in the postnatal ovary.

Heine et al. (2000) found that male and female Esr1 knockout mice had hyperplasia and hypertrophy of adipocytes, insulin resistance, and glucose intolerance. The results provided evidence that estrogen/ESR1 signaling is critical in female and male white adipose tissue; obesity in the knockout males involved a mechanism of reduced energy expenditure rather than increased energy intake. Similar results were obtained by Jones et al. (2000) studying aromatase (CYP19; 107910) knockout mice, which cannot synthesize endogenous estrogens. Both male and female aromatase knockout mice progressively accumulated significantly more intraabdominal adipose tissue than their wildtype littermates and had elevated circulating levels of leptin and cholesterol, as well as elevated insulin levels and a striking accumulation of lipid droplets in the livers.

Davis et al. (2002) noted that studies in humans and rodent models had suggested that estrogen may provide protection against age-related cataracts. The presence of estrogen receptors in the eye indicates that estrogen protection may result from direct interactions with its receptors in the eye, instead of being an indirect consequence from effects on another tissue. Davis et al. (2002) validated the concept that estrogen is beneficial for the eye. In transgenic mice expressing ESR-delta-3, a natural variant of ESR1 with an in-frame deletion of exon 3 resulting from alternative splicing, they found that cortical cataracts spontaneously formed in females after puberty and progressed with age. ESR-delta-3 is a dominant-negative form of ESR-alpha that inhibits ESR-alpha function. Cataract formation could be prevented if the females were ovariectomized before, but not after, sexual maturity. Both male and female ESR-delta-3 mice developed cataracts after neonatal treatment with the potent estrogen DES. The incidence of spontaneous and DES-induced cataracts in ESR-delta-3 mice was 100%, whereas such cataracts were absent from wildtype mice. The data suggested that repression of estrogen action induces cortical cataract formation because estrogen is required to activate the ESR-delta-3 repressor. Evidence of DES-induced cataracts in the ESR-delta-3 males as well as the females suggested that estrogen is important in lens physiology in both sexes.

Using mice lacking functional Esr1, Lee et al. (2003) showed that bone in vivo undergoes an adaptive response to loading that is less effective in the absence of Esr1 and that osteoblast-like cells require Esr1 to proliferate in response to mechanical strain in vitro. Lee et al. (2003) speculated that as ESR1 expression in osteoblasts and osteocytes depends on estrogen concentration, a failure to maintain bone strength after menopause might be due to reduction in the activity of ESR1 in bone cells, thereby limiting their anabolic response to mechanical loading and allowing a loss of bone tissue comparable to that associated with disuse.

Garey et al. (2003) identified a generalized arousal component in the behavior of mice. Analyzed by mathematical/statistical approaches across experiments, investigators, and mouse populations, it accounted for approximately one-third of the variance in arousal-related measures. Knockout of the Esr1 gene greatly reduced arousal responses. In contrast, disrupting the Esr2 gene (601663), a likely gene duplication product which encodes ER-beta, did not have these effects.

In mice, ovariectomy accelerates the progression of the end-stage renal disease glomerulosclerosis. In women, the incidence of this disorder increases after menopause, and estrogen alters its progression. Polymorphisms in the ESR1 gene may constitute a genetic predisposition for lupus nephritis (Liu et al., 2002). Shim et al. (2004) showed that by 1 year of age, mice lacking ER-alpha, but not those lacking ER-beta, exhibited immune complex-type glomerulonephritis, proteinuria, and destruction of tubular cells. The mice also showed spontaneous formation of germinal centers in the spleen in the absence of antigen challenge and infiltration of plasma cells in the kidney and spleen. The results indicated that ER-alpha has indispensable functions in the kidney and in germinal centers, and that defective ER-alpha signaling results in glomerulonephritis.

In female mice and rats, Musatov et al. (2007) used RNAi to focally silence ER-alpha in the ventromedial nucleus of the hypothalamus and observed development of a phenotype characteristic of metabolic syndrome, marked by obesity, hyperphagia, impaired glucose tolerance, and reduced energy expenditure; this phenotype persisted despite normal ER-alpha levels elsewhere in the brain. Although an increase in food intake preceded weight gain, the authors stated that their data suggested that a leading factor of obesity in this model was likely a decline in energy expenditure with all 3 major constituents being affected, including voluntary activity, basal metabolic rate, and diet-induced thermogenesis. Musatov et al. (2007) concluded that ER-alpha in the ventromedial nucleus of the hypothalamus neurons plays an essential role in the control of energy balance and maintenance of normal body weight.

Nakamura et al. (2007) selectively ablated Esr1 in differentiated mouse osteoclasts and found that females, but not males, exhibited trabecular bone loss, similar to the osteoporotic bone phenotype in postmenopausal women. Furthermore, estrogen induced apoptosis and upregulation of Fas ligand (FASL, or TNFSF6; 134638) in osteoclasts of trabecular bones of wildtype mice, but not mutant mice. Expression of Esr1 was also required for induction of apoptosis by tamoxifen and estrogen in cultured osteoclasts. Nakamura et al. (2007) concluded that estrogen regulates the life span of mature osteoclasts via induction of the FAS (TNFRSF6; 134637)/FASL system.

To examine the ability of estrogen response element (ERE)-independent ESR1 signaling pathways to convey estrogen feedback regulation of the female hypothalamic-pituitary axis, Glidewell-Kenney et al. (2007) bred knockin mice expressing a mutant form of Esr1 ('AA') that has ablated ERE activity but intact ERE-independent activity with Esr1 -/- mice. The Esr1 -/AA mice exhibited 70% lower serum LH levels compared to Esr1 -/- mice. In addition, like wildtype mice, Esr1 -/AA mice exhibited elevated LH after ovariectomy, and the postovariectomy rise in LH was significantly suppressed by estrogen treatment in the ovariectomized Esr1 -/AA mice. However, unlike wildtype, both Esr1 -/AA and Esr1 -/- mice failed to exhibit estrous cyclicity, spontaneous ovulation, or an afternoon LH surge response to estrogen. Glidewell-Kenney et al. (2007) concluded that ERE-independent ESR1 signaling is sufficient to convey the majority of estrogen's negative feedback actions, whereas positive feedback and spontaneous ovulatory cyclicity require ERE-dependent ESR1 signaling.

Park et al. (2011) used Esr1 -/AA mice to assess the role of noncanonical ESR1 signaling in energy homeostasis and found that nonclassical ESR1 signaling restored metabolic parameters dysregulated in Esr1 -/- mice to normal or near-normal values. The rescue of body weight and metabolic function by nonclassical ESR1 signaling was mediated by normalization of energy expenditure, including voluntary locomotor activity. Park et al. (2011) concluded that nonclassical ESR1 signaling mediates major effects of estradiol-17-beta on energy balance.


ALLELIC VARIANTS 7 Selected Examples):

.0001   ESTROGEN RECEPTOR MUTANT, TEMPERATURE-SENSITIVE

ESR1, CYS447ALA
SNP: rs121913043, ClinVar: RCV000018058

Reese and Katzenellenbogen (1991) identified an estrogen receptor mutant that had a similar binding affinity for estradiol as wildtype ESR but displayed a dose-response shift for estradiol in transactivation studies. The mutant contained an alanine substitution for cysteine at amino acid 447 in the hormone binding domain of the receptor. Reese and Katzenellenbogen (1992) showed by hormone binding studies that the C447A receptor is a temperature-sensitive mutant, whose instability is only apparent at elevated temperatures, and that ligand can stabilize the mutant receptor. The mutant also showed a temperature-sensitive loss in the DNA binding ability of the receptor. (The mutant was one of several created by Reese and Katzenellenbogen (1991) by in vitro oligonucleotide site-directed mutagenesis of human ESR cDNA. Function of the mutant forms was tested in Chinese hamster ovary (CHO) cells, an estrogen receptor-deficient cell line.)


.0002   ESTROGEN RESISTANCE

ESR1, ARG157TER
SNP: rs104893956, ClinVar: RCV000018060

In a 28-year-old man with estrogen resistance (ESTRR; 615363), Smith et al. (1994) identified homozygosity for a C-T transition in exon 2 of the ESR1 gene, resulting in an arg157-to-ter (R157X) substitution. Both parents were heterozygous carriers of the R157X mutation, and pedigree analysis showed that they were related as second cousins; 3 sisters were also heterozygous for the mutation.


.0003   ESTROGEN RESISTANCE

ESR1, VAL364GLU
SNP: rs121913044, ClinVar: RCV000018061

McInerney et al. (1996) characterized a human ESR mutant, val364 to glu, which has a single amino acid substitution in its hormone-binding domain. While this mutant is fully active or even superactive at saturating levels of estradiol it also acts as a strong dominant-negative inhibitor of the wildtype ESR and it is able to repress ESR-mediated transcription when the mutant and wildtype ESR are present together in cells, even without DNA binding. It is probable that altered interactions with proteins important in ESR-mediated transcription play a key role in the repression of transcription by val364 to glu.


.0004   MYOCARDIAL INFARCTION, SUSCEPTIBILITY TO

ESR1, IVS1AS, T-C, -401
SNP: rs2234693, gnomAD: rs2234693, ClinVar: RCV000018062

Myocardial Infarction, Susceptibility to

In a study of atherosclerotic cardiovascular events in 1,739 unrelated men and women from the Framingham Heart Study, Shearman et al. (2003) found that after adjustment for covariates, the ESR1 C/C genotype was significantly associated with major atherosclerotic cardiovascular disease and myocardial infarction (608446) (odds ratios of 2.0 and 3.0, respectively, compared to individuals with the C/T or T/T genotypes). The results remained significant when analyses were restricted to men; too few women had events to study them separately. Shearman et al. (2003) concluded that individuals with the common ESR1 C/C genotype have a substantial increase in risk of myocardial infarction.

Atherosclerosis, Susceptibility to

Lehtimaki et al. (2002) examined coronary artery specimens from 300 Finnish white men aged 33 to 69 years included in the Helsinki Sudden Death Study and determined the ESR1 IVS1 -401T/C (or PvuII) genotype. After adjusting for age and BMI, men aged 53 years or over with C/T and C/C genotypes had areas of complicated lesions on average 2- and 5-fold larger, respectively, than subjects with the T/T genotype. The age and BMI-adjusted odds ratios for coronary thrombosis were 6.2 for C/T and 10.6 for C/C compared to men with the T/T genotype. After additional adjustment for diabetes and hypertension, ESR1 genotype persisted as an independent predictor of complicated lesions and coronary thrombosis. Lehtimaki et al. (2002) concluded that the ESR1 gene is a potential candidate behind the pathogenesis of acute coronary events.

HDL Cholesterol, Augmented Response of, to Hormone Replacement

Herrington et al. (2002) characterized 309 women with coronary artery disease with respect to 8 previously described and 2 novel ESR1 polymorphisms, and examined the association between these polymorphisms and the response of HDL cholesterol and other lipids to treatment with estrogen alone or estrogen plus progestin. They found that postmenopausal women who had the ESR1 C/C genotype at the -401 position in intron 1, or several other closely related genotypes, had an augmented response of HDL cholesterol to hormone replacement therapy.


.0005   MIGRAINE, SUSCEPTIBILITY TO

ESR1, 594G-A
SNP: rs2228480, gnomAD: rs2228480, ClinVar: RCV001674740, RCV001843377, RCV002506722, RCV003975895

Migraine (157300) is a painful and debilitating disorder with a significant genetic component. Steroid hormones, in particular estrogen, have long been considered to play a role in migraine, as variations in hormone levels are associated with migraine onset in many individuals with the disorder. Steroid hormones mediate their activity via hormone receptors, which have a wide tissue distribution. Estrogen receptors have been localized to the brain in regions considered to be involved in migraine pathogenesis. Colson et al. (2004) examined the ESR1 gene for a potential role in migraine pathogenesis and susceptibility. A population-based cohort of 224 patients with migraine and 224 matched controls were genotyped for the 594G-A polymorphism located in exon 8 of the ESR1 gene. Statistical analysis indicated a significant difference between patients with migraine and those without migraine in both the allele frequencies (P = 0.003) and genotype distributions (P = 0.008). An independent follow-up study using this marker in an additional population-based cohort of 260 patients with migraine and 260 matched controls resulted in a significant association between the 2 groups with regard to allele frequencies and genotype distributions. The findings supported the hypothesis that genetic variation in hormone receptors, in particular the ESR1 gene, may play a role in migraines.


.0006   ESTROGEN RESISTANCE

ESR1, GLN375HIS
SNP: rs397509428, ClinVar: RCV000054445

In an 18-year-old woman with estrogen resistance (ESTRR; 615363), Quaynor et al. (2013) identified homozygosity for a c.1125G-T transversion in exon 5 of the ESR1 gene, resulting in a gln375-to-his (Q375H) substitution at a highly conserved residue within the ligand-binding domain. The patient had been adopted, thus DNA was not available from her parents; however, high-density microarray analysis revealed an approximately 11% region of homozygosity, suggesting that her biologic parents were second-degree relatives and likely heterozygous carriers of the mutation. Functional analysis in transfected COS-7 cells showed greatly reduced activity in the mutated estrogen receptor, with an EC50 that was 240-fold greater than that of the wildtype receptor.


.0007   ESTROGEN RESISTANCE

ESR1, ARG394HIS
SNP: rs1131692059, ClinVar: RCV000495838

In 3 sibs from a consanguineous Algerian family with estrogen resistance (ESTRR; 615363), Bernard et al. (2017) identified homozygosity for a c.1181G-A transition (c.1181G-A, NM_000125.3) in exon 5 of the ESR1 gene, resulting in an arg394-to-his (R394H) substitution at a highly conserved residue in the ligand-binding domain. Their unaffected parents and an unaffected sister were heterozygous for the mutation, which was not found in the Exome Variant Server or ExAC databases. Analysis of transiently transfected HEK293T cells demonstrated highly reduced sensitivity of the mutant receptor to 17-beta-estradiol stimulation, with an ED50 that was 65-fold higher than that of wildtype.


See Also:

Castagnoli et al. (1987); Jeltsch et al. (1987)

REFERENCES

  1. Abbondanza, C., Rossi, V., Roscigno, A., Gallo, L., Belsito, A., Piluso, G., Medici, N., Nigro, V., Molinari, A. M., Moncharmont, B., Puca, G. A. Interaction of vault particles with estrogen receptor in the MCF-7 breast cancer cell. J. Cell Biol. 141: 1301-1310, 1998. [PubMed: 9628887] [Full Text: https://doi.org/10.1083/jcb.141.6.1301]

  2. Aguirre, J. I., Plotkin, L. I., Gortazar, A. R., Martin Millan, M., O'Brien, C. A., Manolagas, S. C., Bellido, T. A novel ligand-independent function of the estrogen receptor is essential for osteocyte and osteoblast mechanotransduction. J. Biol. Chem. 282: 25501-25508, 2007. [PubMed: 17609204] [Full Text: https://doi.org/10.1074/jbc.M702231200]

  3. Albagha, O. M. E., Pettersson, U., Stewart, A., McGuigan, F. E. A., MacDonald, H. M., Reid, D. M., Ralston, S. H. Association of oestrogen receptor alpha gene polymorphisms with postmenopausal bone loss, bone mass, and quantitative ultrasound properties of bone. J. Med. Genet. 42: 240-246, 2005. [PubMed: 15744038] [Full Text: https://doi.org/10.1136/jmg.2004.023895]

  4. Albertson, D. G. Conflicting evidence on the frequency of ESR1 amplification in breast cancer. Nature Genet. 40: 821-822, 2008. [PubMed: 18583976] [Full Text: https://doi.org/10.1038/ng0708-821]

  5. Andersen, T. I., Wooster, R., Laake, K., Collins, N., Warren, W., Skrede, M., Eeles, R., Tveit, K. M., Johnston, S. R. D., Dowsett, M., Olsen, A. O., Moller, P., Stratton, M. R., Borresen-Dale, A.-L. Screening for ESR mutations in breast and ovarian cancer patients. Hum. Mutat. 9: 531-536, 1997. [PubMed: 9195227] [Full Text: https://doi.org/10.1002/(SICI)1098-1004(1997)9:6<531::AID-HUMU6>3.0.CO;2-4]

  6. Auboeuf, D., Honig, A., Berget, S. M., O'Malley, B. W. Coordinate regulation of transcription and splicing by steroid receptor coregulators. Science 298: 416-419, 2002. [PubMed: 12376702] [Full Text: https://doi.org/10.1126/science.1073734]

  7. Balleine, R. L., Hunt, S. M. N., Clarke, C. L. Coexpression of alternatively spliced estrogen and progesterone receptor transcripts in human breast cancer. J. Clin. Endocr. Metab. 84: 1370-1377, 1999. [PubMed: 10199781] [Full Text: https://doi.org/10.1210/jcem.84.4.5637]

  8. Becherini, L., Gennari, L., Masi, L., Mansani, R., Massart, F., Morelli, A., Falchetti, A., Gonnelli, S., Fiorelli, G., Tanini, A., Brandi, M. L. Evidence of a linkage disequilibrium between polymorphisms in the human estrogen receptor-alpha gene and their relationship to bone mass variation in postmenopausal Italian women. Hum. Molec. Genet. 9: 2043-2050, 2000. [PubMed: 10942433] [Full Text: https://doi.org/10.1093/hmg/9.13.2043]

  9. Bernard, V., Kherra, S., Francou, B., Fagart, J., Viengchareun, S., Guechot, J., Ladjouze, A., Guiochon-Mantel, A., Korach, K. S., Binart, N., Lombes, M., Christin-Maitre, S. Familial multiplicity of estrogen insensitivity associated with a loss-of-function ESR1 mutation. J. Clin. Endocr. Metab. 102: 93-99, 2017. [PubMed: 27754803] [Full Text: https://doi.org/10.1210/jc.2016-2749]

  10. Bord, S., Horner, A., Beavan, S., Compston, J. Estrogen receptors alpha and beta are differentially expressed in developing human bone. J. Clin. Endocr. Metab. 86: 2309-2314, 2001. [PubMed: 11344243] [Full Text: https://doi.org/10.1210/jcem.86.5.7513]

  11. Brown, L. A., Hoog, J., Chin, S.-F., Tao, Y., Zayed, A. A., Chin, K., Teschendorff, A. E., Quackenbush, J. F., Marioni, J. C., Leung, S., Perou, C. M., Neilsen, T. O., Ellis, M., Gray, J. W., Bernard, P. S., Huntsman, D. G., Caldas, C. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 806-807, 2008. [PubMed: 18583964] [Full Text: https://doi.org/10.1038/ng0708-806]

  12. Castagnoli, A., Maestri, I., Bernardi, F., Del Senno, L. PvuII RFLP inside the human estrogen receptor gene. Nucleic Acids Res. 15: 866 only, 1987. [PubMed: 2881267] [Full Text: https://doi.org/10.1093/nar/15.2.866]

  13. Chaidarun, S. S., Alexander, J. M. A tumor-specific truncated estrogen receptor splice variant enhances estrogen-stimulated gene expression. Molec. Endocr. 12: 1355-1366, 1998. [PubMed: 9731704] [Full Text: https://doi.org/10.1210/mend.12.9.0170]

  14. Chiang, C.-H., Cheng, K. W., Igarashi, S., Nathwani, P. S., Leung, P. C. K. Hormonal regulation of estrogen receptor alpha and beta gene expression in human granulosa-luteal cells in vitro. J. Clin. Endocr. Metab. 85: 3828-3839, 2000. [PubMed: 11061546] [Full Text: https://doi.org/10.1210/jcem.85.10.6886]

  15. Cho, J., Yu, N.-K., Choi, J.-H., Sim, S.-E., Kang, S. J., Kwak, C., Lee, S.-W., Kim, J., Choi, D. I., Kim, V. N., Kaang, B.-K. Multiple repressive mechanisms in the hippocampus during memory formation. Science 350: 82-87, 2015. [PubMed: 26430118] [Full Text: https://doi.org/10.1126/science.aac7368]

  16. Clark, G. M., McGuire, W. L. Steroid receptors and other prognostic factors in primary breast cancer. Semin. Oncol. 15 (suppl. 1): 20-25, 1988. [PubMed: 3285482]

  17. Clegg, D. J., Palmer, B. F. Effects of an estrogen receptor alpha variant. (Letter) New Eng. J. Med. 369: 1663-1664, 2013. [PubMed: 24152274] [Full Text: https://doi.org/10.1056/NEJMc1310364]

  18. Colin, E. M., Uitterlinden, A. G., Meurs, J. B. J., Bergink, A. P., van de Klift, M., Fang, Y., Arp, P. P., Hofman, A., van Leeuwen, J. P. T. M., Pols, H. A. P. Interaction between vitamin D receptor genotype and estrogen receptor alpha genotype influences vertebral fracture risk. J. Clin. Endocr. Metab. 88: 3777-3784, 2003. [PubMed: 12915669] [Full Text: https://doi.org/10.1210/jc.2002-021861]

  19. Colson, N. J., Lea, R. A., Quinlan, S., MacMillan, J., Griffiths, L. R. The estrogen receptor 1 G594A polymorphism is associated with migraine susceptibility in two independent case/control groups. Neurogenetics 5: 129-133, 2004. [PubMed: 15133719] [Full Text: https://doi.org/10.1007/s10048-004-0181-4]

  20. Couse, J. F., Hewitt, S. C., Bunch, D. O., Sar, M., Walker, V. R., Davis, B. J., Korach, K. S. Postnatal sex reversal of the ovaries in mice lacking estrogen receptors alpha and beta. Science 286: 2328-2331, 1999. [PubMed: 10600740] [Full Text: https://doi.org/10.1126/science.286.5448.2328]

  21. Davis, V. L., Chan, C.-C., Schoen, T. J., Couse, J. F., Chader, G. J., Korach, K. S. An estrogen receptor repressor induces cataract formation in transgenic mice. Proc. Nat. Acad. Sci. 99: 9427-9432, 2002. [PubMed: 12082175] [Full Text: https://doi.org/10.1073/pnas.132247999]

  22. Emingil, G., Buduneli, E., Aliyev, A., Akilli, A., Atilla, G. Association between periodontal disease and acute myocardial infarction. J. Periodont. 71: 1882-1886, 2000. [PubMed: 11156045] [Full Text: https://doi.org/10.1902/jop.2000.71.12.1882]

  23. Esmaeli, B., Harvey, J. T., Hewlett, B. Immunohistochemical evidence for estrogen receptors in meibomian glands. Ophthalmology 107: 180-184, 2000. [PubMed: 10647739] [Full Text: https://doi.org/10.1016/s0161-6420(99)00040-8]

  24. Fan, S., Wang, J.-A., Yuan, R., Ma, Y., Meng, Q., Erdos, M. R., Pestell, R. G., Yuan, F., Auborn, K. J., Goldberg, I. D., Rosen, E. M. BRCA1 inhibition of estrogen receptor signaling in transfected cells. Science 284: 1354-1356, 1999. [PubMed: 10334989] [Full Text: https://doi.org/10.1126/science.284.5418.1354]

  25. 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-alpha) that is encoded by distinct transcripts and that is able to repress hER-alpha activation function 1. EMBO J. 19: 4688-4700, 2000. [PubMed: 10970861] [Full Text: https://doi.org/10.1093/emboj/19.17.4688]

  26. Fullwood, M. J., Liu, M. H., Pan, Y. F., Liu, J., Xu, H., Mohamed, Y. B., Orlov, Y. L., Velkov, S., Ho, A., Mei, P. H., Chew, E. G. Y., Huang, P. Y. H., and 29 others. An oestrogen-receptor-alpha-bound human chromatin interactome. Nature 462: 58-64, 2009. [PubMed: 19890323] [Full Text: https://doi.org/10.1038/nature08497]

  27. Fuqua, S. A. W., Chamness, G. C., McGuire, W. L. Estrogen receptor mutations in breast cancer. J. Cell. Biochem. 51: 135-139, 1993. [PubMed: 8440747] [Full Text: https://doi.org/10.1002/jcb.240510204]

  28. Garcia-Morales, P., Saceda, M., Kenney, N., Kim, N., Salomon, D. S., Gottardis, M. M., Solomon, H. B., Sholler, P. F., Jordan, V. C., Martin, M. B. Effect of cadmium on estrogen receptor levels and estrogen-induced responses in human breast cancer cells. J. Biol. Chem. 269: 16896-16901, 1994. [PubMed: 8207012]

  29. Garey, J., Goodwillie, A., Frohlich, J., Morgan, M., Gustafsson, J.-A., Smithies, O., Korach, K. S., Ogawa, S., Pfaff, D. W. Genetic contributions to generalized arousal of brain and behavior. Proc. Nat. Acad. Sci. 100: 11019-11022, 2003. [PubMed: 12930901] [Full Text: https://doi.org/10.1073/pnas.1633773100]

  30. Glidewell-Kenney, C., Hurley, L. A., Pfaff, L., Weiss, J., Levine, J. E., Jameson, J. L. Nonclassical estrogen receptor alpha signaling mediates negative feedback in the female mouse reproductive axis. Proc. Nat. Acad. Sci. 104: 8173-8177, 2007. [PubMed: 17470805] [Full Text: https://doi.org/10.1073/pnas.0611514104]

  31. Gomez, R., Magana, J. J., Cisneros, B., Perez-Salazar, E., Faugeron, S., Veliz, D., Castro, C., Rubio, J., Casas, L., Valdes-Flores, M. Association of the estrogen receptor alpha gene polymorphisms with osteoporosis in the Mexican population. Clin. Genet. 72: 574-581, 2007. [PubMed: 17953702] [Full Text: https://doi.org/10.1111/j.1399-0004.2007.00898.x]

  32. Gosden, J. R., Middleton, P. G., Rout, D. Localization of the human oestrogen receptor gene to chromosome 6q24-q27 by in situ hybridization. Cytogenet. Cell Genet. 43: 218-220, 1986. [PubMed: 3802924] [Full Text: https://doi.org/10.1159/000132325]

  33. Green, S., Walter, P., Kumar, V., Krust, A., Bornert, J.-M., Argos, P., Chambon, P. Human oestrogen receptor cDNA: sequence, expression and homology to v-erb-A. Nature 320: 134-139, 1986. [PubMed: 3754034] [Full Text: https://doi.org/10.1038/320134a0]

  34. Greene, G. L., Gilna, P., Waterfield, M., Baker, A., Hort, Y., Shine, J. Sequence and expression of human estrogen receptor complementary DNA. Science 231: 1150-1154, 1986. [PubMed: 3753802] [Full Text: https://doi.org/10.1126/science.3753802]

  35. He, Z., Zhang, S. S., Meng, Q., Li, S., Zhu, H. H., Raquil, M.-A., Alderson, N., Zhang, H., Wu, J., Rui, L., Cai, D., Feng, G.-S. Shp2 controls female body weight and energy balance by integrating leptin and estrogen signals. Molec. Cell. Biol. 32: 1867-1878, 2012. [PubMed: 22431513] [Full Text: https://doi.org/10.1128/MCB.06712-11]

  36. Heine, P. A., Taylor, J. A., Iwamoto, G. A., Lubahn, D. B., Cooke, P. S. Increased adipose tissue in male and female estrogen receptor-alpha knockout mice. Proc. Nat. Acad. Sci. 97: 12729-12734, 2000. [PubMed: 11070086] [Full Text: https://doi.org/10.1073/pnas.97.23.12729]

  37. 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. New Eng. J. Med. 346: 967-974, 2002. [PubMed: 11919305] [Full Text: https://doi.org/10.1056/NEJMoa012952]

  38. Holst, F., Stahl, P., Hellwinkel, O., Dancau, A.-M., Krohn, A., Wuth, L., Heupel, C., Lebeau, A., Terracciano, L., Al-Kuraya, K., Janicke, F., Sauter, G., Simon, R. Reply to 'ESR1 gene amplification in breast cancer: a common phenomenon?' (Letter) Nature Genet. 40: 810-812, 2008.

  39. Holst, F., Stahl, P. R., Ruiz, C., Hellwinkel, O., Jehan, Z., Wendland, M., Lebeau, A., Terracciano, L., Al-Kuraya, K., Janicke, F., Sauter, G., Simon, R. Estrogen receptor alpha (ESR1) gene amplification is frequent in breast cancer. Nature Genet. 39: 655-660, 2007. [PubMed: 17417639] [Full Text: https://doi.org/10.1038/ng2006]

  40. Horlings, H. M., Bergamaschi, A., Nordgard, S. H., Kim, Y. H., Han, W., Noh, D.-Y., Salari, K., Joosse, S. A., Reyal, F., Lingjaerde, O. C., Kristensen, V. N., Borresen-Dale, A.-L., Pollack, J., van de Vijver, M. J. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 807-808, 2008. [PubMed: 18583965] [Full Text: https://doi.org/10.1038/ng0708-807]

  41. Hurtado, A., Holmes, K. A., Geistlinger, T. R., Hutcheson, I. R., Nicholson, R. I., Brown, M., Jiang, J., Howat, W. J., Ali, S., Carroll, J. S. Regulation of ERBB2 by oestrogen receptor-PAX2 determines response to tamoxifen. Nature 456: 663-666, 2008. Note: Erratum: Nature 457: 1168 only, 2009. [PubMed: 19005469] [Full Text: https://doi.org/10.1038/nature07483]

  42. Hurtado, A., Holmes, K. A., Ross-Innes, C. S., Schmidt, D., Carroll, J. S. FOXA1 is a key determinant of estrogen receptor function and endocrine response. Nature Genet. 43: 27-33, 2011. [PubMed: 21151129] [Full Text: https://doi.org/10.1038/ng.730]

  43. Issa, J.-P. J., Ottaviano, Y. L., Celano, P., Hamilton, S. R., Davidson, N. E., Baylin, S. B. Methylation of the oestrogen receptor CpG island links ageing and neoplasia in human colon. Nature Genet. 7: 536-540, 1994. [PubMed: 7951326] [Full Text: https://doi.org/10.1038/ng0894-536]

  44. Jeltsch, J. M., Roberts, M., Schatz, C., Garnier, J. M., Brown, A. M. C., Chambon, P. Structure of the human oestrogen-responsive gene pS2. Nucleic Acids Res. 15: 1401-1414, 1987. [PubMed: 3822834] [Full Text: https://doi.org/10.1093/nar/15.4.1401]

  45. Johnson, M. D., Kenney, N., Stoica, A., Hilakivi-Clarke, L., Singh, B., Chepko, G., Clarke, R., Sholler, P. F., Lirio, A. A., Foss, C., Reiter, R., Trock, B., Paik, S., Martin, M. B. Cadmium mimics the in vivo effects of estrogen in the uterus and mammary gland. Nature Med. 9: 1081-1084, 2003. [PubMed: 12858169] [Full Text: https://doi.org/10.1038/nm902]

  46. Jones, M. E. E., Thorburn, A. W., Britt, K. L., Hewitt, K. N., Wreford, N. G., Proietto, J., Oz, O. K., Leury, B. J., Robertson, K. M., Yao, S., Simpson, E. R. Aromatase-deficient (ArKO) mice have a phenotype of increased adiposity. Proc. Nat. Acad. Sci. 97: 12735-12740, 2000. [PubMed: 11070087] [Full Text: https://doi.org/10.1073/pnas.97.23.12735]

  47. Justice, M. J., Siracusa, L. D., Gilbert, D. J., Heisterkamp, N., Groffen, J., Chada, K., Silan, C. M., Copeland, N. G., Jenkins, N. A. A genetic linkage map of mouse chromosome 10: localization of eighteen molecular markers using a single interspecific backcross. Genetics 125: 855-866, 1990. [PubMed: 1975791] [Full Text: https://doi.org/10.1093/genetics/125.4.855]

  48. Khosla, S., Riggs, B. L., Atkinson, E. J., Oberg, A. L., Mavilia, C., Del Monte, F., Melton, L. J., III, Brandi, M. L. Relationship of estrogen receptor genotypes to bone mineral density and to rates of bone loss in men. J. Clin. Endocr. Metab. 89: 1808-1816, 2004. [PubMed: 15070949] [Full Text: https://doi.org/10.1210/jc.2003-031448]

  49. Kim, J., Park, U.-H., Moon, M., Um, S.-J., Kim, E.-J. Negative regulation of ER-alpha by a novel protein CAC1 through association with histone demethylase LSD1. FEBS Lett. 587: 17-22, 2013. [PubMed: 23178685] [Full Text: https://doi.org/10.1016/j.febslet.2012.10.054]

  50. Korach, K. S. Insights from the study of animals lacking functional estrogen receptor. Science 266: 1524-1527, 1994. [PubMed: 7985022] [Full Text: https://doi.org/10.1126/science.7985022]

  51. Kos, M., Denger, S., Reid, G., Gannon, F. Upstream open reading frames regulate the translation of the multiple mRNA variants of the estrogen receptor-alpha. J. Biol. Chem. 277: 37131-37138, 2002. [PubMed: 12147702] [Full Text: https://doi.org/10.1074/jbc.M206325200]

  52. Kos, M., Reid, G., Denger, S., Gannon, F. Minireview: genomic organization of the human ER-alpha gene promoter region. Molec. Endocr. 15: 2057-2063, 2001. [PubMed: 11731608] [Full Text: https://doi.org/10.1210/mend.15.12.0731]

  53. Kumar, R., Wang, R.-A., Mazumdar, A., Talukder, A. H., Mandal, M., Yang, Z., Bagheri-Yarmand, R., Sahin, A., Hortobagyi, G., Adam, L., Barnes, C. J., Vadlamudi, R. K. A naturally occurring MTA1 variant sequesters oestrogen receptor-alpha in the cytoplasm. Nature 418: 654-657, 2002. [PubMed: 12167865] [Full Text: https://doi.org/10.1038/nature00889]

  54. Kun, Y., How, L. C., Hoon, T. P., Bajic, V. B., Lam, T. S., Aggarwal, A., Sze, H. G., Bok, W. S., Yin, W. C., Tan, P. Classifying the estrogen receptor status of breast cancers by expression profiles reveals a poor prognosis subpopulation exhibiting high expression of the ERBB2 receptor. Hum. Molec. Genet. 12: 3245-3258, 2003. [PubMed: 14570715] [Full Text: https://doi.org/10.1093/hmg/ddg347]

  55. Lawson, J. S., Field, A. S., Champion, S., Tran, D., Ishikura, H., Trichopoulos, D. Low oestrogen receptor alpha expression in normal breast tissue underlies low breast cancer incidence in Japan. (Letter) Lancet 354: 1787-1788, 1999. [PubMed: 10577642] [Full Text: https://doi.org/10.1016/s0140-6736(99)04936-3]

  56. Lee, H., Kim, D.-W., Remedios, R., Anthony, T. E., Chang, A., Madisen, L., Zeng, H., Anderson, D. J. Scalable control of mounting and attack by Esr1+ neurons in the ventromedial hypothalamus. Nature 509: 627-632, 2014. [PubMed: 24739975] [Full Text: https://doi.org/10.1038/nature13169]

  57. Lee, K., Jessop, H., Suswillo, R., Zaman, G., Lanyon, L. Bone adaptation requires oestrogen receptor-alpha. Nature 424: 389 only, 2003. [PubMed: 12879058] [Full Text: https://doi.org/10.1038/424389a]

  58. Lehtimaki, T., Kunnas, T. A., Mattila, K. M., Perola, M., Penttila, A., Koivula, T., Karhunen, P. J. Coronary artery wall atherosclerosis in relation to the estrogen receptor 1 gene polymorphism: an autopsy study. J. Molec. Med. 80: 176-180, 2002. [PubMed: 11894143] [Full Text: https://doi.org/10.1007/s00109-001-0311-5]

  59. Li, L., Haynes, M. P., Bender, J. R. Plasma membrane localization and function of the estrogen receptor alpha variant (ER46) in human endothelial cells. Proc. Nat. Acad. Sci. 100: 4807-4812, 2003. [PubMed: 12682286] [Full Text: https://doi.org/10.1073/pnas.0831079100]

  60. Liu, Z.-H., Cheng, Z.-H., Gong, R.-J., Liu, H., Liu, D., Li, L.-S. Sex differences in estrogen receptor gene polymorphism and its association with lupus nephritis in Chinese. Nephron 90: 174-180, 2002. [PubMed: 11818702] [Full Text: https://doi.org/10.1159/000049039]

  61. Lonard, D. M., Nawaz, Z., Smith, C. L., O'Malley, B. W. The 26S proteasome is required for estrogen receptor-alpha and coactivator turnover and for efficient estrogen receptor-alpha transactivation. Molec. Cell 5: 939-948, 2000. [PubMed: 10911988] [Full Text: https://doi.org/10.1016/s1097-2765(00)80259-2]

  62. Lorentzon, M., Lorentzon, R., Backstrom, T., Nordstrom, P. Estrogen receptor gene polymorphism, but not estradiol levels, is related to bone density in healthy adolescent boys: a cross-sectional and longitudinal study. J. Clin. Endocr. Metab. 84: 4597-4601, 1999. [PubMed: 10599725] [Full Text: https://doi.org/10.1210/jcem.84.12.6238]

  63. Mader, S., Kumar, V., de Verneuil, H., Chambon, P. Three amino acids of the oestrogen receptor are essential to its ability to distinguish an oestrogen from a glucocorticoid-responsive element. Nature 338: 271-274, 1989. [PubMed: 2922054] [Full Text: https://doi.org/10.1038/338271a0]

  64. Mattila, K. J., Nieminen, M. S., Valtonen, V. V., Rasi, V. P., Kesaniemi, Y. A., Syrjala, S. L., Jungell, P. S., Isoluoma, M., Hietaniemi K., Jokinen, M. J., Huttunen, J. K. Association between dental health and acute myocardial infarction. Brit. Med. J. 298: 779-782, 1989. [PubMed: 2496855] [Full Text: https://doi.org/10.1136/bmj.298.6676.779]

  65. McGuire, W. L., Chamness, G. C., Fuqua, S. A. W. Estrogen receptor variants in clinical breast cancer. Molec. Endocr. 5: 1571-1577, 1991. [PubMed: 1779964] [Full Text: https://doi.org/10.1210/mend-5-11-1571]

  66. McGuire, W. L., Chamness, G. C., Fuqua, S. A. W. Abnormal estrogen receptor in clinical breast cancer. J. Steroid Biochem. Molec. Biol. 43: 243-247, 1992. [PubMed: 1525064] [Full Text: https://doi.org/10.1016/0960-0760(92)90214-4]

  67. 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. Molec. Endocr. 10: 1519-1526, 1996. [PubMed: 8961262] [Full Text: https://doi.org/10.1210/mend.10.12.8961262]

  68. Menasce, L. P., White, G. R. M., Harrison, C. J., Boyle, J. M. Localization of the estrogen receptor locus (ESR) to chromosome 6q25.1 by FISH and a simple post-FISH banding technique. Genomics 17: 263-265, 1993. [PubMed: 8406468] [Full Text: https://doi.org/10.1006/geno.1993.1320]

  69. Metivier, R., Penot, G., Hubner, M. R., Reid, G., Brand, H., Kos, M., Gannon, F. Estrogen receptor-alpha directs ordered, cyclical, and combinatorial recruitment of cofactors on a natural target promoter. Cell 115: 751-763, 2003. [PubMed: 14675539] [Full Text: https://doi.org/10.1016/s0092-8674(03)00934-6]

  70. Metzger, D., White, J. H., Chambon, P. The human oestrogen receptor functions in yeast. Nature 334: 31-36, 1988. [PubMed: 3290686] [Full Text: https://doi.org/10.1038/334031a0]

  71. Mohammed, H., Russell, I. A., Stark, R., Rueda, O. M., Hickey, T. E., Tarulli, G. A., Serandour, A. A., Birrell, S. N., Bruna, A., Saadi, A., Menon, S., Hadfield, J., and 12 others. Progesterone receptor modulates ER-alpha action in breast cancer. Nature 523: 313-317, 2015. Note: Erratum: Nature 523: 144 only, 2015. [PubMed: 26153859] [Full Text: https://doi.org/10.1038/nature14583]

  72. Murphy, L. C., Wang, M., Coutt, A., Dotzlaw, H. Novel mutations in the estrogen receptor messenger RNA in human breast cancers. J. Clin. Endocr. Metab. 81: 1420-1427, 1996. [PubMed: 8636345] [Full Text: https://doi.org/10.1210/jcem.81.4.8636345]

  73. Musatov, S., Chen, W., Pfaff, D. W., Mobbs, C. V., Yang, X.-J., Clegg, D. J., Kaplitt, M. G., Ogawa, S. Silencing of estrogen receptor-alpha in the ventromedial nucleus of hypothalamus leads to metabolic syndrome. Proc. Nat. Acad. Sci. 104: 2501-2506, 2007. [PubMed: 17284595] [Full Text: https://doi.org/10.1073/pnas.0610787104]

  74. Nakamura, T., Imai, Y., Matsumoto, T., Sato, S., Takeuchi, K., Igarashi, K., Harada, Y., Azuma, Y., Krust, A., Yamamoto, Y., Nishina, H., Takeda, S., Takayanagi, H., Metzger, D., Kanno, J., Takaoka, K., Martin, T. J., Chambon, P., Kato, S. Estrogen prevents bone loss via estrogen receptor alpha and induction of Fas ligand in osteoclasts. Cell 130: 811-823, 2007. [PubMed: 17803905] [Full Text: https://doi.org/10.1016/j.cell.2007.07.025]

  75. Nelson, E. R., Wardell, S. E., Jasper, J. S., Park, S., Suchindran, S., Howe, M. K., Carver, N. J., Pillai, R. V., Sullivan, P. M., Sondhi, V., Umetani, M., Geradts, J., McDonnell, D. P. 27-hydroxycholesterol links hypercholesterolemia and breast cancer pathophysiology. Science 342: 1094-1098, 2013. [PubMed: 24288332] [Full Text: https://doi.org/10.1126/science.1241908]

  76. Ohtake, F., Takeyama, K., Matsumoto, T., Kitagawa, H., Yamamoto, Y., Nohara, K., Tohyama, C., Krust, A., Mimura, J., Chambon, P., Yanagisawa, J., Fujii-Kuriyama, Y., Kato, S. Modulation of oestrogen receptor signalling by association with the activated dioxin receptor. Nature 423: 545-550, 2003. [PubMed: 12774124] [Full Text: https://doi.org/10.1038/nature01606]

  77. Park, C. J., Zhao, Z., Glidewell-Kenney, C., Lazic, M., Chambon, P., Krust, A., Weiss, J., Clegg, D. J., Dunaif, A., Jameson, J. L., Levine, J. E. Genetic rescue of nonclassical ER-alpha signaling normalizes energy balance in obese Er-alpha-null mutant mice. J. Clin. Invest. 121: 604-612, 2011. [PubMed: 21245576] [Full Text: https://doi.org/10.1172/JCI41702]

  78. Pedram, A., Razandi, M., Deschenes, R. J., Levin, E. R. DHHC-7 and -21 are palmitoylacyltransferases for sex steroid receptors. Molec. Biol. Cell 23: 188-199, 2012. [PubMed: 22031296] [Full Text: https://doi.org/10.1091/mbc.E11-07-0638]

  79. Pelletier, G., El-Alfy, M. Immunocytochemical localization of estrogen receptors alpha and beta in the human reproductive organs. J. Clin. Endocr. Metab. 85: 4835-4840, 2000. [PubMed: 11134151] [Full Text: https://doi.org/10.1210/jcem.85.12.7029]

  80. Perillo, B., Ombra, M. N., Bertoni, A., Cuozzo, C., Sacchetti, S., Sasso, A., Chiariotti, L., Malorni, A., Abbondanza, C., Avvedimento, E. V. DNA oxidation as triggered by H3K9me2 demethylation drives estrogen-induced gene expression. Science 319: 202-206, 2008. [PubMed: 18187655] [Full Text: https://doi.org/10.1126/science.1147674]

  81. Ponglikitmongkol, M., Green, S., Chambon, P. Genomic organization of the human oestrogen receptor gene. EMBO J. 7: 3385-3388, 1988. [PubMed: 3145193] [Full Text: https://doi.org/10.1002/j.1460-2075.1988.tb03211.x]

  82. Quaynor, S. D., Stradtman, E. W., Jr., Kim, H.-G., Shen, Y., Chorich, L. P., Schreihofer, D. A., Layman, L. C. Delayed puberty and estrogen resistance in a woman with estrogen receptor alpha variant. New Eng. J. Med. 369: 164-171, 2013. [PubMed: 23841731] [Full Text: https://doi.org/10.1056/NEJMoa1303611]

  83. Reese, J. C., Katzenellenbogen, B. S. Mutagenesis of cysteines in the hormone binding domain of the human estrogen receptor: alterations in binding and transcriptional activation by covalently and reversibly attaching ligands. J. Biol. Chem. 266: 10880-10887, 1991. [PubMed: 2040605]

  84. Reese, J. C., Katzenellenbogen, B. S. Characterization of a temperature-sensitive mutation in the hormone binding domain of the human estrogen receptor: studies in cell extracts and intact cells and their implications for hormone-dependent transcriptional activation. J. Biol. Chem. 267: 9868-9873, 1992. [PubMed: 1577818]

  85. Reid, G., Hubner, M. R., Metivier, R., Brand, H., Denger, S., Manu, D., Beaudouin, J., Ellenberg, J., Gannon, F. Cyclic, proteasome-mediated turnover of unliganded and liganded ER-alpha on responsive promoters is an integral feature of estrogen signaling. Molec. Cell 11: 695-707, 2003. [PubMed: 12667452] [Full Text: https://doi.org/10.1016/s1097-2765(03)00090-x]

  86. Reis-Filho, J. S., Drury, S., Lambros, M. B., Marchio, C., Johnson, N., Natrajan, R., Salter, J., Levey, P., Fletcher, O., Peto, J., Ashworth, A., Dowsett, M. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 809-810, 2008. [PubMed: 18583966] [Full Text: https://doi.org/10.1038/ng0708-809b]

  87. Robinson, D. R., Wu, Y.-M., Vats, P., Su, F., Lonigro, R. J., Cao, X., Kalyana-Sundaram, S., Wang, R., Ning, Y., Hodges, L., Gursky, A., Siddiqui, J., and 13 others. Activating ESR1 mutations in hormone-resistant metastatic breast cancer. Nature Genet. 45: 1446-1451, 2013. [PubMed: 24185510] [Full Text: https://doi.org/10.1038/ng.2823]

  88. Schuit, S. C. E., Oei, H.-H. S., Witteman, J. C. M., Geurts van Kessel, C. H., van Meurs, J. B. J., Nijhuis, R. L., van Leeuwen, J. P. T. M., de Jong, F. H., Zillikens, M. C., Hofman, A., Pols, H. A. P., Uitterlinden, A. G. Estrogen receptor alpha gene polymorphisms and risk of myocardial infarction. JAMA 291: 2969-2977, 2004. [PubMed: 15213208] [Full Text: https://doi.org/10.1001/jama.291.24.2969]

  89. Scott, G. K., Kushner, P., Vigne, J.-L., Benz, C. C. Truncated forms of DNA-binding estrogen receptors in human breast cancer. J. Clin. Invest. 88: 700-706, 1991. [PubMed: 1864980] [Full Text: https://doi.org/10.1172/JCI115356]

  90. Shang, Y., Brown, M. Molecular determinants for the tissue specificity of SERMs. Science 295: 2465-2468, 2002. [PubMed: 11923541] [Full Text: https://doi.org/10.1126/science.1068537]

  91. Shearman, A. M., Cupples, L. A., Demissie, S., Peter, I., Schmid, C. H., Karas, R. H., Mendelsohn, M. E., Housman, D. E., Levy, D. Association between estrogen receptor alpha gene variation and cardiovascular disease. JAMA 290: 2263-2270, 2003. Note: Erratum: JAMA 291: 186 only, 2004. [PubMed: 14600184] [Full Text: https://doi.org/10.1001/jama.290.17.2263]

  92. Shearman, A. M., Demissie, S., Cupples, L. A., Peter, I., Schmid, C. H., Ordovas, J. M., Mendelsohn, M. E., Housman, D. E. Tobacco smoking, estrogen receptor alpha gene variation and small low density lipoprotein level. Hum. Molec. Genet. 14: 2405-2413, 2005. [PubMed: 16014638] [Full Text: https://doi.org/10.1093/hmg/ddi242]

  93. 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 95: 927-937, 1998. [PubMed: 9875847] [Full Text: https://doi.org/10.1016/s0092-8674(00)81717-1]

  94. Shim, G.-J., Kis, L. L., Warner, M., Gustafsson, J.-A. Autoimmune glomerulonephritis with spontaneous formation of splenic germinal centers in mice lacking the estrogen receptor alpha gene. Proc. Nat. Acad. Sci. 101: 1720-1724, 2004. Note: Erratum: Proc. Nat. Acad. Sci. 103: 8298 only, 2006. [PubMed: 14745006] [Full Text: https://doi.org/10.1073/pnas.0307915100]

  95. Shupnik, M. A., Pitt, L. K., Soh, A. Y., Anderson, A., Lopes, M. B., Laws, E. R., Jr. Selective expression of estrogen receptor alpha and beta isoforms in human pituitary tumors. J. Clin. Endocr. Metab. 83: 3965-3972, 1998. [PubMed: 9814476] [Full Text: https://doi.org/10.1210/jcem.83.11.5236]

  96. Simoncini, T., Hafezi-Moghadam, A., Brazil, D. P., Ley, K., Chin, W. W., Liao, J. K. Interaction of oestrogen receptor with the regulatory subunit of phosphatidylinositol-3-OH kinase. Nature 407: 538-541, 2000. [PubMed: 11029009] [Full Text: https://doi.org/10.1038/35035131]

  97. Sluyser, M., Mester, J. Oncogenes homologous to steroid receptors? (Letter) Nature 315: 546 only, 1985. [PubMed: 2989691] [Full Text: https://doi.org/10.1038/315546a0]

  98. Sluyser, M. Mutations in the estrogen receptor gene. Hum. Mutat. 6: 97-103, 1995. [PubMed: 7581411] [Full Text: https://doi.org/10.1002/humu.1380060202]

  99. 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. New Eng. J. Med. 331: 1056-1061, 1994. Note: Erratum: New Eng. J. Med. 332: 131 only, 1995. [PubMed: 8090165] [Full Text: https://doi.org/10.1056/NEJM199410203311604]

  100. Sowers, M., Jannausch, M. L., Liang, W., Willing, M. Estrogen receptor genotypes and their association with the 10-year changes in bone mineral density and osteocalcin concentrations. J. Clin. Endocr. Metab. 89: 733-739, 2004. [PubMed: 14764789] [Full Text: https://doi.org/10.1210/jc.2003-030691]

  101. Stoica, A., Katzenellenbogen, B. S., Martin, M. B. Activation of estrogen receptor-alpha by the heavy metal cadmium. Molec. Endocr. 14: 545-553, 2000. [PubMed: 10770491] [Full Text: https://doi.org/10.1210/mend.14.4.0441]

  102. 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., Bagger, Y., Gulcher, J. R., Eisman, J. A., Christiansen, C., Sigurdsson, G., Kong, A., Thorsteinsdottir, U., Stefansson, K. Multiple genetic loci for bone mineral density and fractures. New Eng. J. Med. 358: 2355-2365, 2008. [PubMed: 18445777] [Full Text: https://doi.org/10.1056/NEJMoa0801197]

  103. Taguchi, A., Kobayashi, J., Suei, Y., Ohtsuka, M., Tanimoto, K., Sanada, M., Ohama, K. Relationship between estrogen-receptor genotype and tooth loss in postmenopausal women. (Letter) JAMA 286: 2234-2235, 2001. [PubMed: 11710886] [Full Text: https://doi.org/10.1001/jama.286.18.2234]

  104. Taguchi, A. Estrogen receptor alpha polymorphisms and the risk of myocardial infarction. (Letter) JAMA 292: 1683 only, 2004. [PubMed: 15479929] [Full Text: https://doi.org/10.1001/jama.292.14.1683-a]

  105. Takeyama, J., Suzuki, T., Inoue, S., Kaneko, C., Nagura, H., Harada, N., Sasano, H. Expression and cellular localization of estrogen receptors alpha and beta in the human fetus. J. Clin. Endocr. Metab. 86: 2258-2262, 2001. [PubMed: 11344236] [Full Text: https://doi.org/10.1210/jcem.86.5.7447]

  106. Tobias, J. H., Steer, C. D., Vilarino-Guell, C., Brown, M. A. Estrogen receptor alpha regulates area-adjusted bone mineral content in late pubertal girls. J. Clin. Endocr. Metab. 92: 641-647, 2007. [PubMed: 17105837] [Full Text: https://doi.org/10.1210/jc.2006-1555]

  107. Toy, W., Shen, Y., Won, H., Green, B., Sakr, R. A., Will, M., Li, Z., Gala, K., Fanning, S., King, T. A., Hudis, C., Chen, D., Taran, T., Hortobagyi, G., Greene, G., Berger, M., Baselga, J., Chandarlapaty, S. ESR1 ligand-binding domain mutations in hormone-resistant breast cancer. Nature Genet. 45: 1439-1445, 2013. [PubMed: 24185512] [Full Text: https://doi.org/10.1038/ng.2822]

  108. van Meurs, J. B. J., Schuit, S. C. E., Weel, A. E. A. M., van der Klift, M., Bergink, A. P., Arp, P. P., Colin, E. M., Fang, Y., Hofman, A., van Duijn, C. M., van Leeuwen, J. P. T. M., Pols, H. A. P., Uitterlinden, A. G. Association of 5-prime estrogen receptor alpha gene polymorphisms with bone mineral density, vertebral bone area and fracture risk. Hum. Molec. Genet. 12: 1745-1754, 2003. [PubMed: 12837697] [Full Text: https://doi.org/10.1093/hmg/ddg176]

  109. Vincent-Salomon, A., Raynal, V., Lucchesi, C., Gruel, N., Delattre, O. ESR1 gene amplification in breast cancer: a common phenomenon? (Letter) Nature Genet. 40: 809 only, 2008. [PubMed: 18583967] [Full Text: https://doi.org/10.1038/ng0708-809a]

  110. Wada-Hiraike, O., Yano, T., Nei, T., Matsumoto, Y., Nagasaka, K., Takizawa, S., Oishi, H., Arimoto, T., Nakagawa, S., Yasugi, T., Kato, S., Taketani, Y. The DNA mismatch repair gene hMSH2 is a potent coactivator of oestrogen receptor-alpha. Brit. J. Cancer 92: 2286-2291, 2005. [PubMed: 15886699] [Full Text: https://doi.org/10.1038/sj.bjc.6602614]

  111. Walter, P., Green, S., Greene, G., Krust, A., Bornert, J.-M., Jeltsch, J.-M., Staub, A., Jensen, E., Scrace, G., Waterfield, M., Chambon, P. Cloning of the human estrogen receptor cDNA. Proc. Nat. Acad. Sci. 82: 7889-7893, 1985. [PubMed: 3865204] [Full Text: https://doi.org/10.1073/pnas.82.23.7889]

  112. Wang, X., Yang, Z., Zhang, H., Ding, L., Li, X., Zhu, C., Zheng, Y., Ye, Q. The estrogen receptor-interacting protein HPIP increases estrogen-responsive gene expression through activation of MAPK and AKT. Biochim. Biophys. Acta 1783: 1220-1228, 2008. [PubMed: 18302941] [Full Text: https://doi.org/10.1016/j.bbamcr.2008.01.026]

  113. Wei, X., Xu, H., Kufe, D. MUC1 oncoprotein stabilizes and activates estrogen receptor alpha. Molec. Cell 21: 295-305, 2006. [PubMed: 16427018] [Full Text: https://doi.org/10.1016/j.molcel.2005.11.030]

  114. Weis, K. E., Ekena, K., Thomas, J. A., Lazennec, G., Katzenellenbogen, B. S. Constitutively active human estrogen receptors containing amino acid substitutions for tyrosine 537 in the receptor protein. Molec. Endocr. 10: 1388-1398, 1996. [PubMed: 8923465] [Full Text: https://doi.org/10.1210/mend.10.11.8923465]

  115. Ye, Q., Chung, L. W. K., Li, S., Zhau, H. E. Identification of a novel FAS/ER-alpha fusion transcript expressed in human cancer cells. Biochim. Biophys. Acta 1493: 373-377, 2000. [PubMed: 11018265] [Full Text: https://doi.org/10.1016/s0167-4781(00)00202-5]

  116. Zhao, C., Koide, A., Abrams, J., Deighton-Collins, S., Martinez, A., Schwartz, J. A., Koide, S., Skafar, D. F. Mutation of Leu-536 in human estrogen receptor-alpha alters the coupling between ligand binding, transcription activation, and receptor conformation. J. Biol. Chem. 278: 27278-27286, 2003. [PubMed: 12736255] [Full Text: https://doi.org/10.1074/jbc.M303840200]

  117. Zuppan, P., Hall, J. M., Lee, M. K., Ponglikitmongkol, M., King, M.-C. Possible linkage of the estrogen receptor gene to breast cancer in a family with late-onset disease. Am. J. Hum. Genet. 48: 1065-1068, 1991. [PubMed: 2035527]

  118. Zuppan, P. J., Hall, J. M., Ponglikitmongkol, M., Spielman, R., King, M. C. Polymorphisms at the estrogen receptor (ESR) locus and linkage relationships on chromosome 6q. (Abstract) Cytogenet. Cell Genet. 51: 1116 only, 1989.


Contributors:
Bao Lige - updated : 03/18/2020
Bao Lige - updated : 01/31/2020
Marla J. F. O'Neill - updated : 12/03/2018
Ada Hamosh - updated : 11/30/2015
Matthew B. Gross - updated : 11/3/2015
Ada Hamosh - updated : 8/26/2014
Ada Hamosh - updated : 1/9/2014
Ada Hamosh - updated : 12/19/2013
Marla J. F. O'Neill - updated : 11/7/2013
Marla J. F. O'Neill - updated : 8/8/2013
Patricia A. Hartz - updated : 4/19/2013
Patricia A. Hartz - updated : 4/27/2012
Patricia A. Hartz - updated : 3/9/2011
Ada Hamosh - updated : 12/29/2009
John A. Phillips, III - updated : 5/7/2009
Patricia A. Hartz - updated : 3/26/2009
Ada Hamosh - updated : 1/6/2009
Ada Hamosh - updated : 8/6/2008
Ada Hamosh - updated : 6/10/2008
Patricia A. Hartz - updated : 5/29/2008
Ada Hamosh - updated : 3/25/2008
Patricia A. Hartz - updated : 2/7/2008
Marla J. F. O'Neill - updated : 1/21/2008
John A. Phillips, III - updated : 12/19/2007
Jane Kelly - updated : 11/27/2007
Victor A. McKusick - updated : 5/24/2007
Marla J. F. O'Neill - updated : 4/19/2007
Patricia A. Hartz - updated : 6/2/2006
Patricia A. Hartz - updated : 2/9/2006
Marla J. F. O'Neill - updated : 1/20/2006
George E. Tiller - updated : 1/10/2006
Patricia A. Hartz - updated : 11/9/2005
Marla J. F. O'Neill - updated : 8/30/2005
John A. Phillips, III - updated : 8/1/2005
Marla J. F. O'Neill - updated : 5/19/2005
Marla J. F. O'Neill - updated : 5/10/2005
George E. Tiller - updated : 5/6/2005
John A. Phillips, III - updated : 4/11/2005
Victor A. McKusick - updated : 1/25/2005
Victor A. McKusick - updated : 12/29/2004
John A. Phillips, III - updated : 10/14/2004
Victor A. McKusick - updated : 5/12/2004
Victor A. McKusick - updated : 12/8/2003
Patricia A. Hartz - updated : 10/10/2003
Ada Hamosh - updated : 8/4/2003
Ada Hamosh - updated : 7/24/2003
Victor A. McKusick - updated : 6/6/2003
Ada Hamosh - updated : 5/29/2003
Stylianos E. Antonarakis - updated : 4/15/2003
Patricia A. Hartz - updated : 1/30/2003
Ada Hamosh - updated : 10/18/2002
Ada Hamosh - updated : 9/20/2002
John A. Phillips, III - updated : 7/29/2002
John A. Phillips, III - updated : 7/26/2002
Victor A. McKusick - updated : 4/5/2002
Ada Hamosh - updated : 4/2/2002
John A. Phillips, III - updated : 10/9/2001
John A. Phillips, III - updated : 9/21/2001
John A. Phillips, III - updated : 7/9/2001
Paul J. Converse - updated : 2/1/2001
Victor A. McKusick - updated : 11/30/2000
John A. Phillips, III - updated : 11/13/2000
George E. Tiller - updated : 10/27/2000
Ada Hamosh - updated : 10/18/2000
John A. Phillips, III - updated : 9/29/2000
Stylianos E. Antonarakis - updated : 8/2/2000
Jane Kelly - updated : 6/22/2000
Wilson H. Y. Lo - updated : 3/21/2000
Ada Hamosh - updated : 2/8/2000
John A. Phillips, III - updated : 9/29/1999
Ada Hamosh - updated : 5/20/1999
John A. Phillips, III - updated : 4/15/1999
John A. Phillips, III - updated : 3/18/1999
Stylianos E. Antonarakis - updated : 2/4/1999
Paul Brennan - updated : 1/16/1998
Victor A. McKusick - updated : 6/23/1997
John A. Phillips, III - updated : 3/20/1997
John A. Phillips, III - updated : 2/25/1997
John A. Phillips, III - updated : 1/8/1997

Creation Date:
Victor A. McKusick : 6/4/1986

Edit History:
alopez : 02/12/2021
mgross : 03/18/2020
mgross : 01/31/2020
carol : 10/01/2019
carol : 09/27/2019
carol : 09/18/2019
carol : 12/03/2018
carol : 11/14/2017
carol : 11/13/2017
alopez : 11/30/2015
alopez : 11/30/2015
mgross : 11/3/2015
mgross : 11/3/2015
carol : 8/14/2015
alopez : 8/26/2014
alopez : 1/9/2014
alopez : 12/19/2013
alopez : 11/8/2013
alopez : 11/8/2013
mcolton : 11/7/2013
carol : 8/9/2013
carol : 8/9/2013
tpirozzi : 8/8/2013
tpirozzi : 8/8/2013
tpirozzi : 8/8/2013
mgross : 4/19/2013
carol : 2/26/2013
mgross : 2/5/2013
carol : 1/2/2013
terry : 10/10/2012
mgross : 4/27/2012
terry : 2/15/2012
terry : 1/13/2012
carol : 11/2/2011
ckniffin : 4/8/2011
mgross : 3/9/2011
terry : 3/9/2011
alopez : 3/1/2011
alopez : 3/1/2011
wwang : 7/1/2010
terry : 6/23/2010
alopez : 6/22/2010
alopez : 1/6/2010
terry : 12/29/2009
mgross : 12/10/2009
terry : 12/3/2009
terry : 6/3/2009
alopez : 5/7/2009
mgross : 3/26/2009
mgross : 3/26/2009
alopez : 3/11/2009
mgross : 2/4/2009
carol : 2/2/2009
wwang : 1/12/2009
alopez : 1/7/2009
terry : 1/6/2009
terry : 9/25/2008
alopez : 8/28/2008
terry : 8/6/2008
alopez : 6/16/2008
terry : 6/10/2008
mgross : 6/2/2008
terry : 5/29/2008
alopez : 3/25/2008
mgross : 2/8/2008
mgross : 2/8/2008
terry : 2/7/2008
carol : 1/21/2008
carol : 12/26/2007
carol : 12/19/2007
carol : 11/27/2007
carol : 11/27/2007
alopez : 6/6/2007
terry : 5/24/2007
wwang : 4/25/2007
wwang : 4/19/2007
wwang : 11/8/2006
mgross : 7/5/2006
terry : 6/27/2006
mgross : 6/8/2006
terry : 6/2/2006
mgross : 2/24/2006
terry : 2/9/2006
terry : 2/3/2006
wwang : 1/27/2006
wwang : 1/20/2006
wwang : 1/20/2006
terry : 1/10/2006
wwang : 11/28/2005
terry : 11/9/2005
wwang : 8/30/2005
alopez : 8/1/2005
wwang : 5/19/2005
wwang : 5/10/2005
tkritzer : 5/6/2005
alopez : 4/11/2005
tkritzer : 1/31/2005
terry : 1/25/2005
tkritzer : 12/30/2004
terry : 12/29/2004
alopez : 10/14/2004
alopez : 7/27/2004
tkritzer : 5/18/2004
terry : 5/12/2004
tkritzer : 12/9/2003
terry : 12/8/2003
ckniffin : 10/15/2003
mgross : 10/10/2003
mgross : 10/10/2003
carol : 8/28/2003
alopez : 8/5/2003
terry : 8/4/2003
tkritzer : 7/25/2003
terry : 7/24/2003
carol : 6/19/2003
tkritzer : 6/17/2003
terry : 6/6/2003
mgross : 5/30/2003
terry : 5/29/2003
mgross : 4/15/2003
mgross : 2/4/2003
terry : 1/30/2003
mgross : 10/18/2002
mgross : 10/18/2002
terry : 10/18/2002
alopez : 9/20/2002
carol : 9/18/2002
tkritzer : 9/13/2002
tkritzer : 9/6/2002
tkritzer : 7/29/2002
tkritzer : 7/26/2002
cwells : 4/15/2002
cwells : 4/5/2002
cwells : 4/5/2002
cwells : 4/3/2002
terry : 4/2/2002
cwells : 10/9/2001
cwells : 10/9/2001
cwells : 10/5/2001
cwells : 9/26/2001
cwells : 9/21/2001
alopez : 9/21/2001
carol : 9/13/2001
alopez : 8/13/2001
alopez : 8/13/2001
alopez : 8/13/2001
alopez : 7/9/2001
cwells : 2/6/2001
cwells : 2/1/2001
mgross : 12/4/2000
terry : 11/30/2000
terry : 11/13/2000
mcapotos : 11/1/2000
mcapotos : 10/27/2000
alopez : 10/18/2000
mgross : 10/3/2000
terry : 9/29/2000
mcapotos : 8/10/2000
mgross : 8/2/2000
alopez : 6/22/2000
carol : 3/21/2000
carol : 3/21/2000
alopez : 2/8/2000
mgross : 9/29/1999
alopez : 5/20/1999
terry : 5/20/1999
mgross : 4/15/1999
mgross : 3/23/1999
mgross : 3/18/1999
mgross : 3/17/1999
carol : 2/4/1999
dkim : 9/11/1998
alopez : 7/6/1998
terry : 6/30/1998
carol : 1/16/1998
terry : 8/7/1997
terry : 8/5/1997
mark : 6/26/1997
jenny : 6/23/1997
jenny : 5/28/1997
jenny : 5/27/1997
jenny : 3/20/1997
mark : 3/18/1997
jenny : 2/25/1997
carol : 11/12/1996
mark : 10/16/1996
terry : 10/31/1995
mark : 10/2/1995
carol : 1/5/1995
pfoster : 4/21/1994
warfield : 4/20/1994
carol : 10/26/1993