Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

Contents lists available at ScienceDirect

Journal of Steroid Biochemistry and Molecular Biology


journal homepage: www.elsevier.com/locate/jsbmb

Review

Estrogens and breast cancer: Mechanisms involved in obesity-related T


development, growth and progression
Priya Bhardwaja,b,1, CheukMan C. Aua,1, Alberto Benito-Martina, Heta Ladumora,c,

Sofya Oshchepkovaa, Ruth Mogesa, Kristy A. Browna,b,d,
a
Department of Medicine, Weill Cornell Medicine, New York, USA
b
Graduate School of Medical Sciences, Weill Cornell Medicine, New York, USA
c
Weill Cornell Medicine - Qatar, Doha, Qatar
d
Department of Physiology, Monash University, Clayton, Victoria, Australia

A R T I C LE I N FO A B S T R A C T

Keywords: Obesity is a risk factor for estrogen receptor-positive (ER+) breast cancer after menopause. The pro-proliferative
Estrogen effects of estrogens are well characterized and there is a growing body of evidence to also suggest an important
Breast cancer role in tumorigenesis. Importantly, obesity not only increases the risk of breast cancer, but it also increases the
Obesity risk of recurrence and cancer-associated death. Aromatase is the rate-limiting enzyme in estrogen biosynthesis
Aromatase
and its expression in breast adipose stromal cells is hypothesized to drive the growth of breast tumors and confer
Tumorigenesis
resistance to endocrine therapy in obese postmenopausal women. The molecular regulation of aromatase has
Metastasis
Endocrine therapy been characterized in response to many obesity-related molecules, including inflammatory mediators and adi-
Endocrine resistance pokines. This review is aimed at providing an overview of our current knowledge in relation to the regulation of
estrogens in adipose tissue and their role in driving breast tumor development, growth and progression.

1. Sources of estrogens in pre- and postmenopausal women which is the precursor to all adrenal and gonadal steroid hormones [5].
The first process in steroidogenesis is the transport of cholesterol to the
Estrogens play an important role in a number of physiological inner mitochondrial membrane by the steroidogenic acute regulator
processes, including regulating energy metabolism, stress responses, (StAR). Next, cholesterol is converted to pregnenolone by the cyto-
mineral balance, as well as sexual development [1]. In premenopausal chrome P450 side-chain cleavage enzyme. The formation of the tes-
women, estrogens are predominantly produced by the ovary [2]. The tosterone precursor androstenedione from pregnenolone is dependent
hypothalamus releases gonadotropin-releasing hormone (GnRH), which on the action of 3β-HSD to produce progesterone and CYP17A1, which
stimulates the secretion of follicle-stimulating hormone (FSH) and lu- converts progesterone to androstenedione via a two-step mechanism.
teinizing hormone (LH). FSH stimulates the biosynthesis of estrogens in Androstenedione is then converted to testosterone by 17β-HSD en-
growing ovarian follicles, which then act on the hypothalamus to in- zymes, and can then be aromatized to estradiol (17β-estradiol/E2). In
duce the production of LH. An acute rise in LH triggers ovulation and postmenopausal women, however, it is circulating dehydroepian-
the development of the corpus luteum. After menopause, the ovaries drosterone sulfate (DHEA-S) from the adrenals that is the source of
produce negligible levels of estrogens. The importance of gonadal androgen for estrogen formation at peripheral sites. The local bio-
steroidogenesis in normal breast development and in the origin of synthesis of estrogens within the breast [6,7] and circulating levels of
breast cancer is emphasized by the fact that early menstruation and late estrogens in blood [8,9], believed to be a reflection of adipose-derived
menopause are linked to a higher risk of breast cancer [3]. Similarly, steroid production, are directly associated with driving breast tumor
late menarche and early menopause (before the age of 40) result in a cell proliferation [10]. The intracrinology that occurs in the breast as a
significant reduction in the risk of developing breast cancer [4]. It is result of the complex interaction of enzymes responsible for the acti-
somewhat paradoxical, therefore, that the majority of breast cancers vation and inactivation of steroid hormones has been the focus of many
occur after menopause, when circulating estrogen levels are low. studies to explain the increased risk of breast cancer after menopause,
The de novo biosynthesis of sex hormones necessitates cholesterol, when gonadal estrogen biosynthesis has ceased [11,12]. Specifically,


Corresponding author at: 1300 York Avenue, Room E-804, New York, NY 10065, USA.
E-mail address: kab2060@med.cornell.edu (K.A. Brown).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.jsbmb.2019.03.002
Received 7 January 2019; Received in revised form 27 February 2019; Accepted 1 March 2019
Available online 06 March 2019
0960-0760/ © 2019 Elsevier Ltd. All rights reserved.
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

the breast expresses all enzymes required for the conversion of DHEA-S which is elevated in obesity, inhibits p53 resulting in elevation in ar-
to E2, including steroid sulfatase, 3β-HSD, 17β-HSD1 and aromatase omatase [43]. PGE2 was also shown to stabilize HIF1α, leading to the
[13,14]. Of these enzymes, the best characterized in terms of its reg- binding and stimulation of aromatase PII [44]. More recently, Sub-
ulation in obesity is the enzyme involved in the rate-limiting step in baramaiah et al. found that PGE2 downregulates SIRT1 in a human ASC
estrogen biosynthesis, aromatase. cell line leading to the upregulation of HIF1α [45]. Further supporting
the role of PGE2 in induction of aromatase, IL-6 in sera from obese
2. Aromatase subjects was found to induce breast cancer cell PGE2 secretion, which in
turn induced aromatase expression in primary ASCs. This effect was
Cytochrome P450 aromatase (P450arom) is a microsomal enzyme nullified by both depletion of IL-6 from sera or treatment with cel-
that is expressed in the endoplasmic reticulum and catalyzes one of the ecoxib, an inhibitor of the enzyme COX-2 which catalyzes the conver-
final steps in estrogen biosynthesis by converting 19-carbon steroids sion of arachidonic acid to PGE2 [46]. While previous findings de-
(androgens, e.g. androstenedione and testosterone) to 18-carbon ster- monstrated that IL-6 can increase the activity of promoter I.4 in the
oids (estrogens, e.g. estrone and estradiol) [15]. Aromatase is found in presence of the IL-6 soluble receptor [18], these findings provide a new
many tissues, including the gonads, brain, adipose tissue, placenta, level of complexity regarding the role of IL-6 in regulating aromatase
blood vessels, skin, bone and in breast cancer tissue [16]. Its expression expression in ASCs. Adipokines have also been examined for their role
in breast adipose is hypothesized to be a major driver of estrogen-de- in regulating aromatase. Leptin, an adipokine increased in obesity, in-
pendent breast cancer after menopause. The aromatase (CYP19A1) hibits p53 in human breast derived ASCs, leading to an increase in
gene is located on chromosome 15q21.2 and is approximately 123 kb aromatase expression [47]. This finding complemented prior work de-
long with nine coding exons (II-X) and a 93 kb regulatory region. monstrating that PGE2 and leptin stimulate aromatase expression by
Eight tissue-specific promoters regulate the expression of the suppressing the activity of energy sensors LKB1/AMPK, thereby alle-
CYP19A1 gene yielding transcripts with unique 5′-untranslated regions viating their suppressive effects on CREB-regulated transcriptional co-
[17]. These are promoters I.1 (placenta major, ≈ 93 kb), I.2a (placenta activators (CRTCs) which stimulate aromatase [48,49]. Interestingly,
minor, ≈ 78 kb), I.4 (skin, adipose tissue and bone, ≈ 73 kb), I.7 LKB1 and AMPK are stimulated by adiponectin, an adipokine produced
(endothelial cell and breast cancer, ≈ 36 kb), I.f (brain, ≈ 33 kb), I.6 by healthy adipocytes, leading to suppression of the PII-specific ex-
(bone, ≈ 0.7 kb), I.3 (adipose tissue and breast cancer, ≈ 0.2 kb) and II pression of aromatase. This suggests that AMPK-activating drugs may
(ovary, adipose tissue, breast cancer and endometriosis, within 1 kb) selectively inhibit aromatase in adipose tissue, including breast. More
[15,17]. In normal breast adipose tissue, low levels of aromatase are recently, the orexigenic hormone ghrelin and its unacylated form, des-
derived from activation of the distal promoter I.4, known to be regu- acyl ghrelin, which are reduced in obesity, were also shown to suppress
lated by glucocorticoids and class 1 cytokines, such as interleukin 6 (IL- PII-driven aromatase expression mediated via suppression of cAMP
6), interleukin 11(IL-11), leukemia inhibitory factor (LIF) and oncos- [50].
tatin M (OSM), via the Janus kinase-1/signal transducer and activator The regulation of PI.4 has also been examined in the context of
of transcription 3 (JAK1/STAT3) pathway [18]. Promoter I.4 is also obesity. Promoter I.4-specific transcripts, present in adipose tissue, can
activated by the inflammatory mediator tumor necrosis factor alpha be stimulated by inflammatory mediators, including IL-6, IL-11, leu-
(TNFα), via the mitogen-activated protein kinase (MAPK)-AP1 pathway kemia inhibitory factor, oncostatin M, as well as TNFα [51,52]. Using
[18]. However, in the breast adipose tissue of women with breast preclinical models and human breast tissue, the TNFα-mediated in-
cancer, the majority of transcripts are derived from the coordinated duction of aromatase was shown to require ERK1/2 activation, an effect
activation of promoters I.3 and II in a cAMP-dependent manner that was blocked by the anti-inflammatory cytokine IL-10 [53]. Taken
[18,19]. together, these studies propose new mechanisms that explain the ele-
vation in estrogen produced by breast adipose tissue in obese post-
2.1. Aromatase regulation in obesity menopausal women.
It is important to highlight that since aromatase is produced by
The prevalence of obesity has been steadily rising worldwide, and dysfunctional breast adipose tissue, obesity as defined by BMI may not
there is now strong evidence to support a causal link between obesity be the best predictor of local estrogen levels since it does not account
and the development of many cancers, including breast, ovarian, renal, for volume of body fat or quality of fat. For example, a recent study
pancreatic, leukemia, multiple myeloma, and esophageal cancers found that breast white adipose tissue inflammation and other systemic
[20–25]. For breast cancer, the link is strongest for postmenopausal correlates of metabolic syndrome (e.g. leptin) were strongly correlated
women and for the development of ER+ breast cancer, suggesting an with aromatase expression and activity in women with a normal BMI
important role of estrogens in driving obesity-associated breast cancer (≤24.9 kg/m2) [54]. In a prospective study of 12,159 postmenopausal
growth and development [20,24]. After menopause, adipose tissue is women in Sweden, body fat % was a better predictor of breast cancer
the primary source of estrogen production in the body [26–28]. Inter- incidence than BMI [55]. It is possible that this discrepancy is due to
estingly, body mass index (BMI) is found to be positively associated body fat % being a superior readout of estrogen levels which are closely
with tissue levels of estrogens [29,30]. Therefore, as fat mass increases tied to breast cancer risk.
with increasing body weight, aromatase expression and consequently
estrogen levels, are also elevated, an effect that is more prominent in 3. Estrogens, estrogen receptors and breast cancer
postmenopausal women [31–37].
A number of studies have further explored the positive association Estrogens have been shown to function predominantly by inter-
between obesity and local estrogen production by highlighting specific acting with two estrogen receptors (ERs), ERα and ERβ [56]. Estrogen
factors dysregulated in obese adipose tissue that induce aromatase ex- receptors are fundamental for mammary gland maturation and phy-
pression in adipose stromal cells (ASCs) (Fig. 1). Following up on the siological events such as puberty and pregnancy. ERα is found in nearly
seminal findings that proinflammatory mediators and cytokines (e.g. 50–80% of breast cancers, and its expression correlates with better
PGE2, TNFα, IL-1, IL-6, COX-2) play key roles in regulating estrogen prognosis and a lower chance of recurrence [57,58]. ERβ has also been
production in ASCs [18,19,38–42], several recent mechanistic studies detected in breast tumors, and is suspected to contribute to hormonal
have provided greater insight into the regulation of aromatase by some sensitivity and resistance [59,60]. Studies show decreased ERβ RNA
of these factors in obese adipose tissue. For example, using both cell levels in invasive breast cancers in comparison with the normal mam-
culture and clinical samples, Wang et al. showed that p53 is a negative mary gland [60]. Although the role and mechanism through which
regulator of aromatase expression in ASCs and treatment with PGE2, decreased ERβ expression results in tumorigenesis is unknown, results

162
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

Fig. 1. Molecular regulation of aromatase in obesity.

from several studies suggest a stimulatory role of ERα and an inhibitory mutagenic and mitogenic effects of estrogens, the obesity-induced ele-
role of ERβ in relation to proliferation of estrogen-dependent cells vation in local estrogen production is likely to drive DNA damage in
[61,62]. breast epithelial cells leading to a greater risk of tumorigenesis.
Interestingly, there is a growing body of literature implicating es-
3.1. Estrogens as mutagens and effects on breast epithelium trogens in the disruption of the DNA damage response (DDR) and DNA
repair machinery. For example, some studies have indicated that ERα
Estrogens are a significant driver of ER+ breast cancer, with studies downregulates ATM and ATR, important initiators of the DDR [75–77].
suggesting a role in ER- breast cancer as well [63–65]. Neighboring the This is hypothesized to result in defective processing of DNA damage.
estrogen-producing ASCs are breast epithelial cells, which are hormone- The mechanisms by which estrogen signaling alters DDR and DNA re-
sensitive and express the ER. Estrogens play an important in role in the pair has been reviewed in detail [78] and provides a novel theory for
normal development of breast epithelium by stimulating proliferation how estrogens promote breast cancer, i.e. not only by inducing DNA
and ductal morphogenesis [66]. However, when exposed to high levels damage but also potentially diminishing the cell’s ability to sense and
of estrogens such as in the setting of obesity, the pro-proliferative effect repair damage.
of these steroids may cause accumulation of replication errors leading
to mutations and the development of breast cancer (Fig. 2). Pro- 3.2. Mechanism of breast cancer growth in response to estrogen
liferating cells also have higher energy demands that require increased
mitochondrial activity, which could potentially lead to an elevation in In breast cancer, estrogens can act via genomic and non-genomic
reactive oxygen species (ROS) as a byproduct of cellular respiration. mechanisms. Genomic actions of ERs are associated with the regulation
Felty et al. found that estradiol can directly stimulate the production of of estrogen-response element (ERE)-dependent and ERE-independent
intracellular ROS from mitochondria in several breast cancer cell lines gene expression [79–81]. In ERE-dependent genomic activation, es-
[67]. trogen binding to its receptor is associated with increased interaction
Additionally, estrogens can be metabolized to catechols followed by with coactivator proteins in order to bind to the ERE in DNA, resulting
further oxidation to semi-quinones and quinones through a process of in changes in gene expression that regulate growth, differentiation,
redox cycling that produces ROS. This is important in the context of apoptosis and angiogenesis [80]. However, estrogens can also facilitate
tumorigenesis because estrogen quinones are mutagenic and can in- gene transcription via pathways that do not require EREs. In ERE-in-
teract directly with DNA to form adducts, a form of DNA damage dependent genomic activation, the estrogen-ER complex can also in-
[68–70]. Several studies have shown that treating normal breast epi- teract with other DNA bound transcription factors such as Fos/Jun in
thelial cells (MCF-10A) with estrogen metabolites induces elevation in order to bind to AP-1 or SP1 sites in the promoter regions of target
intracellular ROS leading to oxidative DNA damage [71–73]. By in- genes, thereby resulting in activation of gene transcription [79,81].
teracting directly with DNA, estrogen metabolites do not require the Non-genomic effects are actions mediated via activation of ER lo-
estrogen receptor to exert their mutagenic effects, which may explain calized closely to or at the plasma membrane [82–84]. Membrane-as-
the role of estrogen in promoting some ER- breast cancers. Indeed, sociated ER may interact with many proteins including adaptor pro-
Savage et al. found that estrogen and estrogen metabolites caused DNA teins, G-proteins, Src, growth factor receptors (EGFR, IGFR1, HER2),
double strand breaks (DSB) in both normal breast epithelial cells and cytoplasmic kinases (MAPKs, PI3K, AKT) as well as signaling enzymes
ER- breast cancer cells [74]. Given the abundance of evidence for (adenyl cyclase) [85–89]. The actions mediated by this mechanism are

163
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

Fig. 2. Estrogens and tumorigenesis.

independent of gene transcriptional changes [85–89]. invasion in vitro [110]. As mentioned above, estrogens not only influ-
Recent findings show that non-genomic effects also involve the or- ence the etiology of ER+ breast tumors, but also of ER- breast cancers.
phan GPCR-like protein, GPR30 (G protein coupled receptor 30) In ER- breast cancer cells, estrogen actions on GPER1 lead to increased
[90,91]; also named GPER [92]. It has been reported that estrogens can invasion and migration, promoting a prometastatic phenotype [111].
act on membrane GPER in order to stimulate release EGF or EGF-related ERα splice variants have also been shown to mediate extra-nuclear
ligands leading to a transient activation of the EGFR, which in turn effects of estrogens through activation of PKC signaling pathway and
activates MAPK and PI3K signaling pathways [93,94]. Interestingly, promote metastasis in ERα-positive and ERα-negative breast cancer cell
activation of GPER leads to the stimulation of adenylyl cyclase activity lines [112]. Estradiol-induced proteins like G1P3 are able to rescue cells
and increases in cAMP formation, which then leads to a marked in- undergoing anoikis [113], favoring prometastatic estrogenic effects.
hibition of cell proliferation [95]. It is hypothesized that GPR30 inhibits Estradiol-induced expression of Proteinase Inhibitor-9 (PI-9) with in-
MAPK activity and can also increase intracellular calcium stores, re- hibitory activity against Granzyme-B contributes to breast cancer im-
sulting in apoptosis and inhibition of MCF7 cell proliferation [96,97]. mune escape [114]. Another estrogen contribution to metastasis is re-
Thus, GPR30 could be a novel potential target for ER+ breast cancers. lated to premetastatic niche formation and bone marrow-derived
There is also an important cross-talk between ER genomic and non- myeloid (BMDM) cell recruitment. ERα expression in BMDM cells is
genomic signaling pathways. Estrogen binding to nuclear ER can in- necessary for estrogen-mediated tumor promotion of ER- breast cancer
crease transforming growth factor (TGFα) and amphiregulin expres- [115]. Estrogens might contribute to BMDM cell recruitment through
sion. Once TGFα and amphiregulin are bound, they stimulate EGFR in VEGF-A and/or SDF1α that were previously described as downstream
order to activate MAPK and AKT [79,81,98]. Conversely, several cy- mediators of estradiol/ER-induced angiogenesis and macrophage che-
tokines, growth factors, EGFR ligands and IGF1R-related pathways ac- motaxis, respectively. Sartorius et al. demonstrated that estradiol pro-
tivate MAPK/ERK, PI3K/AKT, p90rsk and p38 MAPK, which lead to ER motes brain metastasis of ER- breast cancer cells by modulating astro-
phosphorylation. The phosphorylation of the AF-1 serines 118,167 and cyte function, suggesting that existing endocrine therapies may provide
threonine 311, or other domains, results in the activation of ER some clinical benefit towards reducing and managing brain metastases
[99–103]. in patients with ER- breast tumors [116].
Both genomic and non-genomic signaling pathways of the ER play a A number of epidemiological studies supports these laboratory-
critical role in breast cancer development, progression, and survival. based studies. For example, endocrine therapy has been shown to de-
This is due in part to the regulation of the anti-apoptotic gene Bcl2, pro- crease the risk of developing ER+ contralateral breast cancer [117].
apoptotic gene caspase and cell cycle regulator cyclin D1 [104–106]. Interestingly, in this study, the authors found that patients on anti-
Additionally, estrogens increase the growth of breast tumors by in- estrogen therapy presented a higher risk of contralateral ER- breast
creasing the number of G0/G1 cells entering into the cell cycle, there- cancer. In a retrospective analysis of metastatic behavior of breast
fore resulting in greater proliferation [107,108]. Moreover, PI3K in- cancer subtypes [118], data demonstrate that women with ER+ tumors
teracts with Src in order to promote S-phase entry of MCF7 cells in the are more likely to develop bone metastases and have improved disease-
presence of estradiol [109]. free survival compared to women who have ER- tumors. Their study
support a close relationship between ER+ tumors and metastasis-spe-
3.3. Role of estrogens in breast cancer metastasis cific survival, a finding that is consistent with data from previous stu-
dies [119].
Estrogens have also been shown to influence breast cancer pro-
gression. For example, estrogen treatment has been shown to cause
cytoskeletal remodeling, and contribute to cancer cell migration and

164
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

4. Endocrine therapy this therapy once a month via intramuscular injection [139]. However,
side effects such as diarrhea, hot flushes and throat inflammation have
ERα and progesterone receptor (PR) expression have the greatest been reported that can affect the quality of life of breast cancer patients
predictive value for response to hormonal therapy [120]. Despite breast [140]. It is recommended for use in postmenopausal women who
cancer being the most commonly diagnosed cancer among women, cannot be effectively treated with tamoxifen or aromatase inhibitors, as
cancer-related mortality rates are declining, in part, due to advances in fulvestrant acts independently of estrogens [141–144].
adjuvant therapy [121]. Endocrine or hormonal therapies, including
aromatase inhibitors, have revolutionized treatment for breast cancer 4.3. Aromatase inhibitors
patients. Current endocrine therapy holds major therapeutic value,
especially for ER+ breast cancer patients. Aromatase inhibitors (AIs) are a class of drugs that block the action
of the aromatase enzyme in order to reduce the amount of estrogen in
4.1. Tamoxifen the body [145]. There are two types of AIs, irreversible steroidal in-
hibitors such as exemestane (brand name, Aromasin®) [146], and non-
The initial findings by George Beatson regarding the important role steroidal inhibitors such as anastrozole (brand name, Arimidex®) [147]
estrogen plays in breast development have served as the basis for re- and letrozole (brand name, Femara®) [148]. These three inhibitors
search, development, and discovery of tamoxifen in 1967 by Harper were all approved by the FDA and most clinical studies prove that
and Walpole [122,123]. Tamoxifen is a selective ER modulator (SERM) treating breast cancer with exemestane [149], anastrozole [150] or
used to treat estrogen-dependent breast cancer and reduce the risk of letrozole [151] highly reduces breast cancer recurrence compared with
cancer recurrence in premenopausal and postmenopausal women tamoxifen treatment. The levels of FSH increase as a consequence of
[124]. This drug was approved by the food and drug administration estrogen suppression. For this reason, AIs are contra-indicated in pre-
(FDA) in 1998 and is known by the brand names Nolvadex® and Sol- menopausal women as use can lead to the development of polycystic
tamox®. Tamoxifen is a non-steroidal antiestrogen with triphenylethy- ovaries and incomplete inhibition of ovarian estrogen biosynthesis. AI
lene structure [125]. As a prodrug, tamoxifen has little affinity for the use is also associated with well-documented side effects such as os-
ER. It is metabolized in the liver into an active metabolite, 4-hydro- teoporosis [152], joint and muscle pain [153,154] and hot flushes
xytamoxifen [126] which acts as an antagonist of the ER in breast [155] because of the whole-body inhibition of aromatase and estrogen
tissue, leading to the inhibition of binding with coactivator proteins, production.
thereby blocking the G1 phase of the cell cycle and preventing cell
proliferation. Another tamoxifen metabolite, 4-hydroxy-N-desmethyl 5. Endocrine resistance
tamoxifen (endoxifen), is present at greater concentrations in plasma
than 4-hydroxytamoxifen, and thus may be just as important, if not While endocrine therapy has enhanced the lives of many breast
more, to the anti-estrogenic action of tamoxifen [127,128]. cancer patients, emergence of resistance is inevitable with advanced
In ER+ breast cancer, most clinical studies demonstrate that ta- breast cancer and is to be expected over time. Endocrine resistance in
moxifen should be taken continuously for five years or ten years. cancer cells can be divided broadly into two main categories – de novo
According to the worldwide Adjuvant Tamoxifen: Longer Against and acquired resistance. Breast tumors that show no response to first
Shorter (ATLAS) trial, it was suggested that patients who are treated line hormonal therapy are examples of de novo resistance. On the other
with tamoxifen for ten years have reduced risk of breast cancer recur- hand, tumors that exhibit a response initially to endocrine therapy, but
rence and cancer-associated death compared with patients who are then later recur are examples of acquired or secondary resistance.
treated with tamoxifen for only 5 years [129]. Another clinical trial, the Endocrine resistance may be considered to reflect four possible
aTTom trial, showed that patients who were continuously treated with mechanisms: 1) Pharmacological resistance; 2) Changes in expression
tamoxifen for ten years also had significantly reduced risk of breast of ERα and its co-regulators; 3) Alterations in expression of cell cycle
cancer recurrence compared with patients who took tamoxifen for 5 signaling molecules; 4) Alternate growth receptor pathways, which is
years [130]. Although tamoxifen treatment helps to reduce the risk of especially relevant in the context of obesity.
cancer recurrence, side effects, including endometrial and uterine The first possible theory for developing hormonal resistance is by
cancers [131,132], loss of blood flow to parts of the brain and sig- means of pharmacological mechanisms. As previously mentioned, ta-
nificant loss of bone mineral density in premenopausal women moxifen is a pro-drug that undergoes extensive oxidation in the liver,
[125,133], have been reported, thereby highlighting a need for safer primarily by CYP3A and CYP2D6 [126]. However, polymorphisms in
alternative therapies. tamoxifen-metabolizing genes affect the efficacy of the enzymes and
thus plasma concentrations of the active metabolites of tamoxifen.
4.2. Fulvestrant Retrospective clinical data suggests that women with CYP2D6 4*/4*
genotype have null or reduced enzyme activity, resulting in higher risk
Fulvestrant (ICI 164384) is a selective ER degrader (SERD). It is of tumor relapse [156].
classified as a pure steroidal antiestrogen (European Medicines Agency, The second possible mechanism for developing endocrine resistance
brand name Faslodex®) [134]. It is a class of drug that targets ERs is through mutations in the ESR1 gene, which encodes ERα. Expression
approximately 100 times faster than tamoxifen [134] and immediately of ERα has long been used to predict clinical response to anti-estrogen
causes degradation of the ER in breast cancer cells which reduces their therapy. The mutations mainly occur at two residues in the ligand-
ability to be activated by estrogen as well as the ability of ER to be binding domain, which replaces tyrosine with serine or asparagine at
activated via estrogen-independent mechanisms [135]. The loss of the residue 537 and replaces aspartic acid with glycine at residue 538
ER in breast cancer results in inhibition of breast cancer cell growth. in [157]. A study found a mutation rate of 12% among 76 patients with
vitro, fulvestrant inhibits the growth of tamoxifen-resistant ER+ MCF7 metastatic ER + breast cancer, and a 20% mutation rate among in-
breast cancer cells [136]. in vivo, fulvestrant also inhibits tamoxifen- dividuals with heavily pretreated disease [158]. Analysis of ESR1 mu-
resistant and ER+ MCF7 breast cancer xenografts [137]. In clinical tations was performed on plasma samples from patients participating in
studies, a correlative downregulation of ER with increasing dose was the PALOMA3 and SoFEA studies, and a mutation rate of 25% was
observed in postmenopausal women with primary breast cancer treated observed in patients with breast cancer progression on endocrine
with single doses of fulvestrant for 15–22 days before surgery [138]. therapy, and an even higher mutation rate of 29% in patients who re-
The FDA has approved this drug for the treatment of hormone re- ceived prior AI therapy [159]. Similarly, a mutation rate of 39% was
ceptor-positive metastatic breast cancer. Breast cancer patients receive detected in patients with prior AI sensitivity. These findings suggest

165
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

that although ESR1 mutations are a rare cause of primary endocrine 5.1. Endocrine resistance in obesity
resistance, the mutations arise more commonly with acquired sec-
ondary resistance to AI therapy [159]. Several studies have sought to Several studies have shown that obese women are more likely to
clarify the role of ERβ, if any, in relation to response and resistance to face a poor breast cancer prognosis as compared to lean women, and
endocrine therapy. Despite some conflicting evidence, it has been have a greater chance of breast cancer recurrence [181–183]. In the
suggested that low levels of ERβ expression are related to tamoxifen Anastrozole, Tamoxifen Alone or in Combination (ATAC)
resistance [59]. [150,184,185] and Australian Breast Cancer Study Group (ABCSG)
Additionally, overexpression of the ER co-activator AIB1 (or SRC3 [186] trials there was evidence that the relative benefit of anastrozole
or NCoA3) and down regulation of the co-repressor NCoR is associated vs. tamoxifen was greatest when BMI was normal vs. elevated. A 2014
with tamoxifen resistance [160–162]. Finally, increased levels of tran- meta-analysis concluded that there is an association between BMI and
scription factors such as NFkB and AP-1, which increase the interaction outcome, and that there is increased risk of mortality in individuals
of ER with specific gene promoters, have also been linked to endocrine with BMI over 25 kg/m2 at diagnosis [187]. Different studies have
resistance [163,164]. evaluated the efficacy of AI inhibitors in suppressing circulating es-
Endocrine resistance can also occur as a result of alterations in key trogen levels and its relation with BMI. Interestingly, Elliott et al. found
cell cycle checkpoints [165]. The cell cycle involves a complex se- that BMI and estradiol were higher in metastatic patients [188], and
quence of events through which a cell duplicates, and involves many Lonning et al. showed that estrone levels positively correlate with BMI
regulatory proteins such as cyclin proteins, cyclin-dependent kinases, in women receiving AIs [189]. In a study by Sestak et al., obese women
oncogenes and tumor-suppressor genes, and mitotic checkpoint pro- were less likely to observe benefits of AI treatment despite a significant
teins. The balance of proliferative and anti-proliferative signals de- reduction in circulating estrogens [190]. Therefore, it seems that higher
termines if a cell will progress from the G1 phase to the S phase, or BMI may contribute to reduced efficacy of AIs.
withdraw into the dormant phase [166]. Anti-proliferative signals are Many circulating factors found in the serum of obese post-
communicated via the retinoblastoma (Rb) tumor suppressor protein, menopausal women can amplify crosstalk between nongenomic ERα
while Rb itself is regulated via complexes of cyclin and cyclin-depen- signaling and PI3K/AKT/mTOR and RAF/MEK/ERK pathways which
dent kinases [167]. Cyclin-dependent kinase 4 (CDK4), in complex with promotes breast cancer progression [191]. For example, cells grown in
cyclin D1, D2 or D3, controls the phosphorylation of Rb, which in turn media supplemented with sera from obese patients have greater IGF1R
regulates the progression of the cell from G1 to S phase [168]. By means activation in comparison to control sera [191]. As mentioned pre-
of CDK4 inactivation, or cyclin D1 and E1 amplification, tumor cells are viously, FGFR1 is a known mediator of endocrine therapy resistance
able to circumvent cell cycle regulation [169,170]. Interestingly, cyclin [177,179,192]. The FGFR1 regulatory pathway is one of the pathways
D1 amplification is a common occurrence in ER+ breast cancers with that shows shared activation in patients with obesity and those with
58% and 29% incidence rate in luminal B and luminal A cancers, re- resistance to AI therapy [193]. Phosphorylation of FGFR1 and FGFR1
spectively (Cancer Genome Atlas Network, 2012). Similarly, reduced ligand expression is increased with obesity, metabolic dysfunction,
expression of p21 and p27 (negative cell cycle regulators) and in- weight gain and adipocyte hypertrophy [193]. Weight gain leads to a
activation of Rb are also associated with poor response to hormonal positive energy balance, and in the context of obesity/metabolic dys-
therapy, especially tamoxifen [171,172]. As a result, endocrine therapy function, promotes FGFR ligand production from adipose tissue, which
and CDK4/6 inhibitors, in combination or sequentially, are now a vi- may also result in the activation of receptors in nearby breast cancer
able option for the treatment of hormone receptor-positive breast cells to promote growth after estrogen deprivation [193]. Leptin has
cancer as a first line therapy and following development of resistance to also been shown to induce cell proliferation through activation of
endocrine therapy [173]. MAPK signaling, and its receptors are expressed in both normal breast
Another possible mechanism for developing endocrine resistance is tissue and solid tumors [194]. Additionally, leptin mimics the effects of
by the means of enhanced autophagy, which is an intracellular process ERα transactivation in an ER+ breast cancer cell line, including down-
that recycles damaged or unnecessary organelles (macroautophagy) or regulation of ER mRNA and protein levels and up-regulation of the
proteins (microautophagy). Under conditions of compromised autop- estrogen-dependent gene pS2 [195]. Since serum leptin concentrations
hagy, it was noted that cytotoxic effects of 4-hydroxytamoxifen were are correlated with the percentage of body fat [196], obese individuals
significantly increased [174]. Inhibition of autophagosome function have a higher chance of developing endocrine resistance due to the
stimulated a strong caspase-dependent cell death in the 4-hydro- mechanism described above. Moreover, as described earlier, leptin
xytamoxifen treated, anti-estrogen resistant cells [174,175]. Thus, im- stimulates the expression of aromatase through enhanced binding of
paired autophagy increases sensitivity to endocrine therapy, and in- CREB, CRTC and AP-1 to specific sites in the promoter region [48,197],
hibition of the autophagosome may be a potential target to overcome thus amplifying in situ production of E2 and driving breast tumor
resistance and improve efficacy of hormonal treatment of ER+ breast growth.
cancers. Nevertheless, a number of studies found no association between
Finally, overexpression and amplification of growth factor re- BMI and breast cancer outcomes. Coscia et al. showed a reduction in
ceptors, such as FGFR1 (fibroblast growth factor receptor -1), IGF1R plasma estrone and estradiol following anastrozol treatment, but no
(insulin growth factor -1 receptor), HER2 (human epidermal growth significant changes in steroid concentration in association with BMI
receptor -2), HER3 (human epidermal growth receptor -3), and EGFR [198]. Two recent studies showed no significant relationship between
(epidermal growth factor receptor) [176–179], which converge on the high BMI and the efficacy of two different aromatase inhibitors
PI3K/AKT/mTOR and RAF/MEK/ERK pathways, have been shown to [199,200]. In a cohort where more than two thirds of postmenopausal
be associated with sustained tumor proliferation and survival in- women receiving adjuvant letrozole therapy were obese, there was no
dependent of estrogen [176–179]. These pathways provide alternative association for worse outcome in the obese women compared with lean
survival stimuli to the tumors, and can emerge to act as ER-independent women receiving the same treatment [200]. Zewenghiel et al. found no
drivers of tumor growth, thus conferring resistance to endocrine statistically significant difference between the three BMI categories
therapy. In addition, these pathways can be activated by amplification (normal weight, overweight and obesity) and time to progression
of the receptors and/or their respective ligands. Alternatively, dereg- during fulvestrant treatment in their 173 patient cohort [199]. There-
ulation of downstream signaling molecules such as an activating mu- fore, additional studies are required to determine which patient group
tation in the PI3K p110 catalytic subunit or the loss of expression of and/or treatment may be impacted by BMI before altering disease
PTEN tumor suppressor can also lead in activation of the pathways management.
[180].

166
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

6. Conclusions and future directions [24] A.G. Renehan, et al., Body-mass index and incidence of cancer: a systematic review
and meta-analysis of prospective observational studies, Lancet 371 (9612) (2008)
569–578.
Obesity is now an established risk factor for breast cancer in post- [25] H.S. Bae, et al., Obesity and epithelial ovarian cancer survival: a systematic review
menopausal women. The local production of estrogens in the breast and meta-analysis, J. Ovarian Res. 7 (2014) 41.
adipose tissue is suspected to be a key driver of breast cancer devel- [26] K.A. Brown, E.R. Simpson, Obesity and breast cancer: mechanisms and therapeutic
implications, Front. Biosci. (Elite Ed) 4 (2012) 2515–2524.
opment and growth, and a mediator of resistance to endocrine therapy. [27] J.M. Grodin, P.K. Siiteri, P.C. MacDonald, Source of estrogen production in post-
Additional factors in obesity are also important for disease development menopausal women, J. Clin. Endocrinol. Metab. 36 (2) (1973) 207–214.
and progression, and a better understanding of the mechanisms at play [28] R.J. Santen, et al., History of aromatase: saga of an important biological mediator
and therapeutic target, Endocr. Rev. 30 (4) (2009) 343–375.
will inform management of breast cancer in an increasingly obese fe- [29] Y. Kakugawa, et al., Associations of obesity and physical activity with serum and
male population. Additional studies are also required to determine intratumoral sex steroid hormone levels among postmenopausal women with
whether pharmacological or lifestyle interventions will reduce the risk breast cancer: analysis of paired serum and tumor tissue samples, Breast Cancer
Res. Treat. 162 (1) (2017) 115–125.
of breast cancer development and progression in this obese population.
[30] C. Schairer, et al., Quantifying the role of circulating unconjugated estradiol in
mediating the body mass index-breast cancer association, Cancer Epidemiol.
Acknowledgments Biomarkers Prev. 25 (1) (2016) 105–113.
[31] L. Baglietto, et al., Circulating steroid hormone concentrations in postmenopausal
women in relation to body size and composition, Breast Cancer Res. Treat. 115 (1)
This work was supported by NIH/NCIR01CA215797-01 (to K.A. (2009) 171–179.
Brown) and by the Dr. Robert C. and Veronica Atkins Foundation [32] D.P. Rose, D. Komninou, G.D. Stephenson, Obesity, adipocytokines, and insulin
Curriculum in Metabolic Diseases at Weill Cornell Medical College (to resistance in breast cancer, Obes. Rev. 5 (3) (2004) 153–165.
[33] P.G. Morris, et al., Inflammation and increased aromatase expression occur in the
R. Moges). breast tissue of obese women with breast cancer, Cancer Prev. Res. (Phila) 4 (7)
(2011) 1021–1029.
References [34] T.J. Key, et al., Body mass index, serum sex hormones, and breast cancer risk in
postmenopausal women, J. Natl. Cancer Inst. 95 (16) (2003) 1218–1226.
[35] K.A. Brown, et al., Menopause is a determinant of breast aromatase expression and
[1] K.S. Saladin, Anatomy & Physiology : the Unity of Form and Function, 6. ed., its associations with BMI, inflammation, and systemic markers, J. Clin. Endocrinol.
McGraw-Hill, New York, NY, 2012 xxvi, 1136, A-14, G-20, C-3, I-49. Metab. 102 (5) (2017) 1692–1701.
[2] J. Cui, Y. Shen, R. Li, Estrogen synthesis and signaling pathways during aging: [36] A. Lukanova, et al., Body mass index, circulating levels of sex-steroid hormones,
from periphery to brain, Trends Mol. Med. 19 (3) (2013) 197–209. IGF-I and IGF-binding protein-3: a cross-sectional study in healthy women, Eur. J.
[3] J.L. Kelsey, M.D. Gammon, E.M. John, Reproductive factors and breast cancer, Endocrinol. 150 (2) (2004) 161–171.
Epidemiol. Rev. 15 (1) (1993) 36–47. [37] A. McTiernan, et al., Relation of BMI and physical activity to sex hormones in
[4] M.C. Pike, et al., Estrogens, progestogens, normal breast cell proliferation, and postmenopausal women, Obesity (Silver Spring) 14 (9) (2006) 1662–1677.
breast cancer risk, Epidemiol. Rev. 15 (1) (1993) 17–35. [38] A. Singh, et al., The regulation of aromatase activity in breast fibroblasts: the role
[5] H.K. Ghayee, R.J. Auchus, Basic concepts and recent developments in human of interleukin-6 and prostaglandin E2, Endocr. Relat. Cancer 6 (2) (1999)
steroid hormone biosynthesis, Rev. Endocr. Metab. Disord. 8 (4) (2007) 289–300. 139–147.
[6] S.J. Santner, P.D. Feil, R.J. Santen, In situ estrogen production via the estrone [39] J.A. Richards, R.W. Brueggemeier, Prostaglandin E2 regulates aromatase activity
sulfatase pathway in breast tumors: relative importance versus the aromatase and expression in human adipose stromal cells via two distinct receptor subtypes,
pathway, J. Clin. Endocrinol. Metab. 59 (1) (1984) 29–33. J. Clin. Endocrinol. Metab. 88 (6) (2003) 2810–2816.
[7] C. Gunnarsson, E. Hellqvist, O. Stal, 17beta-Hydroxysteroid dehydrogenases in- [40] A. Purohit, S.P. Newman, M.J. Reed, The role of cytokines in regulating estrogen
volved in local oestrogen synthesis have prognostic significance in breast cancer, synthesis: implications for the etiology of breast cancer, Breast Cancer Res. 4 (2)
Br. J. Cancer 92 (3) (2005) 547–552. (2002) 65–69.
[8] H.V. Thomas, et al., A prospective study of endogenous serum hormone con- [41] M.J. Reed, et al., Interleukin-1 and interleukin-6 in breast cyst fluid: their role in
centrations and breast cancer risk in post-menopausal women on the island of regulating aromatase activity in breast cancer cells, J. Endocrinol. 132 (3) (1992)
Guernsey, Br. J. Cancer 76 (3) (1997) 401–405. R5–8.
[9] P.G. Toniolo, et al., A prospective study of endogenous estrogens and breast cancer [42] K. Subbaramaiah, et al., Increased levels of COX-2 and prostaglandin E2 contribute
in postmenopausal women, J. Natl. Cancer Inst. 87 (3) (1995) 190–197. to elevated aromatase expression in inflamed breast tissue of obese women,
[10] A.A. Larionov, L.M. Berstein, W.R. Miller, Local uptake and synthesis of oestrone Cancer Discov. 2 (4) (2012) 356–365.
in normal and malignant postmenopausal breast tissues, J. Steroid Biochem. Mol. [43] X. Wang, et al., Prostaglandin E2 inhibits p53 in human breast adipose stromal
Biol. 81 (1) (2002) 57–64. cells: a novel mechanism for the regulation of aromatase in obesity and breast
[11] F. Labrie, Intracrinology in action: importance of extragonadal sex steroid bio- cancer, Cancer Res. 75 (4) (2015) 645–655.
synthesis and inactivation in peripheral tissues in both women and men, J. Steroid [44] N.U. Samarajeewa, et al., HIF-1alpha stimulates aromatase expression driven by
Biochem. Mol. Biol. 145 (2015) 131–132. prostaglandin E2 in breast adipose stroma, Breast Cancer Res. 15 (2) (2013) R30.
[12] F. Labrie, All sex steroids are made intracellularly in peripheral tissues by the [45] K. Subbaramaiah, et al., Prostaglandin E2 down-regulates sirtuin 1 (SIRT1) leading
mechanisms of intracrinology after menopause, J. Steroid Biochem. Mol. Biol. 145 to elevated levels of aromatase, providing insights into the obesity-breast cancer
(2015) 133–138. connection, J. Biol. Chem. (2018).
[13] T. Suzuki, et al., Sex steroid-producing enzymes in human breast cancer, Endocr. [46] L.W. Bowers, et al., Obesity-associated systemic interleukin-6 promotes pre-adi-
Relat. Cancer 12 (4) (2005) 701–720. pocyte aromatase expression via increased breast cancer cell prostaglandin E2
[14] C. Belanger, et al., Adipose tissue intracrinology: potential importance of local production, Breast Cancer Res. Treat. 149 (1) (2015) 49–57.
androgen/estrogen metabolism in the regulation of adiposity, Horm. Metab. Res. [47] H. Zahid, et al., Leptin regulation of the p53-HIF1alpha/PKM2-aromatase axis in
34 (11-12) (2002) 737–745. breast adipose stromal cells: a novel mechanism for the obesity-breast cancer link,
[15] S. Sebastian, S.E. Bulun, A highly complex organization of the regulatory region of Int. J. Obes. (Lond) 42 (4) (2018) 711–720.
the human CYP19 (aromatase) gene revealed by the Human Genome Project, J. [48] K.A. Brown, et al., Subcellular localization of cyclic AMP-responsive element
Clin. Endocrinol. Metab. 86 (10) (2001) 4600–4602. binding protein-regulated transcription coactivator 2 provides a link between
[16] E.R. Simpson, et al., Aromatase cytochrome P450, the enzyme responsible for obesity and breast cancer in postmenopausal women, Cancer Res. 69 (13) (2009)
estrogen biosynthesis, Endocr. Rev. 15 (3) (1994) 342–355. 5392–5399.
[17] M.S. Mahendroo, C.R. Mendelson, E.R. Simpson, Tissue-specific and hormonally [49] N.U. Samarajeewa, et al., CREB-regulated transcription co-activator family sti-
controlled alternative promoters regulate aromatase cytochrome P450 gene ex- mulates promoter II-driven aromatase expression in preadipocytes, Horm. Cancer
pression in human adipose tissue, J. Biol. Chem. 268 (26) (1993) 19463–19470. 4 (4) (2013) 233–241.
[18] Y. Zhao, et al., Aromatase P450 gene expression in human adipose tissue. Role of a [50] M.M. Docanto, et al., Ghrelin and des-acyl ghrelin inhibit aromatase expression
Jak/STAT pathway in regulation of the adipose-specific promoter, J. Biol. Chem. and activity in human adipose stromal cells: suppression of cAMP as a possible
270 (27) (1995) 16449–16457. mechanism, Breast Cancer Res. Treat. 147 (1) (2014) 193–201.
[19] Y. Zhao, et al., Tumor necrosis factor-alpha stimulates aromatase gene expression [51] V.R. Agarwal, et al., Alternatively spliced transcripts of the aromatase cytochrome
in human adipose stromal cells through use of an activating protein-1 binding site P450 (CYP19) gene in adipose tissue of women, J. Clin. Endocrinol. Metab. 82 (1)
upstream of promoter 1.4, Mol. Endocrinol. 10 (11) (1996) 1350–1357. (1997) 70–74.
[20] Z. Huang, et al., Dual effects of weight and weight gain on breast cancer risk, [52] Y. Zhao, et al., Transcriptional regulation of CYP19 gene (aromatase) expression in
JAMA 278 (17) (1997) 1407–1411. adipose stromal cells in primary culture, J. Steroid Biochem. Mol. Biol. 61 (3-6)
[21] P.A. van den Brandt, et al., Pooled analysis of prospective cohort studies on height, (1997) 203–210.
weight, and breast cancer risk, Am. J. Epidemiol. 152 (6) (2000) 514–527. [53] G. Martinez-Chacon, et al., IL-10 suppresses TNF-alpha-induced expression of
[22] G.K. Reeves, et al., Cancer incidence and mortality in relation to body mass index human aromatase gene in mammary adipose tissue, FASEB J. 32 (6) (2018)
in the Million Women Study: cohort study, BMJ 335 (7630) (2007) 1134. 3361–3370.
[23] E.E. Calle, R. Kaaks, Overweight, obesity and cancer: epidemiological evidence [54] N.M. Iyengar, et al., Metabolic obesity, adipose inflammation and elevated breast
and proposed mechanisms, Nat. Rev. Cancer 4 (8) (2004) 579–591. aromatase in women with normal body mass index, Cancer Prev. Res. (Phila) 10

167
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

(4) (2017) 235–243. protein-coupled receptor homolog, GPR30, and occurs via trans-activation of the
[55] P.H. Lahmann, et al., A prospective study of adiposity and postmenopausal breast epidermal growth factor receptor through release of HB-EGF, Mol. Endocrinol. 14
cancer risk: the Malmo Diet and cancer study, Int. J. Cancer 103 (2) (2003) (10) (2000) 1649–1660.
246–252. [91] P. Thomas, et al., Conserved estrogen binding and signaling functions of the G
[56] G.G. Kuiper, et al., Cloning of a novel receptor expressed in rat prostate and ovary, protein-coupled estrogen receptor 1 (GPER) in mammals and fish, Steroids 75 (8-
Proc. Natl. Acad. Sci. U. S. A. 93 (12) (1996) 5925–5930. 9) (2010) 595–602.
[57] C.K. Osborne, Steroid hormone receptors in breast cancer management, Breast [92] S.P. Alexander, A. Mathie, J.A. Peters, Guide to Receptors and Channels (GRAC),
Cancer Res. Treat. 51 (3) (1998) 227–238. 5th edition, Br. J. Pharmacol. 164 (2011) S1–324 Suppl. 1.
[58] W.L. McGuire, Hormone receptors: their role in predicting prognosis and response [93] P. Thomas, et al., Identity of an estrogen membrane receptor coupled to a G
to endocrine therapy, Semin. Oncol. 5 (4) (1978) 428–433. protein in human breast cancer cells, Endocrinology 146 (2) (2005) 624–632.
[59] M. Esslimani-Sahla, et al., Estrogen receptor beta (ER beta) level but not its ER [94] E.J. Filardo, et al., Estrogen action via the G protein-coupled receptor, GPR30:
beta cx variant helps to predict tamoxifen resistance in breast cancer, Clin. Cancer stimulation of adenylyl cyclase and cAMP-mediated attenuation of the epidermal
Res. 10 (17) (2004) 5769–5776. growth factor receptor-to-MAPK signaling axis, Mol. Endocrinol. 16 (1) (2002)
[60] C. Palmieri, et al., Estrogen receptor beta in breast cancer, Endocr. Relat. Cancer 9 70–84.
(1) (2002) 1–13. [95] T.M. Ahola, et al., G protein-coupled receptor 30 is critical for a progestin-induced
[61] S. Dupont, et al., Effect of single and compound knockouts of estrogen receptors growth inhibition in MCF-7 breast cancer cells, Endocrinology 143 (9) (2002)
alpha (ERalpha) and beta (ERbeta) on mouse reproductive phenotypes, 3376–3384.
Development 127 (19) (2000) 4277–4291. [96] T.M. Ahola, et al., Progestin and G protein-coupled receptor 30 inhibit mitogen-
[62] K.S. Korach, Insights from the study of animals lacking functional estrogen re- activated protein kinase activity in MCF-7 breast cancer cells, Endocrinology 143
ceptor, Science 266 (5190) (1994) 1524–1527. (12) (2002) 4620–4626.
[63] C. Pequeux, et al., Stromal estrogen receptor-alpha promotes tumor growth by [97] E.A. Ariazi, et al., The G protein-coupled receptor GPR30 inhibits proliferation of
normalizing an increased angiogenesis, Cancer Res. 72 (12) (2012) 3010–3019. estrogen receptor-positive breast cancer cells, Cancer Res. 70 (3) (2010)
[64] W. Yue, et al., Estrogen receptor-dependent and independent mechanisms of 1184–1194.
breast cancer carcinogenesis, Steroids 78 (2) (2013) 161–170. [98] J. Frasor, et al., Profiling of estrogen up- and down-regulated gene expression in
[65] J.D. Yager, N.E. Davidson, Estrogen carcinogenesis in breast cancer, N. Engl. J. human breast cancer cells: insights into gene networks and pathways underlying
Med. 354 (3) (2006) 270–282. estrogenic control of proliferation and cell phenotype, Endocrinology 144 (10)
[66] M.D. Sternlicht, Key stages in mammary gland development: the cues that regulate (2003) 4562–4574.
ductal branching morphogenesis, Breast Cancer Res. 8 (1) (2006) 201. [99] M. Sun, et al., Phosphatidylinositol-3-OH Kinase (PI3K)/AKT2, activated in breast
[67] Q. Felty, et al., Estrogen-induced mitochondrial reactive oxygen species as signal- cancer, regulates and is induced by estrogen receptor alpha (ERalpha) via inter-
transducing messengers, Biochemistry 44 (18) (2005) 6900–6909. action between ERalpha and PI3K, Cancer Res. 61 (16) (2001) 5985–5991.
[68] S.V. Fernandez, I.H. Russo, J. Russo, Estradiol and its metabolites 4-hydro- [100] S. Kato, et al., Activation of the estrogen receptor through phosphorylation by
xyestradiol and 2-hydroxyestradiol induce mutations in human breast epithelial mitogen-activated protein kinase, Science 270 (5241) (1995) 1491–1494.
cells, Int. J. Cancer 118 (8) (2006) 1862–1868. [101] P.B. Joel, et al., pp90rsk1 regulates estrogen receptor-mediated transcription
[69] F. Lu, et al., Estrogen metabolism and formation of estrogen-DNA adducts in es- through phosphorylation of Ser-167, Mol. Cell. Biol. 18 (4) (1998) 1978–1984.
tradiol-treated MCF-10F cells. The effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin [102] R.A. Campbell, et al., Phosphatidylinositol 3-kinase/AKT-mediated activation of
induction and catechol-O-methyltransferase inhibition, J. Steroid Biochem. Mol. estrogen receptor alpha: a new model for anti-estrogen resistance, J. Biol. Chem.
Biol. 105 (1-5) (2007) 150–158. 276 (13) (2001) 9817–9824.
[70] E.L. Cavalieri, et al., Molecular origin of cancer: catechol estrogen-3,4-quinones as [103] G. Bunone, et al., Activation of the unliganded estrogen receptor by EGF involves
endogenous tumor initiators, Proc. Natl. Acad. Sci. U. S. A. 94 (20) (1997) the MAP kinase pathway and direct phosphorylation, EMBO J. 15 (9) (1996)
10937–10942. 2174–2183.
[71] S.A. Park, et al., 4-hydroxyestradiol induces anchorage-independent growth of [104] F. Acconcia, et al., Survival versus apoptotic 17beta-estradiol effect: role of ER
human mammary epithelial cells via activation of IkappaB kinase: potential role of alpha and ER beta activated non-genomic signaling, J. Cell. Physiol. 203 (1)
reactive oxygen species, Cancer Res. 69 (6) (2009) 2416–2424. (2005) 193–201.
[72] K.C. Fussell, et al., Catechol metabolites of endogenous estrogens induce redox [105] M. Marino, et al., Distinct nongenomic signal transduction pathways controlled by
cycling and generate reactive oxygen species in breast epithelial cells, 17beta-estradiol regulate DNA synthesis and cyclin D(1) gene transcription in
Carcinogenesis 32 (8) (2011) 1285–1293. HepG2 cells, Mol. Biol. Cell 13 (10) (2002) 3720–3729.
[73] B. Singh, H.K. Bhat, Superoxide dismutase 3 is induced by antioxidants, inhibits [106] M. Marino, F. Acconcia, A. Trentalance, Biphasic estradiol-induced AKT phos-
oxidative DNA damage and is associated with inhibition of estrogen-induced phorylation is modulated by PTEN via MAP kinase in HepG2 cells, Mol. Biol. Cell
breast cancer, Carcinogenesis 33 (12) (2012) 2601–2610. 14 (6) (2003) 2583–2591.
[74] K.I. Savage, et al., BRCA1 deficiency exacerbates estrogen-induced DNA damage [107] R.B. Dickson, M.E. Lippman, Growth factors in breast cancer, Endocr. Rev. 16 (5)
and genomic instability, Cancer Res. 74 (10) (2014) 2773–2784. (1995) 559–589.
[75] X. Guo, et al., Estrogen receptor alpha regulates ATM Expression through miRNAs [108] S.F. Doisneau-Sixou, et al., Estrogen and antiestrogen regulation of cell cycle
in breast cancer, Clin. Cancer Res. 19 (18) (2013) 4994–5002. progression in breast cancer cells, Endocr. Relat. Cancer 10 (2) (2003) 179–186.
[76] L. Song, et al., miR-18a impairs DNA damage response through downregulation of [109] G. Castoria, et al., PI3-kinase in concert with Src promotes the S-phase entry of
ataxia telangiectasia mutated (ATM) kinase, PLoS One 6 (9) (2011) e25454. oestradiol-stimulated MCF-7 cells, EMBO J. 20 (21) (2001) 6050–6059.
[77] A. Pedram, et al., Estrogen inhibits ATR signaling to cell cycle checkpoints and [110] M.S. Giretti, et al., Extra-nuclear signalling of estrogen receptor to breast cancer
DNA repair, Mol. Biol. Cell 20 (14) (2009) 3374–3389. cytoskeletal remodelling, migration and invasion, PLoS One 3 (5) (2008) e2238.
[78] C.E. Caldon, Estrogen signaling and the DNA damage response in hormone de- [111] Q.F. Jiang, et al., 17beta-estradiol promotes the invasion and migration of nuclear
pendent breast cancers, Front. Oncol. 4 (2014) 106. estrogen receptor-negative breast cancer cells through cross-talk between GPER1
[79] B.J. Deroo, K.S. Korach, Estrogen receptors and human disease, J. Clin. Invest. 116 and CXCR1, J. Steroid Biochem. Mol. Biol. 138 (2013) 314–324.
(3) (2006) 561–570. [112] R.A. Chaudhri, et al., Membrane estrogen signaling enhances tumorigenesis and
[80] J.M. Hall, J.F. Couse, K.S. Korach, The multifaceted mechanisms of estradiol and metastatic potential of breast cancer cells via estrogen receptor-alpha36
estrogen receptor signaling, J. Biol. Chem. 276 (40) (2001) 36869–36872. (ERalpha36), J. Biol. Chem. 287 (10) (2012) 7169–7181.
[81] N.J. McKenna, R.B. Lanz, B.W. O’Malley, Nuclear receptor coregulators: cellular [113] V. Cheriyath, et al., G1P3, an interferon- and estrogen-induced survival protein
and molecular biology, Endocr. Rev. 20 (3) (1999) 321–344. contributes to hyperplasia, tamoxifen resistance and poor outcomes in breast
[82] M. Razandi, et al., ERs associate with and regulate the production of caveolin: cancer, Oncogene 31 (17) (2012) 2222–2236.
implications for signaling and cellular actions, Mol. Endocrinol. 16 (1) (2002) [114] M. Lauricella, et al., The analysis of estrogen receptor-alpha positive breast cancer
100–115. stem-like cells unveils a high expression of the serpin proteinase inhibitor PI-9:
[83] M. Razandi, et al., Proximal events in signaling by plasma membrane estrogen possible regulatory mechanisms, Int. J. Oncol. 49 (1) (2016) 352–360.
receptors, J. Biol. Chem. 278 (4) (2003) 2701–2712. [115] V. Iyer, et al., Estrogen promotes ER-negative tumor growth and angiogenesis
[84] M. Marino, P. Ascenzi, Membrane association of estrogen receptor alpha and beta through mobilization of bone marrow-derived monocytes, Cancer Res. 72 (11)
influences 17beta-estradiol-mediated cancer cell proliferation, Steroids 73 (9-10) (2012) 2705–2713.
(2008) 853–858. [116] C.A. Sartorius, et al., Estrogen promotes the brain metastatic colonization of triple
[85] C.W. Wong, et al., Estrogen receptor-interacting protein that modulates its non- negative breast cancer cells via an astrocyte-mediated paracrine mechanism,
genomic activity-crosstalk with Src/Erk phosphorylation cascade, Proc. Natl. Oncogene 35 (22) (2016) 2881–2892.
Acad. Sci. U. S. A. 99 (23) (2002) 14783–14788. [117] C.I. Li, et al., Tamoxifen therapy for primary breast cancer and risk of contralateral
[86] R. Schiff, et al., Cross-talk between estrogen receptor and growth factor pathways breast cancer, J. Natl. Cancer Inst. 93 (13) (2001) 1008–1013.
as a molecular target for overcoming endocrine resistance, Clin. Cancer Res. 10 (1 [118] C.D. Savci-Heijink, et al., Retrospective analysis of metastatic behaviour of breast
Pt 2) (2004) 331S–336S. cancer subtypes, Breast Cancer Res. Treat. 150 (3) (2015) 547–557.
[87] A. Migliaccio, et al., Tyrosine kinase/p21ras/MAP-kinase pathway activation by [119] K.R. Hess, et al., Estrogen receptors and distinct patterns of breast cancer relapse,
estradiol-receptor complex in MCF-7 cells, EMBO J. 15 (6) (1996) 1292–1300. Breast Cancer Res. Treat. 78 (1) (2003) 105–118.
[88] A. Migliaccio, et al., Activation of the Src/p21ras/Erk pathway by progesterone [120] V.J. Bardou, et al., Progesterone receptor status significantly improves outcome
receptor via cross-talk with estrogen receptor, EMBO J. 17 (7) (1998) 2008–2018. prediction over estrogen receptor status alone for adjuvant endocrine therapy in
[89] S. Kahlert, et al., Estrogen receptor alpha rapidly activates the IGF-1 receptor two large breast cancer databases, J. Clin. Oncol. 21 (10) (2003) 1973–1979.
pathway, J. Biol. Chem. 275 (24) (2000) 18447–18453. [121] B.A. Kohler, et al., Annual report to the nation on the status of cancer, 1975-2011,
[90] E.J. Filardo, et al., Estrogen-induced activation of Erk-1 and Erk-2 requires the G featuring incidence of breast cancer subtypes by race/ethnicity, poverty, and state,

168
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

J. Natl. Cancer Inst. 107 (6) (2015) p. djv048. [152] H.T. Mouridsen, Incidence and management of side effects associated with ar-
[122] G.T. Beatson, On the treatment of inoperable cases of carcinoma of the mamma: omatase inhibitors in the adjuvant treatment of breast cancer in postmenopausal
suggestions for a new method of treatment, with illustrative cases, Trans Med Chir women, Curr. Med. Res. Opin. 22 (8) (2006) 1609–1621.
Soc Edinb 15 (1896) 153–179. [153] E. Amir, et al., Toxicity of adjuvant endocrine therapy in postmenopausal breast
[123] M.J. Harper, A.L. Walpole, A new derivative of triphenylethylene: effect on im- cancer patients: a systematic review and meta-analysis, J. Natl. Cancer Inst. 103
plantation and mode of action in rats, J. Reprod. Fertil. 13 (1) (1967) 101–119. (17) (2011) 1299–1309.
[124] V.C. Jordan, Fourteenth Gaddum Memorial Lecture. A current view of tamoxifen [154] K.D. Crew, et al., Prevalence of joint symptoms in postmenopausal women taking
for the treatment and prevention of breast cancer, Br. J. Pharmacol. 110 (2) (1993) aromatase inhibitors for early-stage breast cancer, J. Clin. Oncol. 25 (25) (2007)
507–517. 3877–3883.
[125] M. Clemons, S. Danson, A. Howell, Tamoxifen ("Nolvadex"): a review, Cancer [155] National_Cancer_Institute, Hormone Therapy for the Breast, Available from:
Treat. Rev. 28 (4) (2002) 165–180. (2012) http://www.cancer.gov/cancertopics/factsheet/Therapy/hormone-
[126] Z. Desta, et al., Comprehensive evaluation of tamoxifen sequential bio- therapy-breast.
transformation by the human cytochrome P450 system in vitro: prominent roles [156] M.P. Goetz, et al., Pharmacogenetics of tamoxifen biotransformation is associated
for CYP3A and CYP2D6, J. Pharmacol. Exp. Ther. 310 (3) (2004) 1062–1075. with clinical outcomes of efficacy and hot flashes, J. Clin. Oncol. 23 (36) (2005)
[127] M.D. Johnson, et al., Pharmacological characterization of 4-hydroxy-N-desmethyl 9312–9318.
tamoxifen, a novel active metabolite of tamoxifen, Breast Cancer Res. Treat. 85 (2) [157] W. Toy, et al., ESR1 ligand-binding domain mutations in hormone-resistant breast
(2004) 151–159. cancer, Nat. Genet. 45 (12) (2013) 1439–1445.
[128] Y.C. Lim, et al., Endoxifen (4-hydroxy-N-desmethyl-tamoxifen) has anti-estrogenic [158] R. Jeselsohn, et al., Emergence of constitutively active estrogen receptor-alpha
effects in breast cancer cells with potency similar to 4-hydroxy-tamoxifen, Cancer mutations in pretreated advanced estrogen receptor-positive breast cancer, Clin.
Chemother. Pharmacol. 55 (5) (2005) 471–478. Cancer Res. 20 (7) (2014) 1757–1767.
[129] C. Davies, et al., Long-term effects of continuing adjuvant tamoxifen to 10 years [159] C. Fribbens, et al., Plasma ESR1 mutations and the treatment of estrogen receptor-
versus stopping at 5 years after diagnosis of oestrogen receptor-positive breast positive advanced breast Cancer, J. Clin. Oncol. 34 (25) (2016) 2961–2968.
cancer: ATLAS, a randomised trial, Lancet 381 (9869) (2013) 805–816. [160] R.M. Lavinsky, et al., Diverse signaling pathways modulate nuclear receptor re-
[130] R.G. Gray, D. Rea, K. Handley, ATTom: long-term effects of continouing adjuvant cruitment of N-CoR and SMRT complexes, Proc. Natl. Acad. Sci. U. S. A. 95 (6)
tamoxifen to 10 years versus stopping at 5 years in 6,953 women with early breast (1998) 2920–2925.
cancer, J. Clin. Oncol. (2013) (suppl; abstr 5). [161] C.K. Osborne, et al., Role of the estrogen receptor coactivator AIB1 (SRC-3) and
[131] B. Fisher, et al., Tamoxifen for prevention of breast cancer: report of the national HER-2/neu in tamoxifen resistance in breast cancer, J. Natl. Cancer Inst. 95 (5)
surgical adjuvant breast and bowel project P-1 study, J. Natl. Cancer Inst. 90 (18) (2003) 353–361.
(1998) 1371–1388. [162] J. Shou, et al., Mechanisms of tamoxifen resistance: increased estrogen receptor-
[132] Tamoxifen for early breast cancer: an overview of the randomised trials. Early HER2/neu cross-talk in ER/HER2-positive breast cancer, J. Natl. Cancer Inst. 96
Breast Cancer Trialists’ Collaborative Group, Lancet 351 (9114) (1998) (12) (2004) 926–935.
1451–1467. [163] S.R. Johnston, et al., Increased activator protein-1 DNA binding and c-Jun NH2-
[133] T.J. Powles, et al., Effect of tamoxifen on bone mineral density measured by dual- terminal kinase activity in human breast tumors with acquired tamoxifen re-
energy x-ray absorptiometry in healthy premenopausal and postmenopausal sistance, Clin. Cancer Res. 5 (2) (1999) 251–256.
women, J. Clin. Oncol. 14 (1) (1996) 78–84. [164] Y. Zhou, et al., Enhanced NF kappa B and AP-1 transcriptional activity associated
[134] A.E. Wakeling, M. Dukes, J. Bowler, A potent specific pure antiestrogen with with antiestrogen resistant breast cancer, BMC Cancer 7 (2007) 59.
clinical potential, Cancer Res. 51 (15) (1991) 3867–3873. [165] C.G. Murphy, M.N. Dickler, The role of CDK4/6 inhibition in breast cancer,
[135] A.E. Wakeling, Similarities and distinctions in the mode of action of different Oncologist 20 (5) (2015) 483–490.
classes of antioestrogens, Endocr. Relat. Cancer 7 (1) (2000) 17–28. [166] A.B. Pardee, G1 events and regulation of cell proliferation, Science 246 (4930)
[136] D.J. DeFriend, et al., Effects of 4-hydroxytamoxifen and a novel pure antioestrogen (1989) 603–608.
(ICI 182780) on the clonogenic growth of human breast cancer cells in vitro, Br. J. [167] D.O. Morgan, Cyclin-dependent kinases: engines, clocks, and microprocessors,
Cancer 70 (2) (1994) 204–211. Annu. Rev. Cell Dev. Biol. 13 (1997) 261–291.
[137] C.K. Osborne, et al., Comparison of the effects of a pure steroidal antiestrogen with [168] C.J. Sherr, D-type cyclins, Trends Biochem. Sci. 20 (5) (1995) 187–190.
those of tamoxifen in a model of human breast cancer, J. Natl. Cancer Inst. 87 (10) [169] G.I. Shapiro, Cyclin-dependent kinase pathways as targets for cancer treatment, J.
(1995) 746–750. Clin. Oncol. 24 (11) (2006) 1770–1783.
[138] D.J. DeFriend, et al., Investigation of a new pure antiestrogen (ICI 182780) in [170] P.N. Span, et al., Cyclin-E is a strong predictor of endocrine therapy failure in
women with primary breast cancer, Cancer Res. 54 (2) (1994) 408–414. human breast cancer, Oncogene 22 (31) (2003) 4898–4904.
[139] J.F. Robertson, et al., Comparison of the short-term biological effects of 7alpha-[9- [171] I.M. Chu, L. Hengst, J.M. Slingerland, The Cdk inhibitor p27 in human cancer:
(4,4,5,5,5-pentafluoropentylsulfinyl)-nonyl]estra-1,3,5, (10)-triene-3,17beta-diol prognostic potential and relevance to anticancer therapy, Nat. Rev. Cancer 8 (4)
(Faslodex) versus tamoxifen in postmenopausal women with primary breast (2008) 253–267.
cancer, Cancer Res. 61 (18) (2001) 6739–6746. [172] G. Perez-Tenorio, et al., Cytoplasmic p21WAF1/CIP1 correlates with Akt activa-
[140] A. Di Leo, et al., Results of the CONFIRM phase III trial comparing fulvestrant 250 tion and poor response to tamoxifen in breast cancer, Int. J. Oncol. 28 (5) (2006)
mg with fulvestrant 500 mg in postmenopausal women with estrogen receptor- 1031–1042.
positive advanced breast cancer, J. Clin. Oncol. 28 (30) (2010) 4594–4600. [173] A. Errico, Breast cancer: PALOMA-3 confirms that CDK4/6 is a key therapeutic
[141] L. Perey, et al., Clinical benefit of fulvestrant in postmenopausal women with target, Nat. Rev. Clin. Oncol. 12 (8) (2015) 436.
advanced breast cancer and primary or acquired resistance to aromatase in- [174] P.V. Schoenlein, et al., Autophagy facilitates the progression of ERalpha-positive
hibitors: final results of phase II Swiss Group for Clinical Cancer research Trial breast cancer cells to antiestrogen resistance, Autophagy 5 (3) (2009) 400–403.
(SAKK 21/00), Ann. Oncol. 18 (1) (2007) 64–69. [175] J.S. Samaddar, et al., A role for macroautophagy in protection against 4-hydro-
[142] J.N. Ingle, et al., Fulvestrant in women with advanced breast cancer after pro- xytamoxifen-induced cell death and the development of antiestrogen resistance,
gression on prior aromatase inhibitor therapy: north Central Cancer treatment Mol. Cancer Ther. 7 (9) (2008) 2977–2987.
Group Trial N0032, J. Clin. Oncol. 24 (7) (2006) 1052–1056. [176] M.J. Ellis, et al., Estrogen-independent proliferation is present in estrogen-receptor
[143] S. Chia, et al., Double-blind, randomized placebo controlled trial of fulvestrant HER2-positive primary breast cancer after neoadjuvant letrozole, J. Clin. Oncol.
compared with exemestane after prior nonsteroidal aromatase inhibitor therapy in 24 (19) (2006) 3019–3025.
postmenopausal women with hormone receptor-positive, advanced breast cancer: [177] E.M. Fox, et al., A kinome-wide screen identifies the insulin/IGF-I receptor
results from EFECT, J. Clin. Oncol. 26 (10) (2008) 1664–1670. pathway as a mechanism of escape from hormone dependence in breast cancer,
[144] A. Howell, et al., Fulvestrant, formerly ICI 182,780, is as effective as anastrozole in Cancer Res. 71 (21) (2011) 6773–6784.
postmenopausal women with advanced breast cancer progressing after prior en- [178] T. Frogne, et al., Activation of ErbB3, EGFR and Erk is essential for growth of
docrine treatment, J. Clin. Oncol. 20 (16) (2002) 3396–3403. human breast cancer cell lines with acquired resistance to fulvestrant, Breast
[145] K. Mokbel, The evolving role of aromatase inhibitors in breast cancer, Int. J. Clin. Cancer Res. Treat. 114 (2) (2009) 263–275.
Oncol. 7 (5) (2002) 279–283. [179] N. Turner, et al., FGFR1 amplification drives endocrine therapy resistance and is a
[146] T.R. Evans, et al., Phase I and endocrine study of exemestane (FCE 24304), a new therapeutic target in breast cancer, Cancer Res. 70 (5) (2010) 2085–2094.
aromatase inhibitor, in postmenopausal women, Cancer Res. 52 (21) (1992) [180] N. Shoman, et al., Reduced PTEN expression predicts relapse in patients with
5933–5939. breast carcinoma treated by tamoxifen, Mod. Pathol. 18 (2) (2005) 250–259.
[147] R.A. Yates, et al., Arimidex (ZD1033): a selective, potent inhibitor of aromatase in [181] J.A. Ligibel, et al., Body mass index, PAM50 subtype, and outcomes in node-po-
postmenopausal female volunteers, Br. J. Cancer 73 (4) (1996) 543–548. sitive breast CancerCALGB 9741 (Alliance), J. Natl. Cancer Inst. 107 (9) (2015).
[148] A. Lipton, et al., Letrozole (CGS 20267). A phase I study of a new potent oral [182] M. Protani, M. Coory, J.H. Martin, Effect of obesity on survival of women with
aromatase inhibitor of breast cancer, Cancer 75 (8) (1995) 2132–2138. breast cancer: systematic review and meta-analysis, Breast Cancer Res. Treat. 123
[149] R.C. Coombes, et al., A randomized trial of exemestane after two to three years of (3) (2010) 627–635.
tamoxifen therapy in postmenopausal women with primary breast cancer, N. Engl. [183] S. Sahin, et al., The association between body mass index and im-
J. Med. 350 (11) (2004) 1081–1092. munohistochemical subtypes in breast cancer, Breast 32 (2017) 227–236.
[150] A. Howell, et al., Results of the ATAC (Arimidex, Tamoxifen, alone or in [184] M. Baum, et al., Anastrozole alone or in combination with tamoxifen versus ta-
Combination) trial after completion of 5 years’ adjuvant treatment for breast moxifen alone for adjuvant treatment of postmenopausal women with early-stage
cancer, Lancet 365 (9453) (2005) 60–62. breast cancer: results of the ATAC (Arimidex, Tamoxifen alone or in Combination)
[151] P.E. Goss, et al., A randomized trial of letrozole in postmenopausal women after trial efficacy and safety update analyses, Cancer 98 (9) (2003) 1802–1810.
five years of tamoxifen therapy for early-stage breast cancer, N. Engl. J. Med. 349 [185] M. Baum, et al., Anastrozole alone or in combination with tamoxifen versus ta-
(19) (2003) 1793–1802. moxifen alone for adjuvant treatment of postmenopausal women with early breast

169
P. Bhardwaj, et al. Journal of Steroid Biochemistry and Molecular Biology 189 (2019) 161–170

cancer: first results of the ATAC randomised trial, Lancet 359 (9324) (2002) (+) breast cancer, Clin. Cancer Res. 23 (20) (2017) 6138–6150.
2131–2139. [193] E.A. Wellberg, et al., FGFR1 underlies obesity-associated progression of estrogen
[186] M. Gnant, et al., The predictive impact of body mass index on the efficacy of receptor-positive breast cancer after estrogen deprivation, JCI Insight 3 (14)
extended adjuvant endocrine treatment with anastrozole in postmenopausal pa- (2018).
tients with breast cancer: an analysis of the randomised ABCSG-6a trial, Br. J. [194] K. Laud, et al., Identification of leptin receptors in human breast cancer: functional
Cancer 109 (3) (2013) 589–596. activity in the T47-D breast cancer cell line, Mol. Cell. Endocrinol. 188 (1-2)
[187] D.S. Chan, et al., Body mass index and survival in women with breast cancer- (2002) 219–226.
systematic literature review and meta-analysis of 82 follow-up studies, Ann. [195] S. Catalano, et al., Leptin induces, via ERK1/ERK2 signal, functional activation of
Oncol. 25 (10) (2014) 1901–1914. estrogen receptor alpha in MCF-7 cells, J. Biol. Chem. 279 (19) (2004)
[188] K.M. Elliott, et al., Effects of aromatase inhibitors and body mass index on steroid 19908–19915.
hormone levels in women with early and advanced breast cancer, Br. J. Surg. 101 [196] R.V. Considine, et al., Serum immunoreactive-leptin concentrations in normal-
(8) (2014) 939–948. weight and obese humans, N. Engl. J. Med. 334 (5) (1996) 292–295.
[189] P.E. Lonning, B.P. Haynes, M. Dowsett, Relationship of body mass index with [197] S. Catalano, et al., Leptin enhances, via AP-1, expression of aromatase in the MCF-
aromatisation and plasma and tissue oestrogen levels in postmenopausal breast 7 cell line, J. Biol. Chem. 278 (31) (2003) 28668–28676.
cancer patients treated with aromatase inhibitors, Eur. J. Cancer 50 (6) (2014) [198] E.B. Coscia, et al., Estrone and estradiol levels in breast cancer patients using
1055–1064. anastrozole are not related to body mass index, Rev. Bras. Ginecol. Obstet. 39 (1)
[190] I. Sestak, et al., Effect of body mass index on recurrences in tamoxifen and ana- (2017) 14–20.
strozole treated women: an exploratory analysis from the ATAC trial, J. Clin. [199] L. Zewenghiel, H. Lindman, A. Valachis, Impact of body mass index on the efficacy
Oncol. 28 (21) (2010) 3411–3415. of endocrine therapy in patients with metastatic breast cancer - a retrospective
[191] L.W. Bowers, et al., Obesity enhances nongenomic estrogen receptor crosstalk with two-center cohort study, Breast 40 (2018) 136–140.
the PI3K/Akt and MAPK pathways to promote in vitro measures of breast cancer [200] J. Zekri, K. Farag, A. Allithy, Obesity and outcome of post-menopausal women
progression, Breast Cancer Res. 15 (4) (2013) R59. receiving adjuvant letrozole for breast cancer, Ecancermedicalscience 12 (2018)
[192] L. Formisano, et al., Association of FGFR1 with ERalpha maintains ligand-in- 821.
dependent ER transcription and mediates resistance to estrogen deprivation in ER

170

You might also like