You are on page 1of 36

Obesity-associated Breast Cancer:

Analysis of risk factors 25


Atilla Engin

Abstract

Several studies show that a significantly stronger association is obvious


between increased body mass index (BMI) and higher breast cancer inci-
dence. Furthermore, obese women are at higher risk of all-cause and breast
cancer specific mortality when compared to non-obese women with breast
cancer. In this context, increased levels of estrogens due to excessive aroma-
tization activity of the adipose tissue, overexpression of pro-­inflammatory
cytokines, insulin resistance, hyperactivation of insulin-like growth factors
(IGFs) pathways, adipocyte-derived adipokines, hypercholesterolemia and
excessive oxidative stress contribute to the development of breast cancer in
obese women. While higher breast cancer risk with hormone replacement
therapy is particularly evident among lean women, in postmenopausal
women who are not taking exogenous hormones, general obesity is a signifi-
cant predictor for breast cancer. Moreover, increased plasma cholesterol
leads to accelerated tumor formation and exacerbates their aggressiveness.
In contrast to postmenopausal women, premenopausal women with high
BMI are inversely associated with breast cancer risk. Nevertheless, life-style
of women for breast cancer risk is regulated by avoiding the overweight and
a high-fat diet. Estrogen-plus-progestin hormone therapy users for more
than 5 years have elevated risks of both invasive ductal and lobular breast
cancer. Additionally, these cases are more commonly node-positive and
have a higher cancer-related mortality. Collectively, in this chapter, the
impacts of obesity-related estrogen, cholesterol, saturated fatty acid, leptin
and adiponectin concentrations, aromatase activity, leptin and insulin resis-
tance on breast cancer patients are evaluated. Obesity-related prognostic
factors of breast cancer also are discussed at molecular basis.

A. Engin, M.D., Ph.D. (*)


Faculty of Medicine, Department of General Surgery,
Gazi University, Besevler, Ankara, Turkey
Mustafa Kemal Mah. 2137. Sok. 8/14, 06520,
Cankaya, Ankara, Turkey
e-mail: dr.aengin@gmail.com

© Springer International Publishing AG 2017 571


A.B. Engin, A. Engin (eds.), Obesity and Lipotoxicity, Advances in Experimental
Medicine and Biology 960, DOI 10.1007/978-3-319-48382-5_25
572 A. Engin

Keywords
Breast cancer • Obesity • Fatty acid • Postmenopausal breast cancer •
Premenopausal breast cancer • Estrogen • Leptin • Adiponectin • Aromatase
• Insulin resistance • Mammalian target of rapamycin (mTOR) • Insulin-
like growth factor (IGF) • Human epidermal growth factor receptor 2
(HER2) • Estrogen receptor (ER) • Progesteron receptor (PR) • Cyclin D1
• Liver X receptor (LXR) • Oxysterol 27-hydroxycholesterol (27HC) •
Sex hormone-binding globulin (SHBG) • Estrone • Hypoxia •
Hyperinsulinemia • Aryl hydrocarbon receptor (AHR) • Breast cancer
recurrence • Cancer-associated adipocytes (CAAs) • Leptin resistance

and breast cancer specific mortality when com-


1 Introduction pared to non-obese women with breast cancer
(Demark-Wahnefried et al. 2012). In this respect,
High body mass index (BMI) is associated with overweight women should avoid weight gain
increased risk of some cancer types. In 1402 of during treatment and obese women should lose
108,812 individuals from the general population weight immediately after treatment (Rock et al.
developed breast cancer during a median of 4.7-­ 2012). Of 53,816 women treated for early-stage
years follow-up. In this series of patients, corre- breast cancer, 18,967 patients with complete fol-
sponding risk of breast cancer was 20% higher in low-­up have been evaluated up to 10 years and
postmenopausal women with high BMI (Benn up to 30 years for the loco-regional recurrences
et al. 2016). However, a significantly stronger or distant metastases and for death, respectively.
association between increased BMI and higher The risk of developing distant metastases after 10
breast cancer incidence is observed in the Asia-­ years was significantly increased which was by
Pacific group than in European-Australian or 46%, and the risk of dying as a result of breast
North-American group (Wang et al. 2016). In the cancer after 30 years was significantly increased
Asia-Pacific region, every additional 5 kg/m2 which was by 38% for patients with a BMI of 30
increase in BMI corresponds to a 31% increase in kg/m2 or more. However, BMI had no influence
postmenopausal breast cancer risk (Renehan on the risk of loco-regional recurrences (Ewertz
et al. 2008). Although overweight is related to a et al. 2011). Indeed, high BMI is associated with
higher risk of mortality, the number of obese worse long-term outcomes among obese women,
female breast cancer survivors is growing as a who have breast cancer (Kamineni et al. 2013).
consequence of advances in treatment strategies In addition, increase in post-diagnostic body
and early diagnosis. In this respect, the number of weight is common in women with breast cancer.
breast cancer survivors is increasing by approxi- In this respect, particularly chemotherapy treat-
mately 3% each year (Maddams et al. 2009). ment has been considered as a significant con-
Evaluation of 79 publications from 82 follow-up tributory ­ factor through reduced metabolism
studies clearly indicate that, amongst the breast (Demark-­ Wahnefried et al. 2012). Inter-linked
cancer survivors, higher BMI is consistently molecular mechanisms between obesity and
associated with lower overall and breast cancer breast cancer might be involved in the pathogen-
survival. Analysis of 23,182 deaths of 213,075 esis in postmenopausal women. In this context,
breast cancer survivors confirmed that the current increased levels of estrogens due to excessive
guideline for breast cancer survivors is to remain aromatization activity of the adipose tissue,
as lean as possible within the normal range of ­overexpression of inflammatory cytokines, insu-
body weight (Chan et al. 2014). Another meta-­ lin resistance, hyperactivation of insulin-like
analysis includes results from 43 studies show growth factors (IGFs) pathways, adipocyte-
that obese women are at higher risk of all-cause derived adipokines, and excessive oxidative
25 Obesity-associated Breast Cancer: Analysis of risk factors 573

stress contribute to the development of breast increase in risk of breast cancer among post-
cancer in obese women. These factors also menopausal non-hormone replacement therapy
­interfere with intracellular signaling in the mito- users (Lahmann et al. 2004). Considering the
gen-activated protein kinase (MAPK) and phos- relationship between breast cancer and obesity;
phatydilinositol-3-phosphate (PI3P)/mammalian the postulated mechanisms for the increased risk
target of rapamycin (mTOR) pathway (Simone of breast cancer in obese women, such as elevated
et al. 2016). estrogen levels, insulin resistance and the influ-
The interaction between human-derived adi- ences of IGF, fatty acids, leptin and adiponectin
pocytes and primary breast cancer cells increase will be discussed in this chapter.
the secretion of pro-inflammatory cytokines.
Furthermore, contact with immature adipocytes
increase the abundance of cancer cells with 2  strogens as a Cause
E
tumor-forming and metastatic potential in vivo of Breast Cancer in Obesity
(Picon-Ruiz et al. 2016). The expression level of
the fat mass-obesity associated genes in breast There is abundance of data linking obesity-­
cancer tissues is significantly higher than that in related breast cancer to estrogen receptor signal-
the adjacent healthy breast tissues. However, ing. As comorbidities of obesity; excessive local
there is no correlation between fat mass-obesity production of estrogens in adipose tissue, the
associated gene expression and age, tumor stage, influence of adipokines and inflammatory cyto-
lymph node status, TNM stage, Ki67, and BMI in kines, finally hypercholesterolemia have also
breast cancer. Despite all these, fat mass-obesity been established as independent risk factors for
associated gene expression has very important breast cancer in postmenopausal women (Boyd
role in human epidermal growth factor receptor 2 and McGuire 1990; McDonnell et al. 2014a).
(HER2)-overexpressed breast cancer (Tan et al. Despite accumulated observational evidences,
2015). In the women undergoing mastectomy for risks and benefits of estrogen use are evaluated
breast cancer risk reduction or for breast cancer with the increase in absolute excess risk for
treatment, white adipose tissue inflammation in occurrence of invasive breast cancers. On May
breast specimens is detected in 52% and 41% of 31, 2002, after a mean of 5.2 years of follow-up,
patients, respectively. In these patients, white the data and safety monitoring board recom-
adipose tissue inflammation is defined by the mended stopping the trial of estrogen-plus-­
presence of dead adipocytes surrounded by mac- progestin versus placebo, because the test
rophages forming crown-like structures in the statistic for invasive breast cancer exceeded the
breast adipose tissue (Iyengar et al. 2016). average incidence of this adverse effect (Rossouw
During the 4.7 years-follow-up of 73,542 pre- et al. 2002). Over an average follow-up of 6.8
menopausal and 103,344 postmenopausal women years, the use of 0.625 mg/day of conjugated
from nine European countries, 1879 incidental equine estrogen increases the risk of stroke,
invasive breast cancers have been diagnosed. decreases the risk of hip fracture, and does not
While excess breast cancer risk with hormone affect coronary heart disease incidence in post-
replacement therapy is particularly evident menopausal women with prior hysterectomy.
among lean women, in postmenopausal women Finally, in February 2004, the National Institutes
who are not taking exogenous hormones, general of Health decided to end the intervention phase
obesity is a significant predictor of breast cancer. of the trial early by taking into consideration of
In contrast to postmenopausal women, among these data (Anderson et al. 2004). In a total of
premenopausal women, high BMI is inversely, 16,608 women without hysterectomy and ran-
but not significantly, associated with breast can- domized to the estrogen-plus-progestin trial, the
cer risk. This meta-analysis indicates a 2% reduc- higher breast cancer risk seen during intervention
tion in risk per unit of increase in BMI (1 kg/m2) that is followed by a substantial drop in the risk
among premenopausal women, whereas a 3% of the early post-intervention phase, but the
574 A. Engin

higher breast cancer risk remains during the late lar and ER positive-PR positive tumors compared
post-­intervention phase of follow-up period. The to never users of hormone therapy regardless of
10,739 women with prior hysterectomy are ran- BMI. Current estrogen-plus-progestin hormone
domized according to the estrogen alone trial. therapy users for more than 5 years with a BMI
Despite the lower breast cancer risk that is less than 24.9 kg/m2 also have a 2.6-fold elevated
observed during the early post-intervention risk of ductal breast cancer (Li et al. 2006). In
phase, during the late post-intervention period a eleven original articles, the case-case comparison
higher breast cancer risk is determined showed a significant association between triple
(Chlebowski et al. 2015). Collectively, the results negative breast cancer and obesity. When the
of these data indicate that in estrogen-plus-­ patients are stratified based on menopausal sta-
progestin group, breast cancer incidence is tus, significant results are observed only in the
higher, and the cases are more commonly node- premenopausal group and triple negative tumors
positive. Breast cancer mortality also appears to with ER negative-PR negative/HER2-found to
be increased with combined use of estrogen-plus-­ have positive associations with obesity but only
progestin (Chlebowski et al. 2010). Breast cancer among premenopausal women (Pierobon and
incidence rate decreases with body mass among Frankenfeld 2013). In a prospective cohort of
premenopausal women in high-risk countries, 67,754 postmenopausal women, 1821 cases of
but this ratio rises with the increase in body mass invasive ductal breast cancer and 471 cases of
in all other groups of women (Pathak and invasive lobular or mixed lobular cancer occurred
Whittemore 1992). More detailed analysis during 13 years of follow-up. Use of exogenous
revealed that decreased risk of breast cancer estrogen plus progesterone is associated with an
associates with increased body size among pre- increased risk of both ductal and lobular breast
menopausal women, however this event is lim- cancer. Risk increased within the first 2–3 years
ited to the youngest age group only (Peacock of use and attenuated 2 years after cessation. In
et al. 1999). Furthermore, obesity may influence contrast, use of estrogen-only is not associated
the levels of endogenous sex-steroid and IGF-­ with an overall increased risk of invasive ductal
related hormones in the circulation, especially cancer. Estrogen-only use is associated with a
after menopause. In the post-menopausal group, 50% increased risk of invasive lobular cancer
increases in estrogens, testosterone and andro- after more than 10 years of use (Calle et al. 2009).
stenedione are correlated with increasing BMI. In In this case, estrogens are converted to metabo-
contrast, sex hormone-binding globulin decreases lites, particularly the catechol estrogen-3,4-­
with increasing BMI (Lukanova et al. 2004; quinones (CE-3,4-Q), that can react with DNA to
McTiernan et al. 2006). Obese women have 35% form depurinating adducts. These adducts are
higher concentrations of estrone and 130% higher released from DNA to generate apurinic sites.
concentrations of estradiol compared with Depurinating adducts 4-hydroxyestrone (estra-
normal-­weight women. Testosterone concentra- diol), 4-OHE1(E2)-1-N3Ade and 4-OHE1(E2)-
tions also increase with increasing levels of adi- 1-N7Gua constitute more than 99% of the total
posity. Concentrations of free estradiol and free DNA adducts formed. Increased levels of these
testosterone are two to three times greater in quinones and their reaction with DNA occur
overweight and obese women compared with when estrogen metabolism is unbalanced
normal-weight women (McTiernan et al. 2003). (Cavalieri et al. 2006). Risk for ER negative-PR
Obese never users of hormone therapy have 1.7- negative tumors among postmenopausal women
to 2.3-fold elevated risks of ductal and estrogen slightly increases, but this is significantly differ-
receptor (ER) positive-progesteron receptor (pos- ent from risk for ER positive-PR positive tumors.
itive-­
PR) positive breast cancer, respectively, The association between adult weight gain and
compared to thinner women. Estrogen-plus-­ postmenopausal breast cancer risk is heteroge-
progestin hormone therapy users for more than 5 neous according to ER/PR status and stronger for
years have 2.1 to 9.6-fold elevated risks of lobu-
25 Obesity-associated Breast Cancer: Analysis of risk factors 575

ER positive-PR positive than for ER negative-PR ing pathways (Bowers et al. 2013). Indeed,
negative tumors (Vrieling et al. 2010). increased phosphatidylinositol 3-kinase (PI3K)
activity and Akt phosphorylation is a favorable
setting for mammary tumor growth and metastasis
3  holesterol and Breast
C in hypercholesterolemic animals (Alikhani et al.
Cancer in Obesity 2013) (Fig. 25.1).
Actually, the oxysterol 27-hydroxycholesterol
In association with oncogenic stimuli, increased (27HC) is synthesized from cholesterol by
plasma cholesterol leads to accelerated tumor CYP27A1 which is a rate limiting enzyme. Hence
formation and increases tumor burden. In mam- in humans, the circulating levels of 27HC closely
mary tumors increased plasma cholesterol levels reflects those of cholesterol and hypercholesterol-
are associated with increased expression of cyclin emia (Karuna et al. 2011) 27HC is a liver X recep-
D1, a marker associated with tumor formation, tor (LXR) agonist and as such serves to limit
and decreased expression of markers associated cholesterol accumulation in cells (Fu et al. 2001)
with protection. Both high density lipoproteins LXRs, as antiproliferative factors suppress growth
and scavenger receptor type BI proteins are ele- of both normal and breast cancer cells (Vigushin
vated in animals that are fed a cholesterol-rich et al. 2004). Interestingly, LXRs might have a
diet. An increase in plasma cholesterol levels dual role in this respect. On the one hand, LXRs
accelerates the development of tumors and exac- are antiproliferative in the breast cancer cells. On
erbates their aggressiveness (Llaverias et al. the other hand, LXR-induced expression of lipo-
2011). The increased availability of high density genic genes peaked at lower 27HC concentration
lipoproteins may increase the estradiol access to are needed for the antiproliferative effect of LXR
the cancer tissue. While lipoproteins may regu- (Vedin et al. 2009). Nevertheless, genetic or phar-
late endothelial function in the tumor, they may macological activation of LXR results in estro-
also affect the transport of estrogen. In addition, gen deactivation, which in turn inhibits breast
scavenger receptor type BI may promote estra- cancer growth (Gong et al. 2007). In contrast,
diol uptake (Brodeur et al. 2008). The binding of 27HC is an estrogen receptor agonist in breast
high density lipoprotein (HDL) to scavenger cancer cells and it stimulates the growth and
receptor type BI activates Ras in a protein kinase metastasis of tumors in breast cancer of obese
C (PKC)-independent manner with subsequent women (McDonnell et al. 2014b). In human
induction of the MAPK signaling cascade breast cancer specimens, tumor grade is corre-
(Grewal et al. 2003). Many metabolic pathways, lated with CYP27A1 expression levels. In high-
which have critical roles in progression of breast grade tumors, both tumor cells and
cancer may induce MAPK signaling cascade. In tumor-associated macrophages display high
this respect, fatty acid binding protein 4 (FABP4) expression levels of this enzyme. Thus, 27HC
enhances the proliferation of breast cancer cells actually promotes the proliferation of ER-positive,
by inducing the Akt and MAPK signaling cas- but not ER-negative breast cancer cell lines
cades (Guaita-Esteruelas et al. 2016). Adiponectin (Nelson et al. 2013). In ER-positive breast cancer
receptor 1 (AdipoR1), adaptor protein APPL1, patients, 27HC content in normal breast tissue is
ER-alpha, Insulin like growth factor 1 receptor increased compared to the cancer-­free controls,
(IGF-R1), and c-Src that is also responsible for and the tumor 27HC abundance is further ele-
MAPK signaling activation in ER-delta positive vated. Cancer 27HC content is not influenced by
breast cancer cells (Mauro et al. 2014). circulating levels of 27HC or its precursors.
Furthermore, circulating factors in the serum of Increases in tumor 27HC are directly related to
obese postmenopausal women stimulate diminished expression of the 27HC metabolizing
ER-alpha positive breast cancer cell viability and enzyme CYP7B1. Actually, 27HC is a locally-
growth by facilitating non-genomic ER-alpha modulated, non-aromatized ER ligand that pro-
crosstalk with the PI3K/Akt and MAPK signal- motes ER positive breast cancer growth (Wu et al.
576 A. Engin

Fig. 25.1 The postulated mechanisms for the increased IGFBP insulin-like growth factor binding protein, CYP19
risk of breast cancer in obese women; elevated estrogen aromatase cytochrome P450, product of the CYP19 gene,
levels, insulin resistance and the influences of IGF, fatty CYP1B1 cytochrome P-450 1B1, CYP1A1 cytochrome
acids, leptin and adiponectin (ROS reactive oxygen radi- P-450 1A1, 2-OHE2 2-hydroxy estradiol, 4-OHE2
cals, PI3K phosphatidylinositol 3-kinase, MAPK mitogen-­ 4-hydroxyestradiol, ER-alpha estrogen receptor-alpha,
activated protein-kinase, Akt protein kinase B, FFA free ER-beta estrogen receptor-beta, STAT3 signal transducers
fatty acid, SHBG sex hormone-binding globulin, TNF-­ and activators of transcription-3, ERK extracellular signal-­
alpha tumor necrosis factor-alpha, IL-6 interleukin-6, regulated kinase, AP-1 activating protein-1, MAPK
NF-kappaB nuclear factor-kappaB, IKK inhibitor kappa B mitogen-­activated protein-kinase, TLR4 Toll like recep-
kinase, mTOR mammalian target of rapamycin, VEGF tor-­4, COX-2 cyclooxygenase-2, LXR liver X receptor,
vascular endothelial growth factor, PAI-1 plasminogen Adipo R1/R2 adiponectin receptors, FABP4 fatty acid
activator inhibitor-1, IGF-1 insulin-like growth factor-1, binding protein 4)

2013). Interestingly, tumor infiltration by macro- may play a role in the pathogenesis of breast can-
phages is associated with their ability to produce cer that operates independent of their
27HC within tumors and induce ER-alpha activity well-­
­ documented function in binding the ER
(McDonnell et al. 2014a) (Fig. 25.1). molecules expressed by breast cancer cells. In
this context, despite the lack of clinical responses
of ER-negative breast cancers to tamoxifen ther-
4 Obesity-Related Aromatase apy, aromatase inhibitor is able to block tumor
Activity in Breast Cancer growth. Actually, estrogen also enhances the
growth of ER-negative tumors by increasing the
While the ovaries are the primary source of estro- systemic capacity for angiogenesis and the
gens in premenopausal women, estrogen synthe- recruitment of bone marrow-derived stromal
sis also occurs in the adipose tissue. Estrogens cells (Gupta et al. 2007). Breast cancer patients
25 Obesity-associated Breast Cancer: Analysis of risk factors 577

have increased exposure to unbound circulating mal part of breast tumors (Sasano and Harada
estradiol. An increased percentage of estradiol 1998), as well as in epithelial location. There is a
bound to albumin may influence the availability significant, but inverse, correlation between the
of estradiol, considering its low binding affinity aromatase activity and the estrogen receptor sta-
to albumin (Ota et al. 1986). Weight loss as a tus, indicating the likelihood of negative estrogen
result of caloric restriction and exercise signifi- receptors if substantial aromatase activity is pres-
cantly decreases estrogens and free testosterone, ent (Esteban et al. 1992). Three-fold higher
and increases sex hormone-binding globulin human aromatase mRNA levels are found in the
(SHBG) (Campbell et al. 2012). In contrast, mammary gland of female mice that are fed with
increasing obesity correlates with a progressive a high fat diet compared with their littermates on
fall in SHBG level. Thus increasing upper body normal chow. Because overweight females have
fat localization is inversely proportional to levels twice the mammary gland mass as lean controls,
of SHBG in breast cancer patients and healthy thereupon, the overweight mice could have up to
individuals. Premenopausal breast cancer patients six-­fold higher human aromatase mRNA in their
are found to have significantly lower levels of mammary gland compared with lean counter-
SHBG compared with age-matched and weight- parts. Hence, weight loss may reduce breast can-
matched individuals. Lower levels of SHBG also cer risk via reduction in local aromatase
result in increased levels of circulating unbound production of the breast in premenopausal obese
androgens which may result in tumor progression women (Chen et al. 2012). As mentioned above,
by themselves and further conversion to estro- epithelial cells are the primary site of estrogen
gens by adipose tissue (Schapira et al. 1991). synthesis in the breast and breast cancers.
Thus, the levels of estradiol that are detected in Aromatase expression is detected in the cyto-
breast cancers are 50 to 100-fold higher than the plasm of tumor epithelial cells and the surround-
concentrations of this hormone that are found in ing stromal cells of over 50% of tumors. Although
the circulation of postmenopausal women. The intra-tumoral aromatase activity does not corre-
high estradiol concentrations found in breast can- late with steroid receptors, ER-positive tumors
cers could arise either due to uptake from the cir- express aromatase. These data indicate that some
culation or from endogenous synthesis. Since breast cancers synthesize sufficient estrogen to
similar concentrations of estradiol are present in stimulate their own proliferation (Brodie et al.
tumors with or without estrogen receptors, 1997). Assays of aromatase levels in adipose tis-
endogenous synthesis is considered to be the sue from the different quadrants of mastectomy
major pathway by which tumor estrogens origi- specimens from patients with breast cancer indi-
nate. Three main enzyme complexes are involved cate that activity is always higher in breast quad-
in the synthesis of estrogens in peripheral tissues. rants associated with tumor as compared with
Firstly, aromatase is responsible for the aromati- non-involved quadrants. These results emphasize
sation of androstenedione to estrone. Secondly, the importance of local estrogen synthesis within
estrone sulfatase catalyses the formation of the breast (Miller and O’Neill 1987). Loss of the
estrone from estrone sulfate, and finally estradiol- ovarian function to supply estrogen and
17β-hydroxysteroid dehydrogenase Type 1 is ­progesterone after menopause can cause deregu-
responsible for the reduction of estrone to the lation of the body’s metabolism and the availabil-
biologically active estrogen and estradiol (Purohit ity of systemic estradiol dramatically decreases
et al. 2002). The aromatase enzyme, estrogen (Boonyaratanakornkit and Pateetin 2015).
synthase (CYP19 p450), belongs to a family of Elevated in situ estrogen concentrations in post-
p450 enzymes (Simpson et al. 1981). In post- menopausal human breast cancer patients are
menopausal women, it is located mainly in adi- derived from intratumoral aromatization.
pose tissue but is also present in normal and Furthermore, in patients with estrogen-dependent
malignant breast tissues. Immunocytochemical breast carcinoma, intratumoral estrogens func-
studies indicate that aromatase settles in the stro- tion as an autocrine growth and a mitogenic fac-
578 A. Engin

tor and could impart a growth advantage to these in increased aromatase activity and subsequently
cancer cells, regardless of serum concentration of increased estrogen biosynthesis through auto-
estrogens (Sasano et al. 2006). Furthermore, in crine mechanisms in breast epithelial cells and
postmenopausal breast cancer patients, up to through paracrine mechanisms in breast stromal
50% of deaths have been attributed to obesity cells (Richards et al. 2002). Furthermore, COX-­
(van Kruijsdijk et al. 2009). As mentioned above, 2-­derived PGE2 stimulates the cyclic adenosine
aromatase activity is highest in adipose tissue of monophosphate (cAMP)-phosphokinase A (PKA)
the breast quadrant containing the tumor. In these signal transduction pathway that activates CYP19
breast tissues, a significant correlation is also transcription. Consequently, higher aromatase
found between interleukin-6 (IL-6) production expression and increased progesterone receptor
and aromatase activity (Reed et al. 1993). In levels are observed in breast tissues of overweight
accordance with declining systemic estradiol at and obese women (Subbaramaiah et al. 2012).
menopause, the incidence of ER-alpha-positive Actually COX-2 directly regulates gene expres-
breast cancer dramatically increases (Pfeilschifter sion of specific aromatase promoter regions of
et al. 2002; Sasser et al. 2007). BMI, leptin, IL-6 the p450 Cyp19 enzyme, aromatase and its prod-
and reactive oxygen species (ROS) is higher in uct, estradiol. Inhibition of COX-2 or blocking
ER positive compared with ER negative patients. the biological effects of PGE2 may be useful in
Among ER positive patients, BMI, leptin, IL-6 significantly limiting aromatase activity and pro-
and ROS correlate with tumor size and metasta- liferation of human breast tumor cells despite
sis. Additionally, leptin, IL-6 and ROS positively the presence of COX-2 protein (Prosperi and
correlate also with lymph node metastasis. Robertson 2006).
Eventually, weight gain, inflammation and oxida- Estrogen or other selective estrogen receptor
tive stress are involved in post-menopausal modulators (SERMs) bind to the ER, a ligand-­
estrogen-­dependent breast cancer prognosis activated transcription factor that regulates tran-
(Madeddu et al. 2014). Evidently, obesity is asso- scription of target genes in the nucleus by binding
ciated with a worse breast cancer prognosis, par- to estrogen response element (ERE). The ER
ticularly in ER-alpha-positive, postmenopausal exists in two main forms, ER-alpha and ER-beta,
patients. Obesity-associated systemic IL-6 indi- which have distinct tissue expression patterns
rectly enhances preadipocyte aromatase expres- (Mueller and Korach 2001). ERs are expressed in
sion via increased breast cancer cell prostaglandin white adipose tissue, and play an important role
E2 (PGE2) production (Bowers et al. 2015). in regulating white adipose tissue in females.
Increased nuclear factor-kappaB (NF-kappaB) Absence of ER-alpha causes adipocyte hyperpla-
binding activity and elevated aromatase expres- sia and hypertrophy in white adipose tissue and is
sion and activity are found due to chronic inflam- accompanied by insulin resistance and glucose
mation in breast tissue of overweight and obese intolerance in both males and females (Cooke
women. The severity of breast inflammation of et al. 2001). Obese patients have 35% higher lev-
obese women is defined as the crown-like struc- els of estrone and 130% higher concentrations of
tures of the breast index and this severity status is estradiol. Concentrations of free estradiol and
correlated with both BMI and adipocyte size. The free testosterone are two to three times greater in
obesity-inflammation-aromatase axis is an indi- overweight and obese women compared with
cator of the increased risk of hormone receptor-­ lean-matched controls (McTiernan et al. 2003). A
positive breast cancer and is a sign of poor traditional hypothesis for how obesity affects
prognosis in breast cancer of overweight and breast cancer recurrence risk is based on the
obese women (Morris et al. 2011; Subbaramaiah mitogenic effect of excess estrogen produced in
et al. 2012). Indeed, increased secretion of pros- fat cells via the enzymatic conversion of adrenal
taglandins via constitutive cyclooxygenase steroids to estrogen (Dignam et al. 2003)
(COX)-1 and inducible COX-2 isozymes in the (Fig. 25.1). In addition, cytokines secreted by the
breast cancer tissue microenvironment can result adipocytes can upregulate the aromatase enzyme
25 Obesity-associated Breast Cancer: Analysis of risk factors 579

to further increase the estrogen production, which mechanisms, either through the estrogen recep-
may stimulate breast cancer cell growth (Cleary tor, which is located in or adjacent to the plasma
and Grossmann 2009). Indeed, obesity is associ- membrane, or through other non-estrogen recep-
ated with increased adipose tissue mass and aro- tor, plasma membrane-associated estrogen-­
matase activity, thereby aromatase inhibitors are binding proteins. Eventually, estrogen-mediated
less effective in women who are overweight or cellular responses are increased levels of Ca2+ or
obese (Goodwin 2013). For women with lymph nitric oxide (NO), and activation of kinases
node-negative, ER-positive breast cancer, tamox- (Deroo and Korach 2006). In the second process,
ifen might be offered as the best approach for the estradiol is metabolized to quinone derivatives,
breast cancer risk reduction among obese women, which directly remove base pairs from DNA
because obesity is associated with increased risks through depurination. The third pathway involv-
of contralateral breast cancer, other primary can- ing estrogen metabolism is mediated by cyto-
cers, and overall mortality (Dignam et al. 2003). chrome p450. Reactive electrophilic estrogen
However, in postmenopausal women with early-­ o-quinones and ROS are generated through redox
stage breast cancer and high BMI, after a median cycling of o-quinones (Bolton and Thatcher 2008).
of 8.7 years of follow-up, obesity shown to be All these data indicate that both ER-dependent
an independent adverse prognostic factor for and genotoxic ER-independent effects of estra-
death from breast cancer. Aromatase inhibitors diol mediate breast cancer development (Santen
are more effective than tamoxifen in reducing et al. 2009). The hormonal changes across the
overall deaths, breast cancer recurrences, and dis- perimenopause and increased risk of obesity at
tant metastases (Ewertz et al. 2012). In any way, menopause may be caused by the decrease in cir-
risk factors associated with breast cancer reflect culating estrogen, and, for fat distribution shifts,
cumulative exposure of the breast epithelium to the relative increase in the androgen-estrogen
estrogen (Henderson and Feigelson 2000). Two ratio (Davis et al. 2012; Lovejoy 2003).
current hypotheses exist to explain this relation-
ship (Yue et al. 2005). In the first hypothesis;
according to genomic mechanism, estrogens dif- 5  reast Cancer Recurrence
B
fuse into the cell and bind to the estrogen recep- in Obesity
tor, which is located in the nucleus. This nuclear
estrogen-ER complex binds to ERE sequences Obesity has been repeatedly shown to be a risk
directly or indirectly through protein-protein factor for breast cancer recurrence and poor sur-
interactions with activator protein 1 (AP1) in the vival. In a meta-analysis of 43 studies, compared
promoter region of estrogen-responsive genes with non-obese, obese women with breast cancer
(Deroo and Korach 2006). Thereby, estrogen showed to have poorer survival which is similar
modulates this mechanism by two separate pro- for overall and breast cancer specific survival.
cesses. One involves the binding of estradiol to This association does not differ by menopausal
ER-alpha with stimulation of cell proliferation. status (Protani et al. 2010). Adjusting for clinical
Consequently, errors in DNA occurring during factors and potential confounders, 10% weight
replication result in fixed mutations when not gain and obesity are associated with increased
repaired. The other process results from the for- risk of late recurrence (Nechuta et al. 2016).
mation of genotoxic metabolites of estradiol, Compared to normal-weight women, obese
which can bind to DNA, cause depurination, and women have experience of increased risk of
result in mutations (Santen et al. 2009). In brief, recurrence and risk of breast cancer death,
estrogens increase the rate of cell proliferation by 12.2% and 6.9%, respectively within 10 years of
stimulating ER-mediated transcription. Thus diagnosis. Although, there is no association
increasing number of errors occur during DNA between BMI and all-cause mortality, obese
replication (Yue et al. 2005). At the same time, women have significantly faster growing tumors
estrogen can act more quickly via non-genomic (Kamineni et al. 2013). The cohort consisted of
580 A. Engin

3385 women enrolled in National Surgical Actually, high BMI is significantly associated
Adjuvant Breast and Bowel Project (NSABP) with larger tumor size both in pre and postmeno-
protocol B-14 showed that obesity is not associ- pausal women. Obese premenopausal women
ated with an increase in recurrence risk in lymph show worse histopathologic features including
node-­negative, ER-positive breast cancer cases. more metastatic axillary lymph nodes, and pres-
However, obesity is associated with increased ence of vascular invasion, compared to under or
risks of contralateral breast cancer and of over- normal weight group. Postmenopausal patients
all mortality (Dignam et al. 2003). In 153 case-­ with BMI less than 25 develop more frequently
control pairs of perimenopausal and estrogen and progesterone receptor positive can-
postmenopausal women, total estradiol, bio- cers, while no association is found in premeno-
available estradiol and free estradiol concentra- pausal women (Biglia et al. 2013). There are no
tions are significantly associated with risk for differences in estradiol levels between lean and
recurrence. Recurred women have an average obese women under aromatase inhibitors. The
total estradiol concentration that is double than known impact of obesity on recurrence risk in
that of non-­recurred women (Rock et al. 2008). women under aromatase inhibitors treatment
Serum levels of soluble IL-6 receptor are lower may not be due to incomplete aromatase inhibi-
in patients with ER-positive cancer. However, tion (Diorio et al. 2012). Aromatase inhibitors
higher serum levels of soluble IL-6 receptor at are less efficient at suppressing estradiol serum
diagnosis are associated with significantly levels in obese when compared with non-obese
shorter relapse-free survival in patients with women (Pfeiler et al. 2013). The higher levels of
ER-positive breast cancer (Won et al. 2013). estrogens in overweight postmenopausal breast
Direct application of IL-6 on breast cancer cells cancer patients before and during aromatase inhi-
inhibits proliferation in ER-positive cells, while bition may be due to the effects of BMI on estro-
high circulating IL-6 levels are correlated with a gen metabolism rather than aromatization
poor prognosis in breast cancer patients (Lønning et al. 2014). Actually, weight gain is a
(Dethlefsen et al. 2013). Interleukins could significant determinant of recurrences. In multi-
stimulate many signaling pathways and thus variate analyses, BMI variation more than 5.71%
regulate the transcription of target genes that are is associated with higher rates of recurrences in
involved in tumor growth, invasion and meta- early-stage breast cancer patients (Fedele et al.
static potential (Gelaleti et al. 2012). Firstly, 2014). BMI and weight gain is even more strongly
estradiol antagonizes IL-6 function by repress- associated with fatal postmenopausal breast can-
ing both IL-6 and its signaling receptor, gp130 cer. In these cases, the percentage of postmeno-
(Pfeilschifter et al. 2002). Five-­year follow-up pausal breast cancer accounted for by weight
of 1199 women with hormone receptor positive gain alone is approximately 16% and by hormone
and human HER2-negative invasive breast can- replacement therapy alone was 5%. But when the
cer are evaluated for the impact of obesity on interaction between these variables is taken into
time to either local or distant recurrence or new account, their impact is about one third of post-
breast cancer, or death due to breast cancer. menopausal breast cancers (Huang et al. 1997).
Moderate to severe obesity is associated with a However, the effect on outcomes is not only lim-
poorer invasive breast cancer prognosis; this is ited to excess weight at presentation, as weight
also true for women with Stage 1 disease, and is gain after chemotherapy for breast cancer nega-
independent of age and treatment (Robinson tively impacts disease-free survival, local recur-
et al. 2014). Weight gain after diagnosis of rence, and death as well. Adjuvant chemotherapy
breast cancer is associated with higher all-­cause is associated with greater weight gain in node-­
mortality rates compared with maintaining body positive, postmenopausal breast cancer patients.
weight. Adverse effects are greater for weight The amount of weight gain appears greater for
gains of 10% or higher (Playdon et al. 2015). premenopausal than postmenopausal women. It
is clear that excessive weight gain in premeno-
25 Obesity-associated Breast Cancer: Analysis of risk factors 581

pausal women may be associated with an increase recruitment of macrophages and the formation
in the relapse risk of breast cancer and cancer-­ of crown-like structures. The recruited macro-
related deaths (Camoriano et al. 1990). Thus, phages are subsequently stimulated by CCL2 and
patients included 5204 Nurses’ Health Study par- IL-1beta to secrete CXCL12. In this case, induc-
ticipants diagnosed with incident, invasive, non-­ tion of angiogenesis supports the expansion of
metastatic breast cancer between 1976 and 2000; adipose tissue. Macrophages promote tumor pro-
860 total deaths, 533 breast cancer deaths, and gression from ductal carcinoma in-situ to aggres-
681 recurrences have been recorded during sive breast cancers. Finally, inflammation within
median follow-up of 9 years. Weight at diagnosis the obese mammary gland contributes to angio-
and weight gain is positively associated with genesis through a novel CCL2/IL-1beta/CXCL12
increased rates of breast cancer recurrence and pathway that bypasses the vascular endothelial
mortality (Kroenke et al. 2005). Furthermore, growth factor/vascular endothelial growth factor
during breast cancer chemotherapy, 31% of receptor (VEGF/VEGFR) pathway (Bowers
patients present a notable weight variation which et al. 2015). Both the systemic and local conse-
is greater than 5% of their initial weight. In mul- quences of chronic adipose inflammation provide
tivariate analyses, weight change more than 5% key potential mechanistic links between obesity
is positively associated with an increased risk of and breast cancer (Iyengar et al. 2013). Adipose
both recurrence and death (Thivat et al. 2010). tissue macrophages can comprise up to 40% of
the cells in adipose tissue of obese humans and
represent a rich source of cytokines which are
6  atty Acids and Breast
F key mediators of the increased risk of insulin
Cancer in Obesity resistance associated with obesity (Uysal et al.
1997; Weisberg et al. 2003). Furthermore, TLR4
Obesity is strongly associated with changes in constitutes a molecular link between nutrition,
the physiological function of adipose tissue, lead- lipids, and inflammation and that the innate
ing to insulin resistance, chronic inflammation, immune system participates in the regulation
and altered secretion of adipokines (van of energy balance and insulin resistance (Shi
Kruijsdijk et al. 2009). The increase in circula- et al. 2006). In this respect, saturated fatty acids
tory non-esterified fatty acids, adipose tissue dys- can activate TLR4 and its downstream signaling
function and the activation of toll-like receptors pathways involving myeloid differentiation pri-
(TLR) induce signal transduction pathways that mary response 88/IL-1 receptor-associated
lead to the activation of transcription factors. kinase/TNF receptor associated factor6 (MyD88/
Adipocyte-macrophage-TLR4 axis might be IRAK/TRAF6) and PI3K/Akt. Constitutively,
involved in the chronic low-grade inflammatory active Akt is sufficient to induce NFkappaB acti-
process occurring in obesity. Indeed, the obesity vation and COX-2 expression (Lee et al. 2003).
is associated with a marked infiltration of macro- In brief, obesity-inflammation axis is important
phages within the adipose tissue (Wolowczuk for the development and progression of breast
et al. 2008). Actually, saturated fatty acids plus cancer. Macrophages forming crown-like
TLR4 are mainly responsible for the amplifica- ­structures in the breast adipose tissue correlate
tion of this inflammatory process. In this vicious with both BMI and adipocyte size (Morris et al.
cycle, increased amount of saturated fatty acids 2011). Necrotic adipocytes surrounded by mac-
which are provided either by high-fat feeding or rophages form crown-like structures in both
adipocyte lipolysis could serve as ligands for mammary glands and visceral fat. Saturated fatty
TLR4. Finally, the activated adipocytes and mac- acids, which have been linked to obesity-related
rophages produce more pro-inflammatory prod- inflammation and crown-like structures are asso-
ucts (Wolowczuk et al. 2008) (Fig. 25.1). During ciated with activation of NF-kappaB and
obesity, the enhanced production of CCL2 and increased levels of tumor necrosis factor-alpha
IL-1beta by the breast adipose tissue increases (TNF-alpha), IL-1beta and COX-2. Each of these
582 A. Engin

cytokines contributes to the induction of aroma- lesser extent IL-1beta and IL-13 exhibit levels of
tase in preadipocytes (Subbaramaiah et al. 2011). expression that are inversely correlated to ER and
Seventy-­five per cent of obese and 70% of over- PR status. However, IL-1beta, IL-6, IL-8, IL-10,
weight breast cancer patients have crown-like IL-12, MCP-1 and MIP-1beta are more abundant
structures, while among normal-weight breast can- in high-grade tumors than in low-grade tumors.
cer patients only 8.3% have crown-like structures. In addition, IL-8 and MIP-1beta are expressed to
This corresponds to an approximately nine-fold a greater degree in HER2-positive patients
change in crown-like structures comparing compared to HER2-negative ones (Chavey et al.
obese patients to normal-weight women (Morris 2007). Obese adipose tissue-derived inflamma-
et al. 2011). tory mediator production could exacerbate
Actually, TNF-alpha is the major mediator of inflammation-associated tumorigenic effects.
cancer related inflammation. Besides increasing These mediators and adipokines such as leptin
total estrogen metabolites, TNF-alpha signifi- and inflammatory cytokines with a concomitant
cantly decreases the Estrogen1 (E1)/Estrogen2 reduction in adiponectin are produced in both
(E2) ratio due to a decrease in the level of E1 visceral adipose tissue depots and also within
with a concomitant increase in E2 concentra- mammary adipose tissue depots. All these depots
tion. TNF-alpha increases the expression levels contribute to the development of a more severe
of CYP1A1 and CYP1B1. Estradiol is metabo- breast cancer phenotype via stimulating breast
lized into 2-hydroxyestradiol (2-OHE2) and cancer growth, invasion and metastasis
4-hydroxyestradiol (4-OHE2) by CYP1A1 and (Dalamaga et al. 2012). n-6 polyunsaturated fatty
CYP1B1, respectively. These catechols undergo acid (n-6 PUFA)-derived eicosanoids are known
further oxidation into semiquinones and qui- as pro-inflammatory and pro-carcinogenic lipid
nones that react with DNA to form depurinating mediators. In contrast, n-3 PUFA-derived lipid
adducts leading to mutations, which are associ- mediators functionally oppose the synthesis and
ated with breast cancer (Kamel et al. 2012). activity n-6 PUFA-derived eicosanoids. Thereby,
Additionally, breast inflammation, as determined several of the known risk factors of these eico-
by crown-like structures of the breast, is paral- sanoids for breast cancer may be modified by
leled by increased NF-kappaB binding activity dietary omega-3 fatty acid supplementation
and elevated levels of aromatase mRNA and (Rose and Connolly 1999). Indeed, n-3 PUFA
aromatase activity (Subbaramaiah et al. 2012) have been shown to concurrently target in the
(Fig. 25.1). NF-kappaB activation underlies multiple aspects of the obese breast cancer phe-
many aspects of breast cancer cell proliferation, notype. In this respect, reduction of macrophage
invasion and metastasis (Rose and Vona-Davis adipose tissue infiltration, crown like structure
2014). In this context, seventeen cytokines; formation and down-regulation of critical adipo-
IL-1beta, IL-2, IL-4, IL-5, IL-6, IL-7, IL-8, kine production are the major goals. Collectively,
IL-10, IL-12 [p70], IL-13, IL-17, granulocyte increased n-3 PUFA intake could attenuate
colony-stimulating factor (G-CSF), granulocyte-­ obesity-­associated breast cancer (Monk et al.
macrophage (GM-CSF), interferon-gamma 2014). Eventually, breast cancer risk has been
(IFN-gamma), monocyte chemoattractant protein found to be negatively associated with specific
(MCP)-1, macrophage inflammatory protein n-3 PUFAs (eicosapentaenoic acid and docosa-
(MIP)-1beta and TNF-alpha are expressed in hexaenoic acid) and positively associated with
breast carcinoma, whereas only nine of these n-6 PUFAs (linoleic acid and arachidonic acid) in
could be detected in normal breast. Most cyto- breast adipose tissues, which are collected from
kines are more abundant in breast carcinoma than 73 breast cancer patients and 74 healthy subjects
in normal breast, with IL-6, IL-8, G-CSF, IFN-­ (Bagga et al. 2002). Excessive dietary intake of
gamma, MCP-1 and MIP-1beta being very abun- n-6 PUFA versus n-3 PUFA results in a signifi-
dant. IL-2, IL-6, IL-8, IL-10, IFN-gamma, cantly greater proportion of eicosanoids genera-
MCP-1, MIP-1beta and TNF-alpha, and to a tion from n-6 PUFA (Calder 2012). Even
25 Obesity-associated Breast Cancer: Analysis of risk factors 583

short-term dietary intervention can lead to statis- Marquez et al. 2013). At first, oleic acid medi-
tically significant increases in omega-3/omega-6 ates the production of arachidonic acid (AA),
polyunsaturated fatty acid ratios in plasma and and then AA metabolites mediate focal adhesion
breast adipose tissue of women with high-risk kinase phosphorylation and cell migration in
localized breast cancer (Bagga et al. 1997). breast cancer cells. In contrast, Gi/Go proteins,
Actually, inflammatory process leads to the phospholipase C, LOXs and Src inhibitor pre-
induction of COX-2 expression and a consequent vents focal adhesion kinase phosphorylation and
elevation in prostaglandin (PG) production. cell migration (Navarro-Tito et al. 2010). Insulin,
Together with other pro-inflammatory cytokines, leptin, and adiponectin has no effect on oleic
the eicosanoids promote further development and acid, AA, and eicosapentaenoic acid uptake by
growth of breast cancers by the induction of breast cancer cells. However, preferential uptake
aromatase, particularly in estrogen positive of AA in breast cancer cells, and the fatty acid
breast cancers. Meanwhile, the more aggressive, uptake activity of these cells is influenced by
estrogen-­ independent tumors may develop by TNF-alpha (Kaur et al. 2009). In breast cancer,
direct stimulatory effect of PGE2 and lipoxygen- the free oleic acid induces tumor invasion
ase (LOX) products (Vona-Davis and Rose 2013). through an epidermal growth factor receptor
The high energy consumption is indispensable (EGFR), guanosine triphosphate (GTP)-binding
for both primary tumor growth and secondary proteins; Gi/Go proteins, matrix metalloprotein-
tumor cell metastasis (Phoenix et al. 2010). ases, PKC and Src-dependent pathway, but it is
Thereby, most breast cancer cell types are not able to promote invasion in non-invasive
addicted to fatty acids, which they require for breast cancer cells (Soto-Guzman et al. 2010).
membrane phospholipid synthesis, signaling pro- The phospholipase C, mitogenic-extracellular
cesses, and energy production. Expression of the signal-regulated kinase 1/2 (MEK 1/2), Src, and
enzymes required for fatty acid synthesis is PI3K/Akt signaling pathways are involved in the
closely linked to each of the major classes of sig- proliferative signal induced by oleate. This effect
naling molecules that stimulate breast cancer cell is mediated at least in part via the G protein-cou-
proliferation (Kinlaw et al. 2016). Obesity, which pled receptor (GPR) 40. Eventually, fatty acids
is characterized by hyperlipidemia and an eleva- control breast cancer cell growth via GPR40,
tion of circulating free fatty acids, is also associ- and may provide a link between fat and cancer
ated with enhanced cancer risk (Soto-Guzman (Hardy et al. 2005). Consequently, G-protein-
et al. 2010). Thus, diet-induced obesity has been coupled cell surface receptor for long-chain fatty
shown to promote the incidence of breast cancer acids is involved in human breast cancer cell
development (Cleary et al. 2004a). Additionally, proliferation (Yonezawa et al. 2004). Saturated
fatty acid synthase is strongly associated with the fatty acids released from adipocytes have been
postmenopausal breast cancer. BMI and patho- linked to obesity-related white adipose tissue
logical stage of tumor is also related to fatty acid inflammation. In this respect, saturated fatty acid
synthase expression of breast cancer cells (Porta stimulates Akt-dependent activation of
et al. 2014). By analyzing mastectomy specimen, NF-kappaB. Later on, TNF-alpha, IL-1beta and
the amounts of glandular and fat tissue are quan- COX-2 levels are increased in ­ macrophages.
tified in the resected breasts tissue. Adipose tis- Release of these cytokines strongly provoke aro-
sue is a major component, comprising up to 56% matase activity (Subbaramaiah et al. 2013).
of the total breast volume (Vandeweyer and Actually, lipid desaturation is an essential pro-
Hertens 2002). cess for cancer cell survival. Stearoyl-CoA desat-
As mentioned above, linoleic acid is a dietary urase mRNA and protein expression is elevated
n-6 PUFA that is known to induce proliferation in human breast cancers and predicts poor sur-
and invasion in breast cancer cells. It promotes vival. Therefore, inhibition of stearoyl-CoA
focal adhesion kinase and Src activation, as well desaturase activity could efficiently control syn-
as cell migration in breast cancer cells (Serna-­ thesis of unsaturated fatty acids and limit breast
584 A. Engin

tumor growth (Peck et al. 2016). Lipogenesis is through phosphorylation on a single tyrosine.
regulated by the enzyme fatty acid synthase Activated STATs form dimers, translocate to the
(FASN); and breakdown of fatty acids is regu- nucleus, bind to specific response elements in
lated by carnitine palmitoyltransferase-1 (CPT-I). promotors of target genes, and transcriptionally
FASN is highly expressed in breast cancers (Puig activate these genes (Heim 1999). After the acti-
et al. 2008). Basal expression of rate-limiting vation of JAK 1 tyrosine kinase and STAT1 and
enzymes FASN, stearoyl-CoA desaturase and STAT3 transcription factors by IL-6, expression
lipolytic phospholipase A2 group IVB, as well as of dominant negative STAT3 in the cells strongly
lipogenesis transcription factors peroxisome pro- reduces IL-6-mediated growth inhibition but
liferator activated receptors alpha (PPARalpha), does not prevent IL-6-induced cell migration.
sterol regulatory element binding factor 2 IL-6 treatment leads to activation of the MAPK
(SREBF2) and cAMP responsive element modu- and the PI3K pathways. Inhibition of MAPK or
lator (CREM) are higher in breast cancer cells. PI3K activity reverses IL-6- stimulated migration
Over-expression of lectin-like oxidized-low den- (Badache and Hynes 2001). The cancer cell death
sity lipoprotein (OX-LDL) receptor-1 in breast signaling c-Jun N-terminal kinase (JNK) path-
cancer cells also enhances cell migration, without way is inhibited by IL-6. Consequently, IL-6
affecting their adherence to TNF-alpha-activated effectively protects cancer cells from the apopto-
endothelium or transendothelial migration. sis and allows for the proliferation of malignant
Additionally, lectin-like OX-LDL receptor-1 cells (Lin et al. 2001). Additionally, IL-6 and
inhibits apoptosis and stimulates cancer cell pro- TNF-alpha acts as a regulator of estrogen synthe-
liferation through NF-kappaB signaling pathway sis and aromatase activity in normal and malig-
(Khaidakov et al. 2011). Triple negative breast nant breast tissues (Purohit et al. 1996). Thereby,
cancer cell lines show overexpression of FASN IL-6 and IL-1 beta regulates proliferation of
(Giró-Perafita et al. 2016). Through a PI3K-­ breast cancer cells through estrogen production
dependent pathway, HER2 stimulates the FASN by steroid-catalyzing enzymes in the breast tissue
promoter and ultimately mediates increased fatty (Honma et al. 2002) In this respect, serum IL-6
acid synthesis. Interestingly, pharmacological concentration is significantly higher in patients
inhibition of FASN preferentially induces apop- with breast cancer compared to healthy controls.
tosis of HER2-overexpressing breast epithelial Thus, median IL-6 serum levels are nearly ten
cells relative to matched control cells (Kumar-­ times higher in patients with metastatic breast
Sinha et al. 2003). By contrast, blocking the cancer compared to those with loco-regional dis-
FASN leads to apoptosis of HER2-positive breast ease. Therefore, serum IL-6 level is the most dis-
carcinoma cells (Blancafort et al. 2015). criminative factor separating healthy controls and
A high-fat diet with estrogen deprivation leads the loco-regional and metastatic breast cancer
to development of insulin resistance, which may patient groups (Benoy et al. 2002). Moreover, the
accelerate mammary tumor growth through the patients with high IL-6 levels show significantly
insulin receptor-mediated Akt pathway and inac- poorer survival than patients with low IL-6 levels
tivation of AMP-activated protein kinase (Zhang and Adachi 1999). Indeed, higher
(AMPK) in vivo. High circulating insulin in com- ­production of IL-6 is associated with worse out-
bination with increased expression of insulin comes in breast cancer patients at high risk of
receptor in tumor tissues may result in stimula- relapse, however these effects are limited to those
tion of Akt/mTOR signaling leading to the accel- patients with ER-positive tumors. IL-6 may exert
eration of breast tumor growth (Kim et al. 2015). its effect on breast cancer cells through hormonal
Janus Kinases (JAKs) are a unique class of tyro- pathways (DeMichele et al. 2009). Although
sine kinases that associate with cytokine recep- both IL-6 and TNF-alpha are expressed by
tors. Upon ligand binding, IL-6, IL-4 and G-CSF adipose tissue, there are important differences in
activate members of the Signal Transducers and their systemic release. In contrast to TNF-alpha,
Activators of Transcription (STAT) family IL-6 is released from the subcutaneous adi-
25 Obesity-associated Breast Cancer: Analysis of risk factors 585

pose tissue depot and thereby able to signal cascade of secondary pro-tumorigenic cytokines
systemically (Mohamed-Ali et al. 1997). (Pantschenko et al. 2003) (Fig. 25.1).
Consequently, there is a positive relationship Adipocyte fatty acid binding protein 4
between IL-6 concentration and percentage of (FABP4) is a key adipokine for fatty acid trans-
body fat (Vozarova et al. 2001). IL-6 genotypes port. Exogenous FABP4 enhances the proliferation
may influence breast cancer risk in conjunction of breast cancer cells and induces the Akt and
with central adiposity in postmenopausal women MAPK signaling cascades. The inhibition of
(Slattery et al. 2008). Interestingly, the breast these pathways reduces the exogenous FBAP4-­
tumors of larger size and/or with lymph nodes mediated cell proliferation (Guaita-Esteruelas
involvement exhibit higher levels of IL-6 in et al. 2016). Although serum adipocyte-FABP
tumor surrounding adipocytes. Adipocytes par- (A-FABP) levels are found to be significantly
ticipate in a vicious cycle, which is controlled by higher in obese than in non-obese women, inde-
cancer cells. Thus, invasive cancer cells dramati- pendent of obesity, the serum A-FABP levels are
cally affect surrounding cancer-associated adipo- significantly higher in breast cancer patients than
cytes. Furthermore, cancer-associated adipocytes in healthy controls. Furthermore, A-FABP is pos-
modify the cancer cell phenotypes leading to a itively correlated with tumor size and nodal-­
more aggressive behavior (Dirat et al. 2011). status (Hancke et al. 2010). In fact, FABP4,
Deleterious function of cancer-associated adipo- adiponectin and retinol binding protein 4 (RBP4)
cytes is dependent on their crosstalk with inva- are most down regulated genes in breast cancer.
sive cancer cells. Indeed, this event leads These genes also display strong connections with
to dramatic phenotypic and/or functional modifi- the other molecules of lipid metabolism pathway
cations of both cell types. These adipocytes in breast cancer (Merdad et al. 2015).
exhibit delipidation and acquire a fibroblast-like
shape. A high number of cancer-associated adi-
pocytes might be predicted to be detrimental in 7  diponectin and Breast
A
obesity (Tan et al. 2011). Therefore, IL-6 seems Cancer
to play a key role in the acquired proinvasive fea-
ture of breast cancer cells. Furthermore, IL-6 Adiponectin is synthesized and secreted almost
induces cell migration through the activation of exclusively by the white adipose tissue. AdipoR1
the MAPK pathway, acts as an anti-apoptotic fac- and AdipoR2 serve as receptors for globular and
tor, promotes the osteoclasts formation, and full-length adiponectin. These receptors mediate
inhibits the differentiation of dendritic cells. In the increased AMPK, PPAR-alpha ligand activi-
this wise, IL-6 facilitates the metastases of breast ties, and glucose uptake and fatty-acid oxidation
cancer (Macciò and Madeddu 2011). by adiponectin. Obesity decreases expression lev-
Furthermore, tumor-associated IL-1alpha and els of AdipoR1/R2. Hence, adiponectin sensitivity
IL-1beta are present in the tumor microenviron- is reduced in obesity (Kadowaki and Yamauchi
ment and may play a pivotal role in regulating 2005). Additionally, reduced expressions of
invasive cancer and ductal carcinoma in-situ AdipoR1/R2 in obese animals are significantly
growth and metastasis (Kurtzman et al. 1999). correlated with decreased binding of adiponectin
The presence and distribution of IL-1 cytokines to membrane fractions. This s­ ituation is defined as
and its receptors (IL-1R) in human breast cancer adiponectin resistance. Increase in adiponectin
suggests that the local expression of IL-1 results resistance in turn may play a role in worsening
in the activation of a population of cells within insulin resistance in obesity (Tsuchida et al. 2004).
the human breast cancer microenvironment. This In women with low-­circulating adiponectin levels,
activation of the IL-1/IL-1R cytokine family via breast tumors may present a more aggressive phe-
autocrine and/or paracrine mechanisms leads to a notype, by large tumor size, high-histological
grade, estrogen receptor negativity, and increased
586 A. Engin

angiogenesis and metastasis (Mantzoros et al. through AdipoR1 (Nakayama et al. 2008). In con-
2004). Adiponectin may act on tumor cells directly trast, adiponectin also triggers cellular apoptosis
by binding and activating adiponectin receptors in breast cancer cells in the presence of 17-beta
and downstream signaling pathways (Barb et al. estradiol. A cross-talk between adiponectin and
2006). In fact, breast cancer cells can express both estrogen receptor signaling exists in breast cancer
AdipoR1/R2 but not adiponectin. In contrast, cells (Pfeiler et al. 2008). Higher circulating lev-
women with the highest adiponectin levels have a els of leptin found in obese subjects could be a
65% reduced risk of breast cancer (Körner et al. growth-enhancing factor, whereas low adiponec-
2007). The analysis of 885 cases of breast cancer tin levels in obese women may be permissive for
revealed that adiponectin levels are not related to growth-promoting effect of leptin. In any case,
the risk of breast cancer in premenopausal women, estrogen and its receptors have a definite impact
however, lower adiponectin levels are associated on the response of human breast cancer cell lines
with a higher risk of breast cancer in postmeno- to leptin and adiponectin (Grossmann et al. 2010).
pausal women (Ye et al. 2014). Thus, an inverse Obesity leads to altered expression profiles of
association is found between adiponectin and various adipokines and cytokines including leptin,
postmenopausal breast cancer risk in 1477 inci- adiponectin, IL-6, TNF-alpha and IL-1beta. The
dent breast cancer cases. However, a modest cor- increased levels of leptin and decreased adiponec-
relation is estimated with ductal type of breast tin secretion are directly associated with breast
cancer, but not lobular tumors. Adiponectin may cancer development (Khan et al. 2013). The
only influence breast cancer cell proliferation in a leptin/adiponectin ratio provides significant
low estrogen environment (Tworoger et al. 2007). adjunctive information to the risk of metabolic
Nevertheless, the frequency of lymph node metas- syndrome beyond homeostatic model assessment-
tasis of tumor is significantly increased in the insulin resistance (HOMA-IR)) alone (Yoon et al.
patients with low plasma adiponectin levels. 2011). In patients with breast cancer, extracellular
Furthermore, the frequency of tumors with nega- leptin is higher and adiponectin is lower in tumors
tive estrogen receptor is significantly increased in than in normal adjacent breast tissue. While estro-
the patients who have less than the median adipo- gen exposure increases leptin secretion in post-
nectin level (Kang et al. 2007) (Fig. 25.1). menopausal group, tamoxifen treatment increases
The highest high molecular-weight adiponec- adiponectin and decreases leptin and the leptin/
tin levels and lower BMI have shown a signifi- adiponectin ratio (Morad et al. 2014). A signifi-
cantly reduced risk of breast cancer in both pre cant reduction in tumor volume is observed in
and postmenopausal women in a case-control human ER-alpha-­ negative breast cancer cells,
study, including 66 sets of Japanese female breast which are pretreated with adiponectin, whereas an
cancer cases and age and menopausal status increased tumor growth is evident in the human
matched controls (Minatoya et al. 2014). Low ER-alpha-positive breast cancer cells. Cyclin D1
total or high molecular-weight adiponectin asso- mRNA and protein levels are also up-regulated in
ciates with risk of breast cancer particularly ER-alpha-positive cells by adiponectin (Mauro
among premenopausal and obese women. In con- et al. 2015). Actually, exogenous adiponectin may
trast to cancer tissue, adiponectin expression is induce the inhibition of cell proliferation and pro-
elevated in adjacent non-tumor adipose tissues of motion of apoptosis in breast cancer cells (Chen
breast, compared with controls. In both tumor tis- and Wang 2011). Conversely, breast cancer aris-
sue samples and breast cancer cell lines, AdipoR1 ing in women with the low serum adiponectin lev-
expression is two to four times higher than that of els are more likely to show a biologically
AdipoR2 (Körner et al. 2007). Adiponectin aggressive phenotype (Miyoshi et al. 2003).
decreases breast cancer cell proliferation by Since, adiponectin acts directly on breast cancer
inhibiting the entry into S-phase without inducing cells by inhibiting proliferation and angiogenesis
apoptosis, and this inhibitory effect is mediated or by stimulating apoptosis, it is proposed that
25 Obesity-associated Breast Cancer: Analysis of risk factors 587

replacement of adiponectin may be of major els (Pollak 1998). Circulating insulin levels and
importance in the prevention and the treatment of plasma levels of IGFBP-3 are simultaneously
breast cancer in obese patients (Delort et al. 2012). elevated in women with premenopausal breast
Consequently, serum adiponectin levels in cancer (Del Giudice et al. 1998). Approximately
ER-negative/PR-negative breast cancer show an 1% of circulating IGF-1 remains unbound, while
inverse relationship with the risk of recurrence the rest is mainly bound to IGFBP-3 (Pollak
(Oh et al. 2011). 2008). IGFBP-­binding to cell surface results in
the release of more IGF to the receptors
(McCusker et al. 1990). Secretion of IGFBP-1
8 Insulin Resistance and IGFBP-2 is further suppressed by insulin
and Insulin-Like Growth and is diminished with increasing obesity
Factor in Breast Cancer (Hoeflich and Russo 2015). The insulin-cancer
hypothesis postulates that chronic hyperinsu-
The association of insulin with breast cancer linemia is associated with decreased concentra-
outcomes is nonlinear. Insulin resistance is sig- tions of IGFBP-1 and IGFBP-2, leads to
nificantly correlated with obesity, which is asso- increased availability of IGF-1 and concomitant
ciated with distant recurrence and death changes in the cellular environment that favor
(Goodwin et al. 2002). This connection between tumor formation (Renehan et al. 2006). The
insulin resistance and obesity is strongly valid IGF-1 and IGFBP-2 are amongst the most potent
for only postmenopausal patients with high pro- mitogens for mammary epithelial cells and there
liferative luminal B/HER-2-negative type of are accumulating evidences that they interact
breast cancer (Nam et al. 2016). Evaluation of 22 with the estradiol-axis to regulate mitogenesis,
studies with 7478 cases indicated that fasting apoptosis, adhesion, migration and differentia-
insulin and non-­fasting or fasting C-peptide lev- tion of mammary epithelial cells. Such interac-
els are not different between women with and tions are bi-directional and estradiol has been
without breast cancer, whereas HOMA-IR levels shown to regulate the expression and activity of
in breast cancer patients are significantly higher IGF-axis genes with the general effect of sensi-
than in women without breast cancer (Hernandez tizing breast epithelial cells to the actions of
et al. 2014). In this respect, hyperinsulinemia is IGFs and insulin (Hawsawi et al. 2013). IGF-1
a consequence of insulin resistance or the can specifically target its receptors on human
impaired responsiveness of cells to insulin. This breast epithelial cells to induce mitogenic and
metabolic imbalance is more common in obese anti-apoptotic pathways. Both the PI3K and
women than in normal-weight women (Lazarus MAPK pathways are important for IGF-1-­
et al. 1998). Evaluation of 816 randomly chosen stimulated proliferation of human breast cancer
postmenopausal breast cancer cases revealed cells in vitro. Insulin-receptor substrate (IRS)-1,
2.4-fold increase in breast cancer incidence in but not IRS-2, is the predominant signaling mol-
women with the highest quartile of fasting insu- ecule activated by IGF-I and insulin (Jackson
lin concentrations. Ultimately, hyperinsulinemia et al. 1998). For overall survival of breast cancer
and insulin or insulin like growth factor-1 (IGF- patients, only histological grade and IGF-1R
1)-signaling are independent risk factors for mRNA emerges as significant predictors.
breast cancer and may have a substantial role in Increased IGF-1R mRNA implies poorer prog-
explaining the obesity-breast cancer relationship nosis among the different subtypes of breast can-
(Gunter et al. 2009). Actually, IGFs are mito- cer, and that may be associated with the lack of
genic and anti-apoptotic peptides that influence responsiveness to tamoxifen in cases with a posi-
the proliferative behavior of many cell types, tive hormone receptor status (Peiró et al. 2011).
including normal and transformed breast epithe- In contrast, IGF-1R correlates with good prog-
lial cells. IGF binding proteins (IGFBPs), and nostic markers among patients with early breast
IGFBP proteases regulate circulating IGF-I lev- cancer and is differentially expressed with vari-
588 A. Engin

able prognostic impact among breast cancer sub- pivotal factor in adipocyte regulation of breast
types (Yerushalmi et al. 2012). However, IGF-2 cancer cell growth and it is upregulated at early
is expressed both in the epithelial tumor cells period of adipocyte differentiation (D’Esposito
and in stromal cells, in 84% and 50% of breast et al. 2012). Indeed, obesity is associated with
cancer cases, respectively. IGF-2 expression in significant changes in the growth hormone (GH)/
cancer cells is related to a non-metastatic disease IGF system. In non-­diabetic obese subjects, adi-
at diagnosis or to low tumor grade (Toropainen pocytes produce large amount of free IGF-1 and
et al. 1995). IGF-1R may be overexpressed in all IGF-2, total IGF-2, IGFBP-3 and growth hormone-
breast cancer subtypes, regardless of the hor- binding protein (GHBP), reduced IGFBP-1 and
mone receptor or HER2 status and its expression IGFBP-2 when compared to lean subjects.
rates range from 43.8 to 87% by Allred score Conversely, in obese Type 2 diabetics, free IGF-I
(Shimizu et al. 2004; Yerushalmi et al. 2012). is insignificantly reduced, when compared to
There is a significant correlation between non-diabetic obese subjects (Frystyk et al. 1999).
IGF-1R and ER status, but not between IGF-1R Adipocytes-­derived IGF-1 release is regulated
and PR status. Patients with IGF-1R-positive by glucose and fatty acids and may contribute to
and ER-negative breast cancers are in a worse the control of breast cancer cell growth in obese
situation than IGFR-negative ER-negative can- individuals (D’Esposito et al. 2012). Actually,
cer patients (Railo et al. 1994). Within 4 years of IGF-1 and IGF-2 stimulate both normal growth
initial breast tumor diagnosis, elevated levels of development and breast cancer cell proliferation.
IGF-1R are strongly associated with ipsilateral IGF-2 induction of the aryl hydrocarbon recep-
breast tumor recurrence. IGF-1R predisposes to tor (AHR) promotes the expression of Cyclin D1
early relapses after radiation therapy, which is and the proliferation of breast cancer cells
thought to be true for the recurrences. After (Tomblin and Salisbury 2014). The AHR also
more than 4 years, recurring tumors are more regulates cell cycle in part by binding with
likely to represent de novo primary breast can- Cyclin D1 and cyclin dependent kinase 4
cers, therefore IGF-1R expression would not be (CDK4) in human breast cancer cells (Barhoover
expected as a marker of radioresistance (Turner et al. 2010). Adipocytes secrete IGF-2 at levels
et al. 1997). On the other hand, reduction of that stimulate the proliferation of human
serum IGF-1 levels results in significantly ER-positive breast cancer cells. By contrast, spe-
decreased burden of tumors, despite modestly cific exogenous AHR ligands inhibit the prolif-
elevated levels of circulating insulin and leptin erative effects of mitogenic adipokines, including
(Lashinger et al. 2011). In this context, IGFBP-3 IGF-2 in human ER expressing breast cancer
induces apoptosis in an IGF-IGFR axis-indepen- cells (Salisbury et al. 2013). In the absence of an
dent manner through the activation of caspases exogenous AHR agonist, the AHR is located in
involved in a death receptor-­mediated pathway the cytoplasm bound to chaperon proteins. In the
in human breast cancer cells. Consequently, classical mechanism of AHR action, upon acti-
IGFBP-3 functions as a negative regulator of vation by an agonist, the AHR dissociates from
breast cancer cell growth (Kim et al. 2004) chaperon proteins, translocates into the nucleus,
(Fig. 25.1). The characteristics and phenotypic and stimulates transcription by binding to AHR
behavior of malignant breast ductal epithelial response elements (AHRE) (Denison et al.
cells associate with the synergistic activity of 2011). AHR-­dependent gene transcription is ter-
adipocyte-derived factors. Adipocyte-secreted minated by dissociation of the liganded AHR-
substances can affect tumorigenesis by increas- AHR nuclear translocator (ARNT) complex
ing the stabilization of pro-­ oncogenic factors from the dioxin-­responsive element (DRE), sub-
(Iyengar et al. 2003). Adipocytes from subcuta- sequent nuclear export of the AHR into the cyto-
neous adipose tissue specimens of obese indi- sol is mediated by its N-terminal nuclear export
viduals are capable of producing larger amounts sequence followed by ubiquitin-mediated AHR
of IGF compared with lean women. IGF-1 is a degradation (Pollenz 2002). Although the ER
25 Obesity-associated Breast Cancer: Analysis of risk factors 589

signaling pathway has many mechanistic simi- cancer and is associated with low levels of adipo-
larities to that of the AHR, there are some differ- nectin and shorter breast cancer survival (Duggan
ences. ER isoforms are localized primarily in the et al. 2011). Actually, adiponectin and HOMA-IR
nucleus in their unliganded state and are found have also prognostic significance in breast cancer
complexed with two chaperone proteins recurrence. Treatments related to these factors
(Heldring et al. 2007). However, AHR-ER cross- may protect against recurrence in ER-negative/
talk is multifactorial in nature, effecting both PR-negative patients. Similar results could not be
ERalpha and ERbeta, and involving a combina- achieved in the case of ER-positive/PR-positive
tion of both classical and non-classical AHR- patients (Oh et al. 2011). Interestingly, an inverse,
dependent mechanisms that results in inhibition significant association between insulin resistance
of estrogen response or responsiveness. and cancer recurrence is observed in the
Eventually, induction of gene expression by the ER-positive/PR-positive breast cancer patients.
classical AHR-ARNT-DRE signaling pathway Furthermore, ER-positive/PR-positive patients
can repress ER-dependent signaling (Denison with hyperglycemia showed decreased risk of
et al. 2011). recurrence (Oh et al. 2011).
Considering breast cancers overall, the rela-
tionship between obesity and cancer risk is com-
plex. Obesity is associated with an increased risk 9 Adipose Tissue Hypoxia
of postmenopausal breast cancer. Contrarily, in in Breast Cancer
premenopausal women, obesity has no or even
reducing effect on breast cancer risk. However, Increased adipose tissue hypoxia, accompanies
epidemiologic data have demonstrated that obe- obesity. Obesity-induced hypoxia and oxidative
sity is strongly associated with an increased risk stress affect the production of many adipocyte-­
of triple-negative breast cancer in both pre- and derived proteins involved in angiogenesis,
postmenopausal women (Millikan et al. 2008). inflammation and extracellular matrix remodel-
Actually, triple-negative breast cancers are ing. All these alterations can establish a pro-­
aggressive tumors, which are typically malignant environment for breast tissues
ER-negative, PR-negative and human EGFR-2 (Yao-Borengasser et al. 2015). Cancer-associated
negative tumors and targeted therapies for triple-­ adipocytes (CAAs) are essential for breast tumor
negative breast cancers are almost currently development and progression. Furthermore,
unavailable (Toft and Cryns 2011). CAAs modify the cancer cell characteristics and
Insulin may influence prognosis to the great- phenotype as leading to a more aggressive tumor
est extent during the first 5 years after diagnosis, behavior (Dirat et al. 2011). Adipocytes can
whereas obesity-related factors, particularly modify ER gene expression and promote
leptin continue to be important with longer fol- epithelial-­mesenchymal transition in breast can-
low-­up (Goodwin et al. 2012). IGF-I activates cer cells through upregulation of hypoxia-­
downstream signaling molecules within the inducible factor 1 alpha (HIF-1alpha). Obesity
leptin receptor and IGF-1R pathways. In contrast plays an important role in the development of an
to IGFI, leptin does not induce phosphorylation aggressive breast cancer (Yao-Borengasser et al.
of IGF-1R, indicating that receptor cross-­ 2015). Decreased oxygen availability stimulates
signaling is unidirectional and signal transmis- cells to consume glucose and produce lactate.
sion occurs from IGF-1R to leptin receptor in This response is coordinated by the HIF-1,
human breast cancer cells (Ozbay and Nahta which is expressed under the control of signaling
2008). Elevated HOMA-IR scores and low levels of the PI3K/Akt/mTOR pathway in tumor cells
of adiponectin are both associated with obesity (DeBerardinis et al. 2008). A significant correla-
and increased breast cancer mortality in obese tion is demonstrated between impaired glucose
breast cancer cases. Hyperinsulinemia is an inde- tolerance and disease progression in postmeno-
pendent risk factor for poor prognosis for breast pausal women receiving treatment of breast can-
590 A. Engin

cer. In particular lactate, are predictive of reduced


derived leptin is strongly involved in mammary
response to chemotherapy and is prognostic for carcinogenesis by contributing to the local pro-­
weight gain during early breast cancer chemo- inflammatory mechanisms. Therefore, obese
therapy (Stebbing et al. 2012). Hyperinsulinemia patients with breast cancer have increased meta-
could induce breast cancer progression through static potential and greater risk of mortality
leptin-dependent mechanisms. Thus, insulin (Delort et al. 2015). Serum leptin levels and
stimulated leptin expression is associated with leptin/BMI ratio are significantly increased in
increased activation of the leptin gene promoter. patients with breast cancer (Romero-Figueroa
Excess insulin increases nuclear accumulation et al. 2013). In the primary breast cancer cases,
of transcription factors HIF-1alpha and their expressions of leptin and leptin receptor are
loading on the leptin promoter (Bartella et al. found in 85% and 75%, respectively. In these
2008). Furthermore, serum lactate inhibits lipol- cases, the expression of leptin is significantly
ysis and triglyceride breakdown in adipocytes correlated with leptin receptor. Additionally,
through a direct activation of the orphan leptin receptor expression in primary breast can-
G-protein-coupled receptor, GPR81(Liu et al. cer is positively correlated with estrogen receptor
2009). Breast cancer cells can switch to a form expression and tumor size but not with Ki67 or
that can utilize lactate as a primary source of progesterone receptor expressions (Jardé et al.
energy, allowing them to survive in case of glu- 2008). Both normal epithelial cells and carci-
cose deprivation and this activity confers resis- noma cells express leptin. However, overexpres-
tance to PI3K/mTOR inhibitors. Estrogen-related sion of leptin is observed in 92% of invasive
receptor alpha (ERR-­alpha), is shown to regulate ductal carcinomas but in no normal epithelium.
the expression of genes required for lactate utili-
Furthermore, normal mammary epithelial cells
zation (Park et al. 2016). COX-2 directly regu- do not express a significant level of leptin recep-
lates gene expression of specific aromatase tor, whereas 83% of breast cancer cases have
promoter regions and regulates aromatase leptin receptors. Accordingly, distant metastasis
enzyme activity in breast cancer cells (Prosperi is detected in 34% of all leptin receptor-positive
and Robertson 2006). PGE2 also increases HIF- tumors with leptin overexpression, but none of
1alpha transcript and protein expression, nuclear the patients with leptin receptor-negative and
localization and binding to aromatase promoter- weak leptin expressing tumors are associated
II in human breast adipose stromal cells. with distant metastasis (Ishikawa et al. 2004).
Actually, HIF-1alpha as a modulator of aroma- Higher circulating levels of leptin found in obese
tase promoter II-driven aromatase expression in individuals and act as growth-enhancing factor.
human breast tumor-associated stroma and pro- In contrast, low adiponectin levels in obese
vides a special mechanism for estrogen regula- women may be permissive for leptin’s growth-­
tion in obesity-related, post-­menopausal breast promoting effects (Grossmann et al. 2010)
cancer (Samarajeewa et al. 2013). (Fig. 25.1). Leptin/adiponectin ratio are increased
significantly in the breast cancer patients, in com-
parison to controls. BMI is negatively and posi-
10 Leptin and Breast Cancer tively correlated to serum adiponectin and leptin
levels, respectively. Low serum adiponectin lev-
Adipocytes account for more than 90% of human els and high serum leptin levels are associated
breast volume and secrete adipocytokines, with an increased risk for breast cancer. The
including leptin. Breast adipose tissue-derived blood levels of estradiol increase in parallel to
leptin promotes the growth of developing breast those of leptin. The increased serum ratio of
cancer (Guo et al. 2012; Khandekar et al. 2011). leptin/adiponectin may indicate the presence of
Indeed, the breast cancer cells are surrounded by aggressive breast cancers (Chen et al. 2006).
an adipocyte microenvironment, which is more Furthermore, high leptin levels in obese and
extensive in obese people. Thus, adipocyte-­ overweight individuals are also positively corre-
25 Obesity-associated Breast Cancer: Analysis of risk factors 591

lated with body fat and adipocyte size. In terms SH2B1 directly binds to both IRS-1 and IRS-2
of obesity, leptin is unable to regulate appetite and mediates the formation of a JAK2/SH2B/
and size of fat deposits leading to a “leptin resis- IRS-1 or IRS-2 protein complexes. Consequently,
tance status” (Guo et al. 2012). SH2B dramatically enhances leptin-stimulated
Almost all of the breast cancer samples co-­ tyrosine phosphorylation of IRS-1 and IRS-2 and
express leptin and its two main isoforms of recep- subsequent activation of the PI3K pathway (Duan
tors, leptin receptor-L (long form) OBR-L and et al. 2004). mRNA expression of leptin and
OBR-S (short form) at the mRNA and protein leptin receptors in adipose tissue and matched
levels. OBR-L or OBR-S is found to be positively tumor samples, respectively, are associated with
correlated with the lack of progesterone receptor. obesity status in breast cancer. Increasing insulin
Elevated OBR-S expression has longer relapse-­ resistance is a predominant feature of this higher
free survival, whereas high OBR-L/OBR-S is leptin and leptin receptors expression (Carroll
associated with a shorter relapse-free survival in et al. 2011). In response to leptin, STAT3 binds to
breast cancer patients. This inverse ratio between phospho-Tyr1138, allowing JAK2 to phosphory-
OBR-L and OBR-S could be a marker of tumor late and activate STAT3. The JAK2-dependent
aggressiveness (Révillion et al. 2006). and -independent pathways act coordinately and
Actually, circulating leptin levels are propor- synergistically to promote STAT3 activation
tional to the body fat mass, thus serve as an adi- (Morris and Rui 2009). Indeed, chronic leptin
posity signal of the total body energy stores. signaling causes significant phosphorylation of
However, despite increased leptin levels, animals JAK2 and STAT3 together with the higher breast
fed a high-fat diet without decreasing their caloric cancer cell proliferation rate. In addition to the
intake, make them become obese. This suggests high ERalpha/ERbeta ratio, estrogen-dependent
that a high content of dietary fat changes the “set transcriptional activity, E2-dependent cell growth
point” for body weight, at least in part by limiting and resistance to antiestrogen agents is enhanced
the action of leptin (Frederich et al. 1995). (Valle et al. 2011). Activated STAT3 and G9a his-
Therefore, regulating leptin sensitivity is tone methyltransferase in epigenetic silencing of
extremely important to control body weight. In miR-200c promotes the formation of breast can-
this respect, leptin exerts its biological action cer stem-like cells. These cells elevate the cell
through binding to and activating OBR-L surface levels of the leptin receptors. In breast
(Elmquist et al. 1998; Scott et al. 2009). In paral- cancer subjects with diet-induced obesity, STAT3
lel with its activation of AMPK, leptin suppresses blockade suppresses the leptin receptor expres-
the activity of acetyl coenzyme A carboxylase, sion and inhibits tumor progression (Chang et al.
subsequently stimulates the oxidation of fatty 2015). Most of obese humans are characterized
acids in muscle. (Minokoshi et al. 2002). OBR-L by leptin resistance (Maffei et al. 1995). Leptin
is capable of signaling to IRS-1 and MAPK via specifically induces expression of SOCS-3
JAK, in addition to activating STAT pathways mRNA (Bjørbaek et al. 1998). Thus, increased
(Bjørbaek et al. 1997). The short isoform (OBR-­ levels of SOCS-3 in peripheral tissues may there-
S) is inactive in both proliferation and JAK acti- fore result in leptin resistance at these sites.
vation. Thereby, resistance to leptin occurs SOCS-3 binds to JAK2 in a leptin-dependent
despite the presence of the OBR-S isoforms manner and suggest that SOCS-3 attenuates
(Ghilardi and Skoda 1997). Binding of the SH2 leptin receptor signaling by inhibiting JAK-­
domain of SH2B1 to phospho-Tyr813 in JAK2 induced tyrosine phosphorylation of the receptor
enhances the leptin induction of JAK2. Actually, and of JAK itself (Bjørbaek et al. 1999).
SH2B1 acts as signal transducer and signaling Collectively, the deterioration of the leptin sig-
adaptor in obesity and its deletion is associated nals results in obesity. Leptin resistance in obe-
with severe obesity. Overexpression of SH2B1 sity may occur with various mechanisms such as;
enhances JAK2-mediated tyrosine phosphoryla- reduction in leptin transport into the brain, defects
tion of IRS-1 in response to leptin (Li et al. 2007). in OBR-L trafficking, as well as in OBR-L
592 A. Engin

expression, increase in SOCS3 expression, receptor isoforms positively correlated with pro-
chronic endoplasmic reticulum stress by activa- tein levels of estradiol and progesterone receptors.
tion of various unfolded protein response-­ The PR isoforms’ mRNA levels are inversely cor-
signaling pathways and impairment of related with clinical-pathological markers of
melanocortin-4 or brain-derived neurotrophic tumor aggressiveness. Furthermore, the proges-
factor receptors (Morris and Rui 2009). terone receptor isoforms are positively correlated
On the other hand, in approximately 30% of with the mRNA levels of HER/ErbB receptors
population, leptin is able to stimulate aromatase and ligands which are associated with more dif-
activity in adipose stromal cells at high concentra- ferentiated phenotypes of breast cancer (Lindet
tions. The elevated levels of aromatase activity et al. 2012). An interaction between leptin signal-
may contribute to increase in the circulating estro- ing and the transmembrane tyrosine kinase recep-
gen levels in obese women. In postmenopausal tor HER2 has been shown in breast cancer. Leptin
obese women, adipose tissue is the only place of receptor and HER2 are co-­ expressed in these
estrogen production by aromatization of C19 ste- tumors and leptin/OBR system contributes to the
roid androstenedione. Hence, the total pool of enhanced HER2 activity and reduced sensitivity
estrogens is increased by the aromatization pro- to anti-HER2 treatments (Fiorio et al. 2008).
cess in postmenopausal obese women (Magoffin Overexpression of HER2 in a series of breast car-
et al. 1999). Plasma leptin levels correlate to cinoma cells increases the aldehyde dehydroge-
plasma levels of estradiol and estrone sulfate in nase-expressing cancer stem cell population
breast cancer patients as well as in healthy post- which displays increased invasion and increased
menopausal females. In addition, plasma leptin tumorigenesis (Korkaya et al. 2008). Furthermore,
levels also correlate to total body aromatization in different human breast cancer cell lines, leptin
rate in breast cancer patients with obesity (Geisler enhances the expression of a chaperone pro-
et al. 2007). Factors that increase endogenous tein heat shock protein 90 (Hsp90). Eventually,
estrogen production or reduce the binding of estra- HER2 protein levels are increased. In contrast,
diol to sex hormone-binding globulin may increase silencing of Hsp90 gene expression by RNA
a woman’s risk of developing breast cancer later in interference inhibits leptin-­mediated HER2 up-
life (Toniolo et al. 1995). On the other hand, regulation. The adipocyte-­secreted leptin modu-
PPAR-gamma ligands show an inhibitory effect on lates Hsp90/HER2 expressions in breast cancer
the growth of breast cancer cells. PPAR-gamma is cells. This mechanism links the obesity to breast
a member of the nuclear receptor family of ligand- cancer growth and progression. However, long-
dependent transcription factors (Elstner et al. term leptin exposure reduces sensitivity of breast
1998). In this respect, activation of PPAR-gamma cancer cells to the anti-estrogen agents (Giordano
also decreases leptin receptors, inhibits their trans- et al. 2013). Synergy between the leptin/leptin
ductional pathways, and negatively interferes with receptor/STAT3 signaling pathway and the HER2
estrogen signaling through the down-­regulation of receptor protects tamoxifen-treated HER2 over-­
aromatase gene expression and the inhibition of expressing cells from the inhibitory effect of
ER-alpha transactivation (Catalano et al. 2011). tamoxifen through differential regulation of apop-
Twenty-four per cent of 1352 total patients tosis-related genes (Papanikolaou et al. 2015).
with breast cancer, are classified as obese. When Leptin enhances cyclin D1 gene transcription by
the obese patients are compared with the normal inducing the binding of ER-alpha to the promoter
weight cancer patients, obesity is found to be of cyclin D1 gene. In contrast, leptin receptor
associated with non-palpable tumors, larger deficiency significantly enhances the inhibitory
tumors, a higher incidence of lymph node metas- effects of tamoxifen on tamoxifen-­ resistance-­
tasis, lower incidence of HER2 positivity, lower breast cancer cell proliferation and survival.
incidence of multifocality (Haakinson et al. However, long-term endocrine therapy facilitates
2012). Actually, mRNA levels of the progesterone leptin and leptin receptor overexpression in breast
25 Obesity-associated Breast Cancer: Analysis of risk factors 593

cancer cells, which attenuates the inhibitory effect and are members of the Delta/Serrate/LAG-2
of tamoxifen by activating both the ERK1/2 and family of proteins. Translocation of the Notch
STAT3 signaling pathways and upregulating intracellular domain into the nucleus induces the
cyclin D1 gene expression (Qian et al. 2015). transcriptional activation of Notch target genes.
Collectively, growth of malignant cells could The carcinogenesis process is reinforced by Notch
be regulated by various leptin-induced second crosstalk with many oncogenic signaling path-
messengers like STAT3, AP1, MAPK and extra- ways suggest that Notch signaling may be a criti-
cellular signal-regulated kinases (ERKs). They cal drug target for breast cancer (Guo et al. 2011).
are all involved in aromatase expression, genera-
tion of estrogens and activation of ER-alpha in
malignant breast epithelium (Sulkowska et al. 11 Conclusion
2006). In this respect, leptin switches on tran-
scription of aromatase, and activates this enzyme Obese women have elevated risks of ductal and
with engagement of AP1 promoter, STAT3 and ER-positive-PR-positive breast cancer com-
ERK2 by promoting estradiol synthesis (Catalano pared to lean subjects. Estrogen-plus-progestin
et al. 2003). Leptin also activates ER-alpha via therapy for more than 5 years significantly
the MAPK pathway (Catalano et al. 2004). increases risks of lobular and ER positive-PR
ER-alpha36 is mainly expressed on the cell sur- positive breast cancer compared to never users
face and mediates membrane-initiated or non-­ of hormone therapy regardless of BMI. In addi-
genomic estrogen signaling (Wang et al. 2005). tion, elevated estrogen concentrations in post-
Hence, in ER-negative breast cancer cells Src menopausal breast cancer patients are largely
acts as a switch in ER-alpha36-mediated biphasic derived from obese adipose tissue aromatiza-
estrogen signaling through the Src/EGFR/STAT5 tion. Most breast cancer cell types are addicted
pathway. Similar to ER-positive breast cancer to fatty acids. Hence, elevation of circulating
cells, ER-alpha36 mediates mitogenic signaling free fatty acids in obesity is associated with
of low-concentration estrogen through the EGFR/ enhanced cancer risk. The obesity-inflamma-
Src/STAT5 pathway in ER-negative breast cancer tion-aromatase axis is an indicator of the
cells (Zhang et al. 2012). After exposure to leptin, increased risk of hormone receptor-­ positive
ER-alpha-positive breast cancer cells undergo breast cancer and is a sign of poor prognosis.
proliferation. This process is mediated by active Increase in risk of recurrence and death in breast
STAT3 and ERK1/2. However, elevated serum cancer is significantly higher within 10 years of
levels of leptin maintain resistance to anti-­ diagnosis in obese women. Low-circulating adi-
estrogen drugs during hormonal therapy of breast ponectin levels in breast cancer patients are
cancer (Garofalo et al. 2004). Despite 12 to associated with the increased rate of more
20-fold higher leptin levels of obese mice than aggressive, high-histological grade, estrogen
that of lean mice, malfunction leptin receptors receptor negative, and metastatic tumors.
effectively protect against mammary tumor inci- Hyperinsulinemia is an independent risk factor
dence (Cleary et al. 2004b). for poor prognosis in breast cancer, and is asso-
Actually, there is a crosstalk between Notch, ciated with low levels of adiponectin and shorter
IL-1 and leptin in breast cancer. Leptin induction breast cancer survival in obesity. Serum leptin
of proliferation/migration and upregulation of levels, leptin-BMI and leptin-adiponectin ratios
VEGF/VEGFR-2 in breast cancer cells are related are significantly higher in obese patients with
to an intact Notch signaling axis (Guo and breast cancer.
Gonzalez-Perez 2011). However, leptin upregula- Body weight control and avoiding estrogen-­
tion of VEGF/VEGFR2 is impaired by IL-1 sig- plus-­progestin therapy is an effective strategy
naling blockade (Zhou et al. 2011). The notch for primary prevention of breast cancer among
ligands are single-pass transmembrane proteins post-­menopausal women. Because of these,
594 A. Engin

recommendations should be taken into account lates cell cycle progression in human breast cancer
cells via a functional interaction with cyclin-­dependent
also for young adult pre-menopausal women,
kinase 4. Molecular Pharmacology 77: 195–201.
obesity has impacts on the responsiveness to doi:10.1124/mol.109.059675.
the endocrine therapy with aromatase inhibi- Bartella, V., S. Cascio, E. Fiorio, A. Auriemma, A. Russo,
tors of pre-­menopausal breast cancer patients. and E. Surmacz. 2008. Insulin-dependent leptin
expression in breast cancer cells. Cancer Research 68:
Excessive weight gain in premenopausal
4919–4927. doi:10.1158/0008-5472.CAN-08-0642.
women is associated with the increase in Benn, M., A. Tybjærg-Hansen, G.D. Smith, and
relapse risk and higher mortality rate in breast B.G. Nordestgaard. 2016. High body mass index and
cancer. cancer risk—A Mendelian randomisation study.
European Journal of Epidemiology 31: 879–892.
doi:10.1007/s10654-016-0147-5.
Benoy, I., R. Salgado, C. Colpaert, R. Weytjens,
References P.B. Vermeulen, and L.Y. Dirix. 2002. Serum interleu-
kin 6, plasma VEGF, serum VEGF, and VEGF platelet
Alikhani, N., R.D. Ferguson, R. Novosyadlyy, load in breast cancer patients. Clinical Breast Cancer
E.J. Gallagher, E.J. Scheinman, S. Yakar, and 2: 311–315.
D. LeRoith. 2013. Mammary tumor growth and pul- Biglia, N., E. Peano, P. Sgandurra, G. Moggio, S. Pecchio,
monary metastasis are enhanced in a hyperlipidemic F. Maggiorotto, and P. Sismondi. 2013. Body mass
mouse model. Oncogene 32: 961–967. doi:10.1038/ index (BMI) and breast cancer: Impact on tumor histo-
onc.2012.113. pathologic features, cancer subtypes and recurrence
Anderson, G.L., M. Limacher, A.R. Assaf, T. Bassford, rate in pre and postmenopausal women. Gynecological
S.A.A. Beresford, H. Black, D. Bonds, R. Brunner, Endocrinology 29: 263–267. doi:10.3109/09513590.2
R. Brzyski, B. Caan, R. Chlebowski, D. Curb, 012.736559.
M. Gass, J. Hays, G. Heiss, S. Hendrix, B.V. Howard, Bjørbaek, C., S. Uotani, B. da Silva, and J.S. Flier. 1997.
J. Hsia, A. Hubbell, R. Jackson, K.C. Johnson, Divergent signaling capacities of the long and short
H. Judd, J.M. Kotchen, L. Kuller, A.Z. LaCroix, isoforms of the leptin receptor. The Journal of
D. Lane, R.D. Langer, N. Lasser, C.E. Lewis, Biological Chemistry 272: 32686–32695.
J. Manson, K. Margolis, J. Ockene, M.J. O’Sullivan, Bjørbaek, C., J.K. Elmquist, J.D. Frantz, S.E. Shoelson,
L. Phillips, R.L. Prentice, C. Ritenbaugh, J. Robbins, and J.S. Flier. 1998. Identification of SOCS-3 as a
J.E. Rossouw, G. Sarto, M.L. Stefanick, L. Van Horn, potential mediator of central leptin resistance.
J. Wactawski-Wende, R. Wallace, S. Wassertheil-­ Molecular Cell 1: 619–625.
Smoller, and Women’s Health Initiative Steering Bjørbaek, C., K. El-Haschimi, J.D. Frantz, and J.S. Flier.
Committee. 2004. Effects of conjugated equine estro- 1999. The role of SOCS-3 in leptin signaling and
gen in postmenopausal women with hysterectomy: leptin resistance. The Journal of Biological Chemistry
The Women’s Health Initiative randomized controlled 274: 30059–30065.
trial. Journal of the American Medical Association Blancafort, A., A. Giró-Perafita, G. Oliveras, S. Palomeras,
291: 1701–1712. doi:10.1001/jama.291.14.1701. C. Turrado, Ò. Campuzano, D. Carrión-Salip,
Badache, A., and N.E. Hynes. 2001. Interleukin 6 inhibits A. Massaguer, R. Brugada, M. Palafox, J. Gómez-­
proliferation and, in cooperation with an epidermal Miragaya, E. González-Suárez, and T. Puig. 2015.
growth factor receptor autocrine loop, increases Dual fatty acid synthase and HER2 signaling blockade
migration of T47D breast cancer cells. Cancer shows marked antitumor activity against breast cancer
Research 61: 383–391. models resistant to anti-HER2 drugs. PLoS One 10:
Bagga, D., S. Capone, H.J. Wang, D. Heber, M. Lill, e0131241. doi:10.1371/journal.pone.0131241.
L. Chap, and J.A. Glaspy. 1997. Dietary modulation of Bolton, J.L., and G.R.J. Thatcher. 2008. Potential mecha-
omega-3/omega-6 polyunsaturated fatty acid ratios in nisms of estrogen quinone carcinogenesis. Chemical
patients with breast cancer. Journal of the National Research in Toxicology 21: 93–101. doi:10.1021/
Cancer Institute 89: 1123–1131. tx700191p.
Bagga, D., K.H. Anders, H.-J. Wang, and J.A. Glaspy. Boonyaratanakornkit, V., and P. Pateetin. 2015. The role
2002. Long-chain n-3-to-n-6 polyunsaturated fatty of ovarian sex steroids in metabolic homeostasis,
acid ratios in breast adipose tissue from women with ­obesity, and postmenopausal breast cancer: Molecular
and without breast cancer. Nutrition and Cancer 42: mechanisms and therapeutic implications. BioMed
180–185. doi:10.1207/S15327914NC422_5. Research International 2015: 140196. doi:10.1155/
Barb, D., K. Pazaitou-Panayiotou, and C.S. Mantzoros. 2015/140196.
2006. Adiponectin: A link between obesity and can- Bowers, L.W., D.A. Cavazos, I.X.F. Maximo,
cer. Expert Opinion on Investigational Drugs 15: 917– A.J. Brenner, S.D. Hursting, and L.A. deGraffenried.
931. doi:10.1517/13543784.15.8.917. 2013. Obesity enhances nongenomic estrogen recep-
Barhoover, M.A., J.M. Hall, W.F. Greenlee, and tor crosstalk with the PI3K/Akt and MAPK pathways
R.S. Thomas. 2010. Aryl hydrocarbon receptor regu- to promote in vitro measures of breast cancer progres-
25 Obesity-associated Breast Cancer: Analysis of risk factors 595

sion. Breast Cancer Research: BCR 15: R59. Catalano, S., L. Mauro, S. Marsico, C. Giordano, P. Rizza,
doi:10.1186/bcr3453. V. Rago, D. Montanaro, M. Maggiolini, M.L. Panno,
Bowers, L.W., A.J. Brenner, S.D. Hursting, R.R. Tekmal, and S. Andó. 2004. Leptin induces, via ERK1/ERK2
and L.A. deGraffenried. 2015. Obesity-associated signal, functional activation of estrogen receptor alpha
systemic interleukin-6 promotes pre-adipocyte aro- in MCF-7 cells. The Journal of Biological Chemistry
matase expression via increased breast cancer cell 279: 19908–19915. doi:10.1074/jbc.M313191200.
prostaglandin E2 production. Breast Cancer Research Catalano, S., L. Mauro, D. Bonofiglio, M. Pellegrino,
and Treatment 149: 49–57. doi:10.1007/s10549- H. Qi, P. Rizza, D. Vizza, G. Bossi, and S. Andò. 2011.
014-3223-0. In vivo and in vitro evidence that PPARγ ligands are
Boyd, N.F., and V. McGuire. 1990. Evidence of associa- antagonists of leptin signaling in breast cancer. The
tion between plasma high-density lipoprotein choles- American Journal of Pathology 179: 1030–1040.
terol and risk factors for breast cancer. Journal of the doi:10.1016/j.ajpath.2011.04.026.
National Cancer Institute 82: 460–468. Cavalieri, E., D. Chakravarti, J. Guttenplan, E. Hart,
Brodeur, M.R., L. Brissette, L. Falstrault, V. Luangrath, J. Ingle, R. Jankowiak, P. Muti, E. Rogan, J. Russo,
and R. Moreau. 2008. Scavenger receptor of class B R. Santen, and T. Sutter. 2006. Catechol estrogen qui-
expressed by osteoblastic cells are implicated in the nones as initiators of breast and other human cancers:
uptake of cholesteryl ester and estradiol from LDL Implications for biomarkers of susceptibility and can-
and HDL3. Journal of Bone and Mineral Research: cer prevention. Biochimica et Biophysica Acta 1766:
The Official Journal of the American Society for Bone 63–78. doi:10.1016/j.bbcan.2006.03.001.
and Mineral Research 23: 326–337. doi:10.1359/ Chan, D.S.M., A.R. Vieira, D. Aune, E.V. Bandera,
jbmr.071022. D.C. Greenwood, A. McTiernan, D. Navarro
Brodie, A., Q. Lu, and J. Nakamura. 1997. Aromatase in Rosenblatt, I. Thune, R. Vieira, and T. Norat. 2014.
the normal breast and breast cancer. The Journal of Body mass index and survival in women with breast
Steroid Biochemistry and Molecular Biology 61: cancer-systematic literature review and meta-analysis
281–286. of 82 follow-up studies. Annals of Oncology 25: 1901–
Calder, P.C. 2012. The role of marine omega-3 (n-3) fatty 1914. doi:10.1093/annonc/mdu042.
acids in inflammatory processes, atherosclerosis and Chang, C.-C., M.-J. Wu, J.-Y. Yang, I.G. Camarillo, and
plaque stability. Molecular Nutrition & Food Research C.-J. Chang. 2015. Leptin-STAT3-G9a signaling pro-
56: 1073–1080. doi:10.1002/mnfr.201100710. motes obesity-mediated breast cancer progression.
Calle, E.E., H.S. Feigelson, J.S. Hildebrand, L.R. Teras, Cancer Research 75: 2375–2386. doi:10.1158/0008-
M.J. Thun, and C. Rodriguez. 2009. Postmenopausal ­5472.CAN-14-3076.
hormone use and breast cancer associations differ by Chavey, C., F. Bibeau, S. Gourgou-Bourgade,
hormone regimen and histologic subtype. Cancer 115: S. Burlinchon, F. Boissière, D. Laune, S. Roques, and
936–945. doi:10.1002/cncr.24101. G. Lazennec. 2007. Oestrogen receptor negative breast
Camoriano, J.K., C.L. Loprinzi, J.N. Ingle, T.M. Therneau, cancers exhibit high cytokine content. Breast Cancer
J.E. Krook, and M.H. Veeder. 1990. Weight change in Research: BCR 9: R15. doi:10.1186/bcr1648.
women treated with adjuvant therapy or observed fol- Chen, X., and Y. Wang. 2011. Adiponectin and breast can-
lowing mastectomy for node-positive breast cancer. cer. Medical Oncology (Northwood, London, England)
Journal of Clinical Oncology: Official Journal of the 28: 1288–1295. doi:10.1007/s12032-010-9617-x.
American Society of Clinical Oncology 8: 1327–1334. Chen, D.-C., Y.-F. Chung, Y.-T. Yeh, H.-C. Chaung,
doi:10.1200/jco.1990.8.8.1327. F.-C. Kuo, O.-Y. Fu, H.-Y. Chen, M.-F. Hou, and
Campbell, K.L., K.E. Foster-Schubert, C.M. Alfano, S.-S.F. Yuan. 2006. Serum adiponectin and leptin lev-
C.-C. Wang, C.-Y. Wang, C.R. Duggan, C. Mason, els in Taiwanese breast cancer patients. Cancer Letters
I. Imayama, A. Kong, L. Xiao, C.E. Bain, 237: 109–114. doi:10.1016/j.canlet.2005.05.047.
G.L. Blackburn, F.Z. Stanczyk, and A. McTiernan. Chen, D., H. Zhao, J.S. Coon, M. Ono, E.K. Pearson, and
2012. Reduced-calorie dietary weight loss, exercise, S.E. Bulun. 2012. Weight gain increases human aro-
and sex hormones in postmenopausal women: matase expression in mammary gland. Molecular and
Randomized controlled trial. Journal of Clinical Cellular Endocrinology 355: 114–120. doi:10.1016/j.
Oncology: Official Journal of the American Society of mce.2012.01.027.
Clinical Oncology 30: 2314–2326. doi:10.1200/ Chlebowski, R.T., G.L. Anderson, M. Gass, D.S. Lane,
JCO.2011.37.9792. A.K. Aragaki, L.H. Kuller, J.E. Manson,
Carroll, P.A., L. Healy, J. Lysaght, T. Boyle, J.V. Reynolds, M.L. Stefanick, J. Ockene, G.E. Sarto, K.C. Johnson,
M.J. Kennedy, G. Pidgeon, and E.M. Connolly. 2011. J. Wactawski-Wende, P.M. Ravdin, R. Schenken,
Influence of the metabolic syndrome on leptin and S.L. Hendrix, A. Rajkovic, T.E. Rohan, S. Yasmeen,
leptin receptor in breast cancer. Molecular R.L. Prentice, and WHI Investigators. 2010. Estrogen
Carcinogenesis 50: 643–651. doi:10.1002/mc.20764. plus progestin and breast cancer incidence and mortal-
Catalano, S., S. Marsico, C. Giordano, L. Mauro, P. Rizza, ity in postmenopausal women. Journal of the American
M.L. Panno, and S. Andò. 2003. Leptin enhances, via Medical Association 304: 1684–1692. doi:10.1001/
AP-1, expression of aromatase in the MCF-7 cell line. jama.2010.1500.
The Journal of Biological Chemistry 278: 28668– Chlebowski, R.T., T.E. Rohan, J.E. Manson, A.K. Aragaki,
28676. doi:10.1074/jbc.M301695200. A. Kaunitz, M.L. Stefanick, M.S. Simon,
596 A. Engin

K.C. Johnson, J. Wactawski-Wende, M.J. O’Sullivan, Delort, L., A. Rossary, M.-C. Farges, M.-P. Vasson, and
L.L. Adams-Campbell, R. Nassir, L.S. Lessin, and F. Caldefie-Chézet. 2015. Leptin, adipocytes and
R.L. Prentice. 2015. Breast cancer after use of estro- breast cancer: Focus on inflammation and anti-tumor
gen plus progestin and estrogen alone: Analyses of immunity. Life Sciences 140: 37–48. doi:10.1016/j.
data from 2 Women’s Health Initiative randomized lfs.2015.04.012.
clinical trials. JAMA Oncology 1: 296–305. Demark-Wahnefried, W., K.L. Campbell, and S.C. Hayes.
doi:10.1001/jamaoncol.2015.0494. 2012. Weight management and its role in breast cancer
Cleary, M.P., and M.E. Grossmann. 2009. Minireview: rehabilitation. Cancer 118: 2277–2287. doi:10.1002/
Obesity and breast cancer: The estrogen connection. cncr.27466.
Endocrinology 150: 2537–2542. doi:10.1210/ DeMichele, A., R. Gray, M. Horn, J. Chen, R. Aplenc,
en.2009-0070. W.P. Vaughan, and M.S. Tallman. 2009. Host
Cleary, M.P., J.P. Grande, S.C. Juneja, and N.J. Maihle. genetic variants in the interleukin-6 promoter pre-
2004a. Diet-induced obesity and mammary tumor dict poor outcome in patients with estrogen recep-
development in MMTV-neu female mice. Nutrition tor-positive, node-positive breast cancer. Cancer
and Cancer 50: 174–180. doi:10.1207/s153279 Research 69: 4184–4191. doi:10.1158/0008-5472.
14nc5002_7. CAN-08-2989.
Cleary, M.P., S.C. Juneja, F.C. Phillips, X. Hu, J.P. Grande, Denison, M.S., A.A. Soshilov, G. He, D.E. DeGroot, and
and N.J. Maihle. 2004b. Leptin receptor-deficient B. Zhao. 2011. Exactly the same but different:
MMTV-TGF-alpha/Lepr(db)Lepr(db) female mice do Promiscuity and diversity in the molecular mecha-
not develop oncogene-induced mammary tumors. nisms of action of the aryl hydrocarbon (dioxin) recep-
Experimental Biology and Medicine (Maywood, NJ) tor. Toxicological Sciences 124: 1–22. doi:10.1093/
229: 182–193. toxsci/kfr218.
Cooke, P.S., P.A. Heine, J.A. Taylor, and D.B. Lubahn. Deroo, B.J., and K.S. Korach. 2006. Estrogen receptors
2001. The role of estrogen and estrogen receptor-alpha and human disease. The Journal of Clinical
in male adipose tissue. Molecular and Cellular Investigation 116: 561–570. doi:10.1172/JCI27987.
Endocrinology 178: 147–154. Dethlefsen, C., G. Højfeldt, and P. Hojman. 2013. The
D’Esposito, V., F. Passaretti, A. Hammarstedt, D. Liguoro, role of intratumoral and systemic IL-6 in breast can-
D. Terracciano, G. Molea, L. Canta, C. Miele, cer. Breast Cancer Research and Treatment 138: 657–
U. Smith, F. Beguinot, and P. Formisano. 2012. 664. doi:10.1007/s10549-013-2488-z.
Adipocyte-released insulin-like growth factor-1 is Dignam, J.J., K. Wieand, K.A. Johnson, B. Fisher, L. Xu,
regulated by glucose and fatty acids and controls and E.P. Mamounas. 2003. Obesity, tamoxifen use,
breast cancer cell growth in vitro. Diabetologia 55: and outcomes in women with estrogen receptor-­
2811–2822. doi:10.1007/s00125-012-2629-7. positive early-stage breast cancer. Journal of the
Dalamaga, M., K.N. Diakopoulos, and C.S. Mantzoros. National Cancer Institute 95: 1467–1476.
2012. The role of adiponectin in cancer: A review of Diorio, C., J. Lemieux, L. Provencher, J.-C. Hogue, and
current evidence. Endocrine Reviews 33: 547–594. E. Vachon. 2012. Aromatase inhibitors in obese breast
doi:10.1210/er.2011-1015. cancer patients are not associated with increased
Davis, S.R., C. Castelo-Branco, P. Chedraui, plasma estradiol levels. Breast Cancer Research and
M.A. Lumsden, R.E. Nappi, D. Shah, P. Villaseca, and Treatment 136: 573–579. doi:10.1007/
Writing Group of the International Menopause Society s10549-012-2278-z.
for World Menopause Day 2012. 2012. Understanding Dirat, B., L. Bochet, M. Dabek, D. Daviaud, S. Dauvillier,
weight gain at menopause. Climacteric: The Journal B. Majed, Y.Y. Wang, A. Meulle, B. Salles, S. Le
of the International Menopause Society 15: 419–429. Gonidec, I. Garrido, G. Escourrou, P. Valet, and
doi:10.3109/13697137.2012.707385. C. Muller. 2011. Cancer-associated adipocytes exhibit
DeBerardinis, R.J., J.J. Lum, G. Hatzivassiliou, and an activated phenotype and contribute to breast cancer
C.B. Thompson. 2008. The biology of cancer: invasion. Cancer Research 71: 2455–2465.
Metabolic reprogramming fuels cell growth and pro- doi:10.1158/0008-5472.CAN-10-3323.
liferation. Cell Metabolism 7: 11–20. doi:10.1016/j. Duan, C., M. Li, and L. Rui. 2004. SH2-B promotes insu-
cmet.2007.10.002. lin receptor substrate 1 (IRS1)- and IRS2-mediated
Del Giudice, M.E., I.G. Fantus, S. Ezzat, G. McKeown-­ activation of the phosphatidylinositol 3-kinase path-
Eyssen, D. Page, and P.J. Goodwin. 1998. Insulin and way in response to leptin. The Journal of Biological
related factors in premenopausal breast cancer risk. Chemistry 279: 43684–43691. doi:10.1074/jbc.
Breast Cancer Research and Treatment 47: 111–120. M408495200.
Delort, L., T. Jardé, V. Dubois, M.-P. Vasson, and Duggan, C., M.L. Irwin, L. Xiao, K.D. Henderson,
F. Caldefie-Chézet. 2012. New insights into anticar- A.W. Smith, R.N. Baumgartner, K.B. Baumgartner,
cinogenic properties of adiponectin: A potential thera- L. Bernstein, R. Ballard-Barbash, and A. McTiernan.
peutic approach in breast cancer? Vitamins and 2011. Associations of insulin resistance and adiponec-
Hormones 90: 397–417. doi:10.1016/ tin with mortality in women with breast cancer.
B978-0-12-398313-8.00015-4. Journal of Clinical Oncology: Official Journal of the
25 Obesity-associated Breast Cancer: Analysis of risk factors 597

American Society of Clinical Oncology 29: 32–39. Fu, X., J.G. Menke, Y. Chen, G. Zhou, K.L. MacNaul,
doi:10.1200/JCO.2009.26.4473. S.D. Wright, C.P. Sparrow, and E.G. Lund. 2001.
Elmquist, J.K., C. Bjørbaek, R.S. Ahima, J.S. Flier, and 27-hydroxycholesterol is an endogenous ligand for
C.B. Saper. 1998. Distributions of leptin receptor liver X receptor in cholesterol-loaded cells. The
mRNA isoforms in the rat brain. The Journal of Journal of Biological Chemistry 276: 38378–38387.
Comparative Neurology 395: 535–547. doi:10.1074/jbc.M105805200.
Elstner, E., C. Müller, K. Koshizuka, E.A. Williamson, Garofalo, C., D. Sisci, and E. Surmacz. 2004. Leptin
D. Park, H. Asou, P. Shintaku, J.W. Said, D. Heber, interferes with the effects of the antiestrogen ICI
and H.P. Koeffler. 1998. Ligands for peroxisome 182,780 in MCF-7 breast cancer cells. Clinical
proliferator-­activated receptorgamma and retinoic acid Cancer Research 10: 6466–6475. doi:10.1158/1078-
receptor inhibit growth and induce apoptosis of human 0432.CCR-04-0203.
breast cancer cells in vitro and in BNX mice. Geisler, J., B. Haynes, D. Ekse, M. Dowsett, and
Proceedings of the National Academy of Sciences of P.E. Lønning. 2007. Total body aromatization in post-
the United States of America 95: 8806–8811. menopausal breast cancer patients is strongly corre-
Esteban, J.M., Z. Warsi, M. Haniu, P. Hall, J.E. Shively, lated to plasma leptin levels. The Journal of Steroid
and S. Chen. 1992. Detection of intratumoral aroma- Biochemistry and Molecular Biology 104: 27–34.
tase in breast carcinomas. An immunohistochemical doi:10.1016/j.jsbmb.2006.09.040.
study with clinicopathologic correlation. The Gelaleti, G.B., B.V. Jardim, C. Leonel, M.G. Moschetta,
American Journal of Pathology 140: 337–343. and D.A. Zuccari. 2012. Interleukin-8 as a prognostic
Ewertz, M., M.-B. Jensen, K.Á. Gunnarsdóttir, I. Højris, serum marker in canine mammary gland neoplasias.
E.H. Jakobsen, D. Nielsen, L.E. Stenbygaard, Veterinary Immunology and Immunopathology 146:
U.B. Tange, and S. Cold. 2011. Effect of obesity on 106–112. doi:10.1016/j.vetimm.2012.02.005.
prognosis after early-stage breast cancer. Journal of Ghilardi, N., and R.C. Skoda. 1997. The leptin receptor
Clinical Oncology: Official Journal of the American activates janus kinase 2 and signals for proliferation in
Society of Clinical Oncology 29: 25–31. doi:10.1200/ a factor-dependent cell line. Molecular Endocrinology
JCO.2010.29.7614. (Baltimore, Md.) 11: 393–399. doi:10.1210/
Ewertz, M., K.P. Gray, M.M. Regan, B. Ejlertsen, mend.11.4.9907.
K.N. Price, B. Thürlimann, H. Bonnefoi, J.F. Forbes, Giordano, C., D. Vizza, S. Panza, I. Barone, D. Bonofiglio,
R.J. Paridaens, M. Rabaglio, R.D. Gelber, M. Colleoni, M. Lanzino, D. Sisci, F. De Amicis, S.A.W. Fuqua,
I. Láng, I.E. Smith, A.S. Coates, A. Goldhirsch, and S. Catalano, and S. Andò. 2013. Leptin increases HER2
H.T. Mouridsen. 2012. Obesity and risk of recurrence protein levels through a STAT3-mediated up-­regulation
or death after adjuvant endocrine therapy with letro- of Hsp90 in breast cancer cells. Molecular Oncology 7:
zole or tamoxifen in the breast international group 379–391. doi:10.1016/j.molonc.2012.11.002.
1-98 trial. Journal of Clinical Oncology: Official Giró-Perafita, A., S. Palomeras, D.H. Lum, A. Blancafort,
Journal of the American Society of Clinical Oncology G. Viñas, G. Oliveras, F. Pérez-Bueno, A. Sarrats,
30: 3967–3975. doi:10.1200/JCO.2011.40.8666. A.L. Welm, and T. Puig. 2016. Preclinical evaluation
Fedele, P., L. Orlando, P. Schiavone, A. Quaranta, of fatty acid synthase and EGFR inhibition in triple-­
A.M. Lapolla, M. De Pasquale, A. Ardizzone, E. Bria, negative breast cancer. Clinical Cancer Research 22:
I. Sperduti, N. Calvani, A. Marino, C. Caliolo, 4687–4697. doi:10.1158/1078-0432.CCR-15-3133.
E. Mazzoni, and S. Cinieri. 2014. BMI variation Gong, H., P. Guo, Y. Zhai, J. Zhou, H. Uppal,
increases recurrence risk in women with early-stage M.J. Jarzynka, W.-C. Song, S.-Y. Cheng, and W. Xie.
breast cancer. Future Oncology (London, England) 10: 2007. Estrogen deprivation and inhibition of breast
2459–2468. doi:10.2217/fon.14.180. cancer growth in vivo through activation of the orphan
Fiorio, E., A. Mercanti, M. Terrasi, R. Micciolo, A. Remo, nuclear receptor liver X receptor. Molecular
A. Auriemma, A. Molino, V. Parolin, B. Di Stefano, Endocrinology (Baltimore, Md.) 21: 1781–1790.
F. Bonetti, A. Giordano, G.L. Cetto, and E. Surmacz. doi:10.1210/me.2007-0187.
2008. Leptin/HER2 crosstalk in breast cancer: In vitro Goodwin, P.J. 2013. Obesity and endocrine therapy: Host
study and preliminary in vivo analysis. BMC Cancer factors and breast cancer outcome. Breast (Edinburgh,
8: 305. doi:10.1186/1471-2407-8-305. Scotland) 22(Suppl 2): S44–S47. doi:10.1016/j.
Frederich, R.C., A. Hamann, S. Anderson, B. Löllmann, breast.2013.07.008.
B.B. Lowell, and J.S. Flier. 1995. Leptin levels reflect Goodwin, P.J., M. Ennis, K.I. Pritchard, M.E. Trudeau,
body lipid content in mice: Evidence for diet-induced J. Koo, Y. Madarnas, W. Hartwick, B. Hoffman, and
resistance to leptin action. Nature Medicine 1: N. Hood. 2002. Fasting insulin and outcome in early-­
1311–1314. stage breast cancer: Results of a prospective cohort
Frystyk, J., C. Skjaerbaek, E. Vestbo, S. Fisker, and study. Journal of Clinical Oncology: Official Journal
H. Orskov. 1999. Circulating levels of free insulin-like of the American Society of Clinical Oncology 20:
growth factors in obese subjects: The impact of type 2 42–51. doi:10.1200/jco.2002.20.1.42.
diabetes. Diabetes/Metabolism Research and Reviews Goodwin, P.J., M. Ennis, K.I. Pritchard, M.E. Trudeau,
15: 314–322. J. Koo, S.K. Taylor, and N. Hood. 2012. Insulin- and
598 A. Engin

obesity-related variables in early-stage breast cancer: Hancke, K., D. Grubeck, N. Hauser, R. Kreienberg, and
Correlations and time course of prognostic associa- J.M. Weiss. 2010. Adipocyte fatty acid-binding pro-
tions. Journal of Clinical Oncology: Official Journal tein as a novel prognostic factor in obese breast cancer
of the American Society of Clinical Oncology 30: 164– patients. Breast Cancer Research and Treatment 119:
171. doi:10.1200/JCO.2011.36.2723. 367–367. doi:10.1007/s10549-009-0577-9.
Grewal, T., I. de Diego, M.F. Kirchhoff, F. Tebar, Hardy, S., G.G. St-Onge, E. Joly, Y. Langelier, and
J. Heeren, F. Rinninger, and C. Enrich. 2003. High M. Prentki. 2005. Oleate promotes the proliferation of
density lipoprotein-induced signaling of the MAPK breast cancer cells via the G protein-coupled receptor
pathway involves scavenger receptor type BI-mediated GPR40. The Journal of Biological Chemistry 280:
activation of Ras. The Journal of Biological Chemistry 13285–13291. doi:10.1074/jbc.M410922200.
278: 16478–16481. doi:10.1074/jbc.C300085200. Hawsawi, Y., R. El-Gendy, C. Twelves, V. Speirs, and
Grossmann, M.E., A. Ray, K.J. Nkhata, D.A. Malakhov, J. Beattie. 2013. Insulin-like growth factor—
O.P. Rogozina, S. Dogan, and M.P. Cleary. 2010. Oestradiol crosstalk and mammary gland tumourigen-
Obesity and breast cancer: Status of leptin and esis. Biochimica et Biophysica Acta 1836: 345–353.
adiponectin in pathological processes. Cancer doi:10.1016/j.bbcan.2013.10.005.
Metastasis Reviews 29: 641–653. doi:10.1007/ Heim, M.H. 1999. The Jak-STAT pathway: Cytokine sig-
s10555-010-9252-1. nalling from the receptor to the nucleus. Journal of
Guaita-Esteruelas, S., A. Bosquet, P. Saavedra, J. Gumà, Receptor and Signal Transduction Research 19:
J. Girona, E.W.-F. Lam, K. Amillano, J. Borràs, and 75–120. doi:10.3109/10799899909036638.
L. Masana. 2016. Exogenous FABP4 increases breast Heldring, N., A. Pike, S. Andersson, J. Matthews,
cancer cell proliferation and activates the expression G. Cheng, J. Hartman, M. Tujague, A. Ström,
of fatty acid transport proteins. Molecular E. Treuter, M. Warner, and J.-A. Gustafsson. 2007.
Carcinogenesis 56(1): 208–217. doi:10.1002/ Estrogen receptors: How do they signal and what are
mc.22485. their targets. Physiological Reviews 87: 905–931.
Gunter, M.J., D.R. Hoover, H. Yu, S. Wassertheil-Smoller, doi:10.1152/physrev.00026.2006.
T.E. Rohan, J.E. Manson, J. Li, G.Y.F. Ho, X. Xue, Henderson, B.E., and H.S. Feigelson. 2000. Hormonal
G.L. Anderson, R.C. Kaplan, T.G. Harris, carcinogenesis. Carcinogenesis 21: 427–433.
B.V. Howard, J. Wylie-Rosett, R.D. Burk, and Hernandez, A.V., M. Guarnizo, Y. Miranda, V. Pasupuleti,
H.D. Strickler. 2009. Insulin, insulin-like growth fac- A. Deshpande, S. Paico, H. Lenti, S. Ganoza,
tor-­
I, and risk of breast cancer in postmenopausal L. Montalvo, P. Thota, and H. Lazaro. 2014.
women. Journal of the National Cancer Institute 101: Association between insulin resistance and breast car-
48–60. doi:10.1093/jnci/djn415. cinoma: A systematic review and meta-analysis. PLoS
Guo, S., and R.R. Gonzalez-Perez. 2011. Notch, IL-1 and One 9: e99317. doi:10.1371/journal.pone.0099317.
leptin crosstalk outcome (NILCO) is critical for leptin-­ Hoeflich, A., and V.C. Russo. 2015. Physiology and
induced proliferation, migration and VEGF/VEGFR-2 pathophysiology of IGFBP-1 and IGFBP-2—
expression in breast cancer. PLoS One 6: e21467. Consensus and dissent on metabolic control and
doi:10.1371/journal.pone.0021467. malignant potential. Best Practice & Research.
Guo, S., M. Liu, and R.R. Gonzalez-Perez. 2011. Role of Clinical Endocrinology & Metabolism 29: 685–700.
Notch and its oncogenic signaling crosstalk in breast doi:10.1016/j.beem.2015.07.002.
cancer. Biochimica et Biophysica Acta 1815: 197–213. Honma, S., K. Shimodaira, Y. Shimizu, N. Tsuchiya,
doi:10.1016/j.bbcan.2010.12.002. H. Saito, T. Yanaihara, and T. Okai. 2002. The influ-
Guo, S., M. Liu, G. Wang, M. Torroella-Kouri, and ence of inflammatory cytokines on estrogen produc-
R.R. Gonzalez-Perez. 2012. Oncogenic role and thera- tion and cell proliferation in human breast cancer
peutic target of leptin signaling in breast cancer and cells. Endocrine Journal 49: 371–377.
cancer stem cells. Biochimica et Biophysica Acta Huang, Z., S.E. Hankinson, G.A. Colditz, M.J. Stampfer,
1825: 207–222. doi:10.1016/j.bbcan.2012.01.002. D.J. Hunter, J.E. Manson, C.H. Hennekens, B. Rosner,
Gupta, P.B., D. Proia, O. Cingoz, J. Weremowicz, F.E. Speizer, and W.C. Willett. 1997. Dual effects of
S.P. Naber, R.A. Weinberg, and C. Kuperwasser. 2007. weight and weight gain on breast cancer risk. Journal
Systemic stromal effects of estrogen promote the of the American Medical Association 278:
growth of estrogen receptor-negative cancers. Cancer 1407–1411.
Research 67: 2062–2071. doi:10.1158/0008-5472. Ishikawa, M., J. Kitayama, and H. Nagawa. 2004.
CAN-06-3895. Enhanced expression of leptin and leptin receptor
Haakinson, D.J., S.G. Leeds, A.C. Dueck, R.J. Gray, (OB-R) in human breast cancer. Clinical Cancer
N. Wasif, C.-C.H. Stucky, D.W. Northfelt, H.A. Apsey, Research 10: 4325–4331. doi:10.1158/1078-0432.
and B. Pockaj. 2012. The impact of obesity on breast CCR-03-0749.
cancer: A retrospective review. Annals of Surgical Iyengar, P., T.P. Combs, S.J. Shah, V. Gouon-Evans,
Oncology 19: 3012–3018. doi:10.1245/ J.W. Pollard, C. Albanese, L. Flanagan,
s10434-012-2320-8. M.P. Tenniswood, C. Guha, M.P. Lisanti, R.G. Pestell,
and P.E. Scherer. 2003. Adipocyte-secreted factors
25 Obesity-associated Breast Cancer: Analysis of risk factors 599

synergistically promote mammary tumorigenesis Prostaglandins, Leukotrienes, and Essential Fatty


through induction of anti-apoptotic transcriptional Acids 80: 93–99. doi:10.1016/j.plefa.2009.01.002.
programs and proto-oncogene stabilization. Oncogene Khaidakov, M., S. Mitra, B.-Y. Kang, X. Wang,
22: 6408–6423. doi:10.1038/sj.onc.1206737. S. Kadlubar, G. Novelli, V. Raj, M. Winters,
Iyengar, N.M., C.A. Hudis, and A.J. Dannenberg. 2013. W.C. Carter, and J.L. Mehta. 2011. Oxidized LDL
Obesity and inflammation: New insights into breast receptor 1 (OLR1) as a possible link between obesity,
cancer development and progression. American dyslipidemia and cancer. PLoS One 6: e20277.
Society of Clinical Oncology Educational Book. doi:10.1371/journal.pone.0020277.
American Society of Clinical Oncology. Meeting Khan, S., S. Shukla, S. Sinha, and S.M. Meeran. 2013.
46–51. doi:10.1200/EdBook_AM.2013.33.46. Role of adipokines and cytokines in obesity-­associated
Iyengar, N.M., X.K. Zhou, A. Gucalp, P.G. Morris, breast cancer: Therapeutic targets. Cytokine & Growth
L.R. Howe, D.D. Giri, M. Morrow, H. Wang, Factor Reviews 24: 503–513. doi:10.1016/j.
M. Pollak, L.W. Jones, C.A. Hudis, and cytogfr.2013.10.001.
A.J. Dannenberg. 2016. Systemic correlates of Khandekar, M.J., P. Cohen, and B.M. Spiegelman. 2011.
white adipose tissue inflammation in early-stage breast Molecular mechanisms of cancer development in obe-
cancer. Clinical Cancer Research 22: 2283–2289. sity. Nature Reviews. Cancer 11: 886–895.
doi:10.1158/1078-0432.CCR-15-2239. doi:10.1038/nrc3174.
Jackson, J.G., M.F. White, and D. Yee. 1998. Insulin Kim, H.-S., A.R. Ingermann, J. Tsubaki, S.M. Twigg,
receptor substrate-1 is the predominant signaling mol- G.E. Walker, and Y. Oh. 2004. Insulin-like growth
ecule activated by insulin-like growth factor-I, insulin, factor-binding protein 3 induces caspase-dependent
and interleukin-4 in estrogen receptor-positive human apoptosis through a death receptor-mediated pathway
breast cancer cells. The Journal of Biological in MCF-7 human breast cancer cells. Cancer Research
Chemistry 273: 9994–10003. 64: 2229–2237.
Jardé, T., F. Caldefie-Chézet, M. Damez, F. Mishellany, Kim, J., Y.H. Lee, J.H.Y. Park, and M.-K. Sung. 2015.
F. Penault-Llorca, J. Guillot, and M.P. Vasson. 2008. Estrogen deprivation and excess energy supply accel-
Leptin and leptin receptor involvement in cancer erate 7,12-dimethylbenz(a)anthracene-induced mam-
development: A study on human primary breast carci- mary tumor growth in C3H/HeN mice. Nutrition
noma. Oncology Reports 19: 905–911. Research and Practice 9: 628–636. doi:10.4162/
Kadowaki, T., and T. Yamauchi. 2005. Adiponectin and nrp.2015.9.6.628.
adiponectin receptors. Endocrine Reviews 26: 439– Kinlaw, W.B., P.W. Baures, L.E. Lupien, W.L. Davis, and
451. doi:10.1210/er.2005-0005. N.B. Kuemmerle. 2016. Fatty acids and breast cancer:
Kamel, M., S. Shouman, M. El-Merzebany, G. Kilic, Make them on site or have them delivered. Journal of
T. Veenstra, M. Saeed, M. Wagih, C. Diaz-Arrastia, Cellular Physiology 231: 2128–2141. doi:10.1002/
D. Patel, and S. Salama. 2012. Effect of tumour necro- jcp.25332.
sis factor-alpha on estrogen metabolic pathways in Korkaya, H., A. Paulson, F. Iovino, and M.S. Wicha.
breast cancer cells. Journal of Cancer 3: 310–321. 2008. HER2 regulates the mammary stem/progeni-
doi:10.7150/jca.4584. tor cell population driving tumorigenesis and inva-
Kamineni, A., M.L. Anderson, E. White, S.H. Taplin, sion. Oncogene 27: 6120–6130. doi:10.1038/
P. Porter, R. Ballard-Barbash, K. Malone, and onc.2008.207.
D.S.M. Buist. 2013. Body mass index, tumor charac- Körner, A., K. Pazaitou-Panayiotou, T. Kelesidis,
teristics, and prognosis following diagnosis of early-­ I. Kelesidis, C.J. Williams, A. Kaprara, J. Bullen,
stage breast cancer in a mammographically screened A. Neuwirth, S. Tseleni, N. Mitsiades, W. Kiess, and
population. Cancer Causes & Control: CCC 24: 305– C.S. Mantzoros. 2007. Total and high-molecular-­
312. doi:10.1007/s10552-012-0115-7. weight adiponectin in breast cancer: In vitro and
Kang, J.-H., B.-Y. Yu, and D.-S. Youn. 2007. Relationship in vivo studies. The Journal of Clinical Endocrinology
of serum adiponectin and resistin levels with breast and Metabolism 92: 1041–1048. doi:10.1210/
cancer risk. Journal of Korean Medical Science 22: jc.2006-1858.
117–121. doi:10.3346/jkms.2007.22.1.117. Kroenke, C.H., W.Y. Chen, B. Rosner, and M.D. Holmes.
Karuna, R., A.G. Holleboom, M.M. Motazacker, 2005. Weight, weight gain, and survival after breast
J.A. Kuivenhoven, R. Frikke-Schmidt, A. Tybjaerg-­ cancer diagnosis. Journal of Clinical Oncology:
Hansen, S. Georgopoulos, M. van Eck, T.J.C. van Official Journal of the American Society of Clinical
Berkel, A. von Eckardstein, and K.M. Rentsch. 2011. Oncology 23: 1370–1378. doi:10.1200/
Plasma levels of 27-hydroxycholesterol in humans and JCO.2005.01.079.
mice with monogenic disturbances of high density Kumar-Sinha, C., K.W. Ignatoski, M.E. Lippman,
lipoprotein metabolism. Atherosclerosis 214: 448– S.P. Ethier, and A.M. Chinnaiyan. 2003. Transcriptome
455. doi:10.1016/j.atherosclerosis.2010.10.042. analysis of HER2 reveals a molecular connection to
Kaur, B., A. Jørgensen, and A.K. Duttaroy. 2009. Fatty fatty acid synthesis. Cancer Research 63: 132–139.
acid uptake by breast cancer cells (MDA-MB-231): Kurtzman, S.H., K.H. Anderson, Y. Wang, L.J. Miller,
Effects of insulin, leptin, adiponectin, and TNFalpha. M. Renna, M. Stankus, R.R. Lindquist, G. Barrows,
and D.L. Kreutzer. 1999. Cytokines in human breast
600 A. Engin

cancer: IL-1alpha and IL-1beta expression. Oncology F. Kamme, and T.W. Lovenberg. 2009. Lactate inhib-
Reports 6: 65–70. its lipolysis in fat cells through activation of an orphan
Lahmann, P.H., K. Hoffmann, N. Allen, C.H. van Gils, G-protein-coupled receptor, GPR81. The Journal of
K.-T. Khaw, B. Tehard, F. Berrino, A. Tjønneland, Biological Chemistry 284: 2811–2822. doi:10.1074/
J. Bigaard, A. Olsen, K. Overvad, F. Clavel-Chapelon, jbc.M806409200.
G. Nagel, H. Boeing, D. Trichopoulos, G. Economou, Llaverias, G., C. Danilo, I. Mercier, K. Daumer,
G. Bellos, D. Palli, R. Tumino, S. Panico, C. Sacerdote, F. Capozza, T.M. Williams, F. Sotgia, M.P. Lisanti,
V. Krogh, P.H.M. Peeters, H.B. Bueno-de-Mesquita, and P.G. Frank. 2011. Role of cholesterol in the devel-
E. Lund, E. Ardanaz, P. Amiano, G. Pera, J.R. Quirós, opment and progression of breast cancer. The
C. Martínez, M.J. Tormo, E. Wirfält, G. Berglund, American Journal of Pathology 178: 402–412.
G. Hallmans, T.J. Key, G. Reeves, S. Bingham, doi:10.1016/j.ajpath.2010.11.005.
T. Norat, C. Biessy, R. Kaaks, and E. Riboli. 2004. Lønning, P.E., B.P. Haynes, and M. Dowsett. 2014.
Body size and breast cancer risk: Findings from the Relationship of body mass index with aromatisation
European Prospective Investigation into Cancer And and plasma and tissue oestrogen levels in postmeno-
Nutrition (EPIC). International Journal of Cancer pausal breast cancer patients treated with aromatase
111: 762–771. doi:10.1002/ijc.20315. inhibitors. European Journal of Cancer (Oxford,
Lashinger, L.M., L.M. Malone, M.J. McArthur, England: 1990) 1990(50): 1055–1064. doi:10.1016/j.
J.A. Goldberg, E.A. Daniels, A. Pavone, J.K. Colby, ejca.2014.01.007.
N.C. Smith, S.N. Perkins, S.M. Fischer, and Lovejoy, J.C. 2003. The menopause and obesity. Primary
S.D. Hursting. 2011. Genetic reduction of insulin- Care 30: 317–325.
like growth factor-1 mimics the anticancer effects of Lukanova, A., E. Lundin, A. Zeleniuch-Jacquotte, P. Muti,
calorie restriction on cyclooxygenase-2-driven pan- A. Mure, S. Rinaldi, L. Dossus, A. Micheli, A. Arslan,
creatic neoplasia. Cancer Prevention Research P. Lenner, R.E. Shore, V. Krogh, K.L. Koenig,
(Philadelphia, Pa.) 4: 1030–1040. doi:10.1158/1940- E. Riboli, F. Berrino, G. Hallmans, P. Stattin,
6207.CAPR-11-0027. P. Toniolo, and R. Kaaks. 2004. Body mass index, cir-
Lazarus, R., D. Sparrow, and S. Weiss. 1998. Temporal culating levels of sex-steroid hormones, IGF-I and
relations between obesity and insulin: Longitudinal IGF-binding protein-3: A cross-sectional study in
data from the Normative Aging Study. American healthy women. European Journal of Endocrinology
Journal of Epidemiology 147: 173–179. 150: 161–171.
Lee, J.Y., J. Ye, Z. Gao, H.S. Youn, W.H. Lee, L. Zhao, Macciò, A., and C. Madeddu. 2011. Obesity, inflamma-
N. Sizemore, and D.H. Hwang. 2003. Reciprocal tion, and postmenopausal breast cancer: Therapeutic
modulation of Toll-like receptor-4 signaling pathways implications. ScientificWorldJournal 11: 2020–2036.
involving MyD88 and phosphatidylinositol 3-kinase/ doi:10.1100/2011/806787.
AKT by saturated and polyunsaturated fatty acids. The Maddams, J., D. Brewster, A. Gavin, J. Steward, J. Elliott,
Journal of Biological Chemistry 278: 37041–37051. M. Utley, and H. Møller. 2009. Cancer prevalence in the
doi:10.1074/jbc.M305213200. United Kingdom: Estimates for 2008. British Journal of
Li, C.I., K.E. Malone, and J.R. Daling. 2006. Interactions Cancer 101: 541–547. doi:10.1038/sj.bjc.6605148.
between body mass index and hormone therapy and Madeddu, C., G. Gramignano, C. Floris, G. Murenu,
postmenopausal breast cancer risk (United States). G. Sollai, and A. Macciò. 2014. Role of inflammation
Cancer Causes & Control: CCC 17: 695–703. and oxidative stress in post-menopausal oestrogen-­
doi:10.1007/s10552-005-0001-7. dependent breast cancer. Journal of Cellular and
Li, Z., Y. Zhou, C. Carter-Su, M.G. Myers, and L. Rui. Molecular Medicine 18: 2519–2529. doi:10.1111/
2007. SH2B1 enhances leptin signaling by both Janus jcmm.12413.
kinase 2 Tyr813 phosphorylation-dependent and Maffei, M., J. Halaas, E. Ravussin, R.E. Pratley, G.H. Lee,
-independent mechanisms. Molecular Endocrinology Y. Zhang, H. Fei, S. Kim, R. Lallone, and
(Baltimore, Md.) 21: 2270–2281. doi:10.1210/ S. Ranganathan. 1995. Leptin levels in human and
me.2007-0111. rodent: Measurement of plasma leptin and ob RNA in
Lin, M.T., C.Y. Juan, K.J. Chang, W.J. Chen, and obese and weight-reduced subjects. Nature Medicine
M.L. Kuo. 2001. IL-6 inhibits apoptosis and retains 1: 1155–1161.
oxidative DNA lesions in human gastric cancer AGS Magoffin, D.A., S.R. Weitsman, S.K. Aagarwal, and
cells through up-regulation of anti-apoptotic gene A.J. Jakimiuk. 1999. Leptin regulation of aromatase
mcl-1. Carcinogenesis 22: 1947–1953. activity in adipose stromal cells from regularly cycling
Lindet, C., F. Révillion, V. Lhotellier, L. Hornez, women. Ginekologia Polska 70: 1–7.
J.-P. Peyrat, and J. Bonneterre. 2012. Relationships Mantzoros, C., E. Petridou, N. Dessypris, C. Chavelas,
between progesterone receptor isoforms and the HER/ M. Dalamaga, D.M. Alexe, Y. Papadiamantis,
ErbB receptors and ligands network in 299 primary C. Markopoulos, E. Spanos, G. Chrousos, and
breast cancers. The International Journal of Biological D. Trichopoulos. 2004. Adiponectin and breast cancer
Markers 27: e111–e117. doi:10.5301/JBM.2012.9198. risk. The Journal of Clinical Endocrinology and
Liu, C., J. Wu, J. Zhu, C. Kuei, J. Yu, J. Shelton, Metabolism 89: 1102–1107. doi:10.1210/
S.W. Sutton, X. Li, S.J. Yun, T. Mirzadegan, C. Mazur, jc.2003-031804.
25 Obesity-associated Breast Cancer: Analysis of risk factors 601

Mauro, L., M. Pellegrino, F. De Amicis, E. Ricchio, Millikan, R.C., B. Newman, C.-K. Tse, P.G. Moorman,
F. Giordano, P. Rizza, S. Catalano, D. Bonofiglio, K. Conway, L.G. Dressler, L.V. Smith, M.H. Labbok,
D. Sisci, M.L. Panno, and S. Andò. 2014. Evidences J. Geradts, J.T. Bensen, S. Jackson, S. Nyante,
that estrogen receptor α interferes with adiponectin C. Livasy, L. Carey, H.S. Earp, and C.M. Perou. 2008.
effects on breast cancer cell growth. Cell Cycle Epidemiology of basal-like breast cancer. Breast
(Georgetown, Texas) 13: 553–564. doi:10.4161/ Cancer Research and Treatment 109: 123–139.
cc.27455. doi:10.1007/s10549-007-9632-6.
Mauro, L., M. Pellegrino, F. Giordano, E. Ricchio, Minatoya, M., G. Kutomi, H. Shima, S. Asakura,
P. Rizza, F. De Amicis, S. Catalano, D. Bonofiglio, S. Otokozawa, H. Ohnishi, H. Akasaka, T. Miura,
M.L. Panno, and S. Andò. 2015. Estrogen receptor-α M. Mori, and K. Hirata. 2014. Relation of serum adi-
drives adiponectin effects on cyclin D1 expression in ponectin levels and obesity with breast cancer: A
breast cancer cells. FASEB Journal 29: 2150–2160. Japanese case-control study. Asian Pacific Journal of
doi:10.1096/fj.14-262808. Cancer Prevention: APJCP 15: 8325–8330.
McCusker, R.H., C. Camacho-Hubner, M.L. Bayne, Minokoshi, Y., Y.-B. Kim, O.D. Peroni, L.G.D. Fryer,
M.A. Cascieri, and D.R. Clemmons. 1990. Insulin-­ C. Müller, D. Carling, and B.B. Kahn. 2002. Leptin
like growth factor (IGF) binding to human fibroblast stimulates fatty-acid oxidation by activating AMP-­
and glioblastoma cells: The modulating effect of cell activated protein kinase. Nature 415: 339–343.
released IGF binding proteins (IGFBPs). Journal of doi:10.1038/415339a.
Cellular Physiology 144: 244–253. doi:10.1002/ Miyoshi, Y., T. Funahashi, S. Kihara, T. Taguchi,
jcp.1041440210. Y. Tamaki, Y. Matsuzawa, and S. Noguchi. 2003.
McDonnell, D.P., C.-Y. Chang, and E.R. Nelson. 2014a. Association of serum adiponectin levels with breast
The estrogen receptor as a mediator of the pathologi- cancer risk. Clinical Cancer Research 9: 5699–5704.
cal actions of cholesterol in breast cancer. Climacteric: Mohamed-Ali, V., S. Goodrick, A. Rawesh, D.R. Katz,
The Journal of the International Menopause Society J.M. Miles, J.S. Yudkin, S. Klein, and S.W. Coppack.
17(Suppl 2): 60–65. doi:10.3109/13697137.2014.966 1997. Subcutaneous adipose tissue releases interleu-
949. kin-­6, but not tumor necrosis factor-alpha, in vivo. The
McDonnell, D.P., S. Park, M.T. Goulet, J. Jasper, Journal of Clinical Endocrinology and Metabolism
S.E. Wardell, C.-Y. Chang, J.D. Norris, J.R. Guyton, 82: 4196–4200. doi:10.1210/jcem.82.12.4450.
and E.R. Nelson. 2014b. Obesity, cholesterol metabo- Monk, J.M., H.F. Turk, D.M. Liddle, A.A. De Boer,
lism, and breast cancer pathogenesis. Cancer Research K.A. Power, D.W.L. Ma, and L.E. Robinson. 2014.
74: 4976–4982. doi:10.1158/0008-5472. n-3 polyunsaturated fatty acids and mechanisms to
CAN-14-1756. mitigate inflammatory paracrine signaling in obesity-­
McTiernan, A., K.B. Rajan, S.S. Tworoger, M. Irwin, associated breast cancer. Nutrients 6: 4760–4793.
L. Bernstein, R. Baumgartner, F. Gilliland, doi:10.3390/nu6114760.
F.Z. Stanczyk, Y. Yasui, and R. Ballard-Barbash. Morad, V., A. Abrahamsson, and C. Dabrosin. 2014.
2003. Adiposity and sex hormones in postmenopausal Estradiol affects extracellular leptin:adiponectin ratio
breast cancer survivors. Journal of Clinical Oncology: in human breast tissue in vivo. The Journal of Clinical
Official Journal of the American Society of Clinical Endocrinology and Metabolism 99: 3460–3467.
Oncology 21: 1961–1966. doi:10.1200/ doi:10.1210/jc.2014-1129.
JCO.2003.07.057. Morris, D.L., and L. Rui. 2009. Recent advances in under-
McTiernan, A., L. Wu, C. Chen, R. Chlebowski, standing leptin signaling and leptin resistance.
Y. Mossavar-Rahmani, F. Modugno, M.G. Perri, American Journal of Physiology. Endocrinology and
F.Z. Stanczyk, L. Van Horn, C.Y. Wang, and Women’s Metabolism 297: E1247–E1259. doi:10.1152/
Health Initiative Investigators. 2006. Relation of BMI ajpendo.00274.2009.
and physical activity to sex hormones in postmeno- Morris, P.G., C.A. Hudis, D. Giri, M. Morrow,
pausal women. Obesity (Silver Spring, Md.) 14: 1662– D.J. Falcone, X.K. Zhou, B. Du, E. Brogi,
1677. doi:10.1038/oby.2006.191. C.B. Crawford, L. Kopelovich, K. Subbaramaiah, and
Merdad, A., S. Karim, H.-J. Schulten, M. Jayapal, A.J. Dannenberg. 2011. Inflammation and increased
A. Dallol, A. Buhmeida, F. Al-Thubaity, M.A. GariI, aromatase expression occur in the breast tissue of
A.G.A. Chaudhary, A.M. Abuzenadah, and obese women with breast cancer. Cancer Prevention
M.H. Al-Qahtani. 2015. Transcriptomics profiling Research (Philadelphia, Pa.) 4: 1021–1029.
study of breast cancer from Kingdom of Saudi Arabia doi:10.1158/1940-6207.CAPR-11-0110.
revealed altered expression of Adiponectin and Fatty Mueller, S.O., and K.S. Korach. 2001. Estrogen receptors
Acid Binding Protein4: Is lipid metabolism associated and endocrine diseases: Lessons from estrogen recep-
with breast cancer? BMC Genomics 16(Suppl 1): S11. tor knockout mice. Current Opinion in Pharmacology
doi:10.1186/1471-2164-16-S1-S11. 1: 613–619.
Miller, W.R., and J. O’Neill. 1987. The importance of Nakayama, S., Y. Miyoshi, H. Ishihara, and S. Noguchi.
local synthesis of estrogen within the breast. Steroids 2008. Growth-inhibitory effect of adiponectin via adi-
50: 537–548. ponectin receptor 1 on human breast cancer cells
602 A. Engin

through inhibition of S-phase entry without inducing K.C. Wood, J.W. Locasale, and D.P. McDonnell.
apoptosis. Breast Cancer Research and Treatment 2016. ERRα-regulated lactate metabolism contributes
112: 405–410. doi:10.1007/s10549-007-9874-3. to resistance to targeted therapies in breast cancer.
Nam, S., S. Park, H.S. Park, S. Kim, J.Y. Kim, and Cell Reports 15: 323–335. doi:10.1016/j.celrep.2016.
S.I. Kim. 2016. Association between insulin resistance 03.026.
and luminal B subtype breast cancer in postmeno- Pathak, D.R., and A.S. Whittemore. 1992. Combined
pausal women. Medicine (Baltimore) 95: e2825. effects of body size, parity, and menstrual events on
doi:10.1097/MD.0000000000002825. breast cancer incidence in seven countries. American
Navarro-Tito, N., A. Soto-Guzman, L. Castro-Sanchez, Journal of Epidemiology 135: 153–168.
R. Martinez-Orozco, and E.P. Salazar. 2010. Oleic Peacock, S.L., E. White, J.R. Daling, L.F. Voigt, and
acid promotes migration on MDA-MB-231 breast K.E. Malone. 1999. Relation between obesity and
cancer cells through an arachidonic acid-dependent breast cancer in young women. American Journal of
pathway. The International Journal of Biochemistry & Epidemiology 149: 339–346.
Cell Biology 42: 306–317. doi:10.1016/j.biocel. Peck, B., Z.T. Schug, Q. Zhang, B. Dankworth, D.T. Jones,
2009.11.010. E. Smethurst, R. Patel, S. Mason, M. Jiang,
Nechuta, S., W.Y. Chen, H. Cai, E.M. Poole, M.L. Kwan, R. Saunders, M. Howell, R. Mitter, B. Spencer-Dene,
S.W. Flatt, R.E. Patterson, J.P. Pierce, B.J. Caan, and G. Stamp, L. McGarry, D. James, E. Shanks,
X. Ou Shu. 2016. A pooled analysis of post-diagnosis E.O. Aboagye, S.E. Critchlow, H.Y. Leung,
lifestyle factors in association with late estrogen-­ A.L. Harris, M.J.O. Wakelam, E. Gottlieb, and
receptor-­positive breast cancer prognosis. A. Schulze. 2016. Inhibition of fatty acid desaturation
International Journal of Cancer 138: 2088–2097. is detrimental to cancer cell survival in metabolically
doi:10.1002/ijc.29940. compromised environments. Cancer & Metabolism 4:
Nelson, E.R., S.E. Wardell, J.S. Jasper, S. Park, 6. doi:10.1186/s40170-016-0146-8.
S. Suchindran, M.K. Howe, N.J. Carver, R.V. Pillai, Peiró, G., E. Adrover, L. Sánchez-Tejada, E. Lerma,
P.M. Sullivan, V. Sondhi, M. Umetani, J. Geradts, and M. Planelles, J. Sánchez-Payá, F.I. Aranda, D. Giner,
D.P. McDonnell. 2013. 27-Hydroxycholesterol links and F.J. Gutiérrez-Aviñó. 2011. Increased insulin-like
hypercholesterolemia and breast cancer pathophysiol- growth factor-1 receptor mRNA expression predicts
ogy. Science 342: 1094–1098. doi:10.1126/ poor survival in immunophenotypes of early breast
science.1241908. carcinoma. Modern Pathology: An Official Journal of
Oh, S.W., C.-Y. Park, E.S. Lee, Y.S. Yoon, E.S. Lee, the United States and Canadian Academy of Pathology,
S.S. Park, Y. Kim, N.J. Sung, Y.H. Yun, K.S. Lee, Inc. 24: 201–208. doi:10.1038/modpathol.2010.191.
H.S. Kang, Y. Kwon, and J. Ro. 2011. Adipokines, Pfeiler, G.H., C. Buechler, M. Neumeier, A. Schäffler,
insulin resistance, metabolic syndrome, and breast G. Schmitz, O. Ortmann, and O. Treeck. 2008.
cancer recurrence: A cohort study. Breast Cancer Adiponectin effects on human breast cancer cells are
Research: BCR 13: R34. doi:10.1186/bcr2856. dependent on 17-beta estradiol. Oncology Reports 19:
Ota, D.M., L.A. Jones, G.L. Jackson, P.M. Jackson, 787–793.
K. Kemp, and D. Bauman. 1986. Obesity, non-protein-­ Pfeiler, G., R. Königsberg, P. Hadji, F. Fitzal, M. Maroske,
bound estradiol levels, and distribution of estradiol in G. Dressel-Ban, J. Zellinger, R. Exner, M. Seifert,
the sera of breast cancer patients. Cancer 57: C. Singer, M. Gnant, and P. Dubsky. 2013. Impact of
558–562. body mass index on estradiol depletion by aromatase
Ozbay, T., and R. Nahta. 2008. A novel unidirectional inhibitors in postmenopausal women with early breast
cross-talk from the insulin-like growth factor-I recep- cancer. British Journal of Cancer 109: 1522–1527.
tor to leptin receptor in human breast cancer cells. doi:10.1038/bjc.2013.499.
Molecular Cancer Research: MCR 6: 1052–1058. Pfeilschifter, J., R. Köditz, M. Pfohl, and H. Schatz. 2002.
doi:10.1158/1541-7786.MCR-07-2126. Changes in proinflammatory cytokine activity after
Pantschenko, A.G., I. Pushkar, K.H. Anderson, Y. Wang, menopause. Endocrine Reviews 23: 90–119.
L.J. Miller, S.H. Kurtzman, G. Barrows, and doi:10.1210/edrv.23.1.0456.
D.L. Kreutzer. 2003. The interleukin-1 family of cyto- Phoenix, K.N., F. Vumbaca, M.M. Fox, R. Evans, and
kines and receptors in human breast cancer: K.P. Claffey. 2010. Dietary energy availability affects
Implications for tumor progression. International primary and metastatic breast cancer and metformin
Journal of Oncology 23: 269–284. efficacy. Breast Cancer Research and Treatment 123:
Papanikolaou, V., N. Stefanou, S. Dubos, I. Papathanasiou, 333–344. doi:10.1007/s10549-009-0647-z.
M. Palianopoulou, V. Valiakou, and A. Tsezou. 2015. Picon-Ruiz, M., C. Pan, K. Drews-Elger, K. Jang,
Synergy of leptin/STAT3 with HER2 receptor induces A.H. Besser, D. Zhao, C. Morata-Tarifa, M. Kim,
tamoxifen resistance in breast cancer cells through T.A. Ince, D.J. Azzam, S.A. Wander, B. Wang,
regulation of apoptosis-related genes. Cellular B. Ergonul, R.H. Datar, R.J. Cote, G.A. Howard,
Oncology (Dordrecht) 38: 155–164. doi:10.1007/ D. El-Ashry, P. Torné-Poyatos, J.A. Marchal, and
s13402-014-0213-5. J.M. Slingerland. 2016. Interactions between adipo-
Park, S., C.-Y. Chang, R. Safi, X. Liu, R. Baldi, J.S. Jasper, cytes and breast cancer cells stimulate cytokine pro-
G.R. Anderson, T. Liu, J.C. Rathmell, M.W. Dewhirst, duction and drive Src/Sox2/miR-302b-mediated
25 Obesity-associated Breast Cancer: Analysis of risk factors 603

malignant progression. Cancer Research 76: 491–504. Tumour Biology: The Journal of the International
doi:10.1158/0008-5472.CAN-15-0927. Society for Oncodevelopmental Biology and Medicine
Pierobon, M., and C.L. Frankenfeld. 2013. Obesity as a 36: 6813–6821. doi:10.1007/s13277-015-3375-5.
risk factor for triple-negative breast cancers: A sys- Railo, M.J., K. von Smitten, and F. Pekonen. 1994. The
tematic review and meta-analysis. Breast Cancer prognostic value of insulin-like growth factor-I in
Research and Treatment 137: 307–314. doi:10.1007/ breast cancer patients. Results of a follow-up study on
s10549-012-2339-3. 126 patients. European Journal of Cancer (Oxford,
Playdon, M.C., M.B. Bracken, T.B. Sanft, J.A. Ligibel, England: 1990) 30A: 307–311.
M. Harrigan, and M.L. Irwin. 2015. Weight gain after Reed, M.J., L. Topping, N.G. Coldham, A. Purohit,
breast cancer diagnosis and all-cause mortality: M.W. Ghilchik, and V.H. James. 1993. Control of aro-
Systematic review and meta-analysis. Journal of the matase activity in breast cancer cells: The role of cyto-
National Cancer Institute 107: djv275. doi:10.1093/ kines and growth factors. The Journal of Steroid
jnci/djv275. Biochemistry and Molecular Biology 44: 589–596.
Pollak, M.N. 1998. Endocrine effects of IGF-I on normal Renehan, A.G., J. Frystyk, and A. Flyvbjerg. 2006.
and transformed breast epithelial cells: Potential rele- Obesity and cancer risk: The role of the insulin-IGF
vance to strategies for breast cancer treatment and pre- axis. Trends in Endocrinology and Metabolism: TEM
vention. Breast Cancer Research and Treatment 47: 17: 328–336. doi:10.1016/j.tem.2006.08.006.
209–217. Renehan, A.G., M. Tyson, M. Egger, R.F. Heller, and
Pollak, M. 2008. Insulin and insulin-like growth factor M. Zwahlen. 2008. Body-mass index and incidence of
signalling in neoplasia. Nature Reviews. Cancer 8: cancer: A systematic review and meta-analysis of
915–928. doi:10.1038/nrc2536. prospective observational studies. Lancet (London,
Pollenz, R.S. 2002. The mechanism of AH receptor pro- England) 371: 569–578. doi:10.1016/S0140-
tein down-regulation (degradation) and its impact on 6736(08)60269-X.
AH receptor-mediated gene regulation. Chemico-­ Révillion, F., M. Charlier, V. Lhotellier, L. Hornez,
Biological Interactions 141: 41–61. S. Giard, M.-C. Baranzelli, J. Djiane, and J.-P. Peyrat.
Porta, R., A. Blancafort, G. Casòliva, M. Casas, J. Dorca, 2006. Messenger RNA expression of leptin and leptin
M. Buxo, G. Viñas, G. Oliveras, and T. Puig. 2014. receptors and their prognostic value in 322 human pri-
Fatty acid synthase expression is strongly related to mary breast cancers. Clinical Cancer Research 12:
menopause in early-stage breast cancer patients. 2088–2094. doi:10.1158/1078-0432.CCR-05-1904.
Menopause (New York, N.Y.) 21: 188–191. Richards, J.A., T.A. Petrel, and R.W. Brueggemeier. 2002.
doi:10.1097/GME.0b013e31829d17dc. Signaling pathways regulating aromatase and cyclo-
Prosperi, J.R., and F.M. Robertson. 2006. oxygenases in normal and malignant breast cells. The
Cyclooxygenase-2 directly regulates gene expression Journal of Steroid Biochemistry and Molecular
of P450 Cyp19 aromatase promoter regions pII, pI.3 Biology 80: 203–212.
and pI.7 and estradiol production in human breast Robinson, P.J., R.J. Bell, and S.R. Davis. 2014. Obesity is
tumor cells. Prostaglandins & Other Lipid Mediators associated with a poorer prognosis in women with hor-
81: 55–70. doi:10.1016/j.prostaglandins.2006.07.003. mone receptor positive breast cancer. Maturitas 79:
Protani, M., M. Coory, and J.H. Martin. 2010. Effect of 279–286. doi:10.1016/j.maturitas.2014.07.004.
obesity on survival of women with breast cancer: Rock, C.L., S.W. Flatt, G.A. Laughlin, E.B. Gold,
Systematic review and meta-analysis. Breast Cancer C.A. Thomson, L. Natarajan, L.A. Jones, B.J. Caan,
Research and Treatment 123: 627–635. doi:10.1007/ M.L. Stefanick, R.A. Hajek, W.K. Al-Delaimy,
s10549-010-0990-0. F.Z. Stanczyk, J.P. Pierce, and Women’s Healthy
Puig, T., A. Vázquez-Martín, J. Relat, J. Pétriz, Eating and Living Study Group. 2008. Reproductive
J.A. Menéndez, R. Porta, G. Casals, P.F. Marrero, steroid hormones and recurrence-free survival in
D. Haro, J. Brunet, and R. Colomer. 2008. Fatty acid women with a history of breast cancer. Cancer
metabolism in breast cancer cells: Differential inhibi- Epidemiology, Biomarkers & Prevention 17: 614–620.
tory effects of epigallocatechin gallate (EGCG) and doi:10.1158/1055-9965.EPI-07-0761.
C75. Breast Cancer Research and Treatment 109: Rock, C.L., C. Doyle, W. Demark-Wahnefried,
471–479. doi:10.1007/s10549-007-9678-5. J. Meyerhardt, K.S. Courneya, A.L. Schwartz,
Purohit, A., D.Y. Wang, M.W. Ghilchik, and M.J. Reed. E.V. Bandera, K.K. Hamilton, B. Grant,
1996. Regulation of aromatase and sulphatase in M. McCullough, T. Byers, and T. Gansler. 2012.
breast tumour cells. The Journal of Endocrinology Nutrition and physical activity guidelines for cancer
150(Suppl): S65–S71. survivors. CA: A Cancer Journal for Clinicians 62:
Purohit, A., S.P. Newman, and M.J. Reed. 2002. The role 243–274. doi:10.3322/caac.21142.
of cytokines in regulating estrogen synthesis: Romero-Figueroa, M.S., J.J. Garduño-García, J. Duarte-­
Implications for the etiology of breast cancer. Breast Mote, G. Matute-González, A. Gómez-Villanueva,
Cancer Research: BCR 4: 65–69. and J. De la Cruz-Vargas. 2013. Insulin and leptin lev-
Qian, Y., D. Shi, J. Qiu, F. Zhu, J. Qian, S. He, Y. Shu, els in obese patients with and without breast cancer.
Y. Yin, and X. Chen. 2015. ObRb downregulation Clinical Breast Cancer 13: 482–485. doi:10.1016/j.
increases breast cancer cell sensitivity to tamoxifen. clbc.2013.08.001.
604 A. Engin

Rose, D.P., and J.M. Connolly. 1999. Omega-3 fatty acids Oncology (Dordrecht) 36: 65–77. doi:10.1007/
as cancer chemopreventive agents. Pharmacology & s13402-012-0114-4.
Therapeutics 83: 217–244. Shi, H., M.V. Kokoeva, K. Inouye, I. Tzameli, H. Yin, and
Rose, D.P., and L. Vona-Davis. 2014. Biochemical and J.S. Flier. 2006. TLR4 links innate immunity and fatty
molecular mechanisms for the association between acid-induced insulin resistance. The Journal of
obesity, chronic inflammation, and breast cancer. Clinical Investigation 116: 3015–3025. doi:10.1172/
BioFactors (Oxford, England) 40: 1–12. doi:10.1002/ JCI28898.
biof.1109. Shimizu, C., T. Hasegawa, Y. Tani, F. Takahashi,
Rossouw, J.E., G.L. Anderson, R.L. Prentice, A.Z. LaCroix, M. Takeuchi, T. Watanabe, M. Ando, N. Katsumata,
C. Kooperberg, M.L. Stefanick, R.D. Jackson, and Y. Fujiwara. 2004. Expression of insulin-like
S.A.A. Beresford, B.V. Howard, K.C. Johnson, growth factor 1 receptor in primary breast cancer:
J.M. Kotchen, J. Ockene, and Writing Group for the Immunohistochemical analysis. Human Pathology 35:
Women’s Health Initiative Investigators. 2002. Risks 1537–1542.
and benefits of estrogen plus progestin in healthy post- Simone, V., M. D’Avenia, A. Argentiero, C. Felici,
menopausal women: Principal results from the Women’s F.M. Rizzo, G. De Pergola, and F. Silvestris. 2016.
Health Initiative randomized controlled trial. Journal of Obesity and breast cancer: Molecular interconnec-
the American Medical Association 288: 321–333. tions and potential clinical applications. The
Salisbury, T.B., G.Z. Morris, J.K. Tomblin, A.R. Chaudhry, Oncologist 21: 404–417. doi:10.1634/
C.R. Cook, and N. Santanam. 2013. Aryl hydrocarbon theoncologist.2015-0351.
receptor ligands inhibit IGF-II and adipokine stimulated Simpson, E.R., G.E. Ackerman, M.E. Smith, and
breast cancer cell proliferation. ISRN Endocrinology C.R. Mendelson. 1981. Estrogen formation in stromal
2013: 104850. doi:10.1155/2013/104850. cells of adipose tissue of women: Induction by gluco-
Samarajeewa, N.U., F. Yang, M.M. Docanto, M. Sakurai, corticosteroids. Proceedings of the National Academy
K.M. McNamara, H. Sasano, S.B. Fox, E.R. Simpson, of Sciences of the United States of America 78:
and K.A. Brown. 2013. HIF-1α stimulates aromatase 5690–5694.
expression driven by prostaglandin E2 in breast adi- Slattery, M.L., K. Curtin, C. Sweeney, R.K. Wolff,
pose stroma. Breast Cancer Research: BCR 15: R30. R.N. Baumgartner, K.B. Baumgartner, A.R. Giuliano,
doi:10.1186/bcr3410. and T. Byers. 2008. Modifying effects of IL-6 poly-
Santen, R., E. Cavalieri, E. Rogan, J. Russo, J. Guttenplan, morphisms on body size-associated breast cancer risk.
J. Ingle, and W. Yue. 2009. Estrogen mediation of Obesity (Silver Spring, Md.) 16: 339–347. ­doi:10.1038/
breast tumor formation involves estrogen receptor-­ oby.2007.44.
dependent, as well as independent, genotoxic effects. Soto-Guzman, A., N. Navarro-Tito, L. Castro-Sanchez,
Annals of the New York Academy of Sciences 1155: R. Martinez-Orozco, and E.P. Salazar. 2010. Oleic
132–140. doi:10.1111/j.1749-6632.2008.03685.x. acid promotes MMP-9 secretion and invasion in breast
Sasano, H., and N. Harada. 1998. Intratumoral aromatase cancer cells. Clinical & Experimental Metastasis 27:
in human breast, endometrial, and ovarian malignan- 505–515. doi:10.1007/s10585-010-9340-1.
cies. Endocrine Reviews 19: 593–607. doi:10.1210/ Stebbing, J., A. Sharma, B. North, T.J. Athersuch,
edrv.19.5.0342. A. Zebrowski, D. Pchejetski, R.C. Coombes,
Sasano, H., T. Suzuki, T. Nakata, and T. Moriya. 2006. J.K. Nicholson, and H.C. Keun. 2012. A metabolic
New development in intracrinology of breast carci- phenotyping approach to understanding relationships
noma. Breast Cancer (Tokyo, Japan) 13: 129–136. between metabolic syndrome and breast tumour
Sasser, A.K., N.J. Sullivan, A.W. Studebaker, L.F. Hendey, responses to chemotherapy. Annals of Oncology 23:
A.E. Axel, and B.M. Hall. 2007. Interleukin-6 is a 860–866. doi:10.1093/annonc/mdr347.
potent growth factor for ER-alpha-positive human Subbaramaiah, K., L.R. Howe, P. Bhardwaj, B. Du,
breast cancer. FASEB Journal 21: 3763–3770. C. Gravaghi, R.K. Yantiss, X.K. Zhou, V.A. Blaho,
doi:10.1096/fj.07-8832com. T. Hla, P. Yang, L. Kopelovich, C.A. Hudis, and
Schapira, D.V., N.B. Kumar, and G.H. Lyman. 1991. A.J. Dannenberg. 2011. Obesity is associated with
Obesity, body fat distribution, and sex hormones in inflammation and elevated aromatase expression in the
breast cancer patients. Cancer 67: 2215–2218. mouse mammary gland. Cancer Prevention Research
Scott, M.M., J.L. Lachey, S.M. Sternson, C.E. Lee, (Philadelphia, Pa.) 4: 329–346. doi:10.1158/1940-
C.F. Elias, J.M. Friedman, and J.K. Elmquist. 2009. ­6207.CAPR-10-0381.
Leptin targets in the mouse brain. The Journal of Subbaramaiah, K., P.G. Morris, X.K. Zhou, M. Morrow,
Comparative Neurology 514: 518–532. doi:10.1002/ B. Du, D. Giri, L. Kopelovich, C.A. Hudis, and
cne.22025. A.J. Dannenberg. 2012. Increased levels of COX-2
Serna-Marquez, N., S. Villegas-Comonfort, O. Galindo-­ and prostaglandin E2 contribute to elevated aromatase
Hernandez, N. Navarro-Tito, A. Millan, and expression in inflamed breast tissue of obese women.
E.P. Salazar. 2013. Role of LOXs and COX-2 on FAK Cancer Discovery 2: 356–365. doi:10.1158/2159-
activation and cell migration induced by linoleic acid ­8290.CD-11-0241.
in MDA-MB-231 breast cancer cells. Cellular
25 Obesity-associated Breast Cancer: Analysis of risk factors 605

Subbaramaiah, K., E. Sue, P. Bhardwaj, B. Du, C.A. Hudis, D. Carter, R. Baserga, and P.M. Glazer. 1997. Insulin-­
D. Giri, L. Kopelovich, X.K. Zhou, and like growth factor-I receptor overexpression mediates
A.J. Dannenberg. 2013. Dietary polyphenols suppress cellular radioresistance and local breast cancer recur-
elevated levels of proinflammatory mediators and aro- rence after lumpectomy and radiation. Cancer
matase in the mammary gland of obese mice. Cancer Research 57: 3079–3083.
Prevention Research (Philadelphia, Pa.) 6: 886–897. Tworoger, S.S., A.H. Eliassen, T. Kelesidis, G.A. Colditz,
doi:10.1158/1940-6207.CAPR-13-0140. W.C. Willett, C.S. Mantzoros, and S.E. Hankinson.
Sulkowska, M., J. Golaszewska, A. Wincewicz, M. Koda, 2007. Plasma adiponectin concentrations and risk of
M. Baltaziak, and S. Sulkowski. 2006. Leptin—From incident breast cancer. The Journal of Clinical
regulation of fat metabolism to stimulation of breast Endocrinology and Metabolism 92: 1510–1516.
cancer growth. Pathology & Oncology Research: POR doi:10.1210/jc.2006-1975.
12: 69–72. doi:PAOR.2006.12.2.0069. Uysal, K.T., S.M. Wiesbrock, M.W. Marino, and
Tan, J., E. Buache, M.-P. Chenard, N. Dali-Youcef, and G.S. Hotamisligil. 1997. Protection from obesity-­
M.-C. Rio. 2011. Adipocyte is a non-trivial, dynamic induced insulin resistance in mice lacking TNF-alpha
partner of breast cancer cells. The International function. Nature 389: 610–614. doi:10.1038/39335.
Journal of Developmental Biology 55: 851–859. Valle, A., J. Sastre-Serra, J. Oliver, and P. Roca. 2011.
doi:10.1387/ijdb.113365jt. Chronic leptin treatment sensitizes MCF-7 breast can-
Tan, A., Y. Dang, G. Chen, and Z. Mo. 2015. cer cells to estrogen. Cellular Physiology and
Overexpression of the fat mass and obesity associated Biochemistry. International Journal of Experimental
gene (FTO) in breast cancer and its clinical implica- Cellular Physiology, Biochemistry and Pharmacology
tions. International Journal of Clinical and 28: 823–832. doi:10.1159/000335796.
Experimental Pathology 8: 13405–13410. van Kruijsdijk, R.C.M., E. van der Wall, and
Thivat, E., S. Thérondel, O. Lapirot, C. Abrial, F.L.J. Visseren. 2009. Obesity and cancer: The role of
P. Gimbergues, E. Gadéa, E. Planchat, F. Kwiatkowski, dysfunctional adipose tissue. Cancer Epidemiology,
M.A. Mouret-Reynier, P. Chollet, and X. Durando. Biomarkers & Prevention 18: 2569–2578.
2010. Weight change during chemotherapy changes doi:10.1158/1055-9965.EPI-09-0372.
the prognosis in non metastatic breast cancer for the Vandeweyer, E., and D. Hertens. 2002. Quantification of
worse. BMC Cancer 10: 648. doi:10.1186/ glands and fat in breast tissue: An experimental deter-
1471-2407-10-648. mination. Annals of Anatomy 184: 181–184.
Toft, D.J., and V.L. Cryns. 2011. Minireview: Basal-like doi:10.1016/S0940-9602(02)80016-4.
breast cancer: From molecular profiles to targeted Vedin, L.-L., S.A. Lewandowski, P. Parini,
therapies. Molecular Endocrinology (Baltimore, Md.) J.-A. Gustafsson, and K.R. Steffensen. 2009. The oxy-
25: 199–211. doi:10.1210/me.2010-0164. sterol receptor LXR inhibits proliferation of human
Tomblin, J.K., and T.B. Salisbury. 2014. Insulin like breast cancer cells. Carcinogenesis 30: 575–579.
growth factor 2 regulation of aryl hydrocarbon recep- doi:10.1093/carcin/bgp029.
tor in MCF-7 breast cancer cells. Biochemical and Vigushin, D.M., Y. Dong, L. Inman, N. Peyvandi,
Biophysical Research Communications 443: 1092– J.P. Alao, C. Sun, S. Ali, E.J. Niesor, C.L. Bentzen,
1096. doi:10.1016/j.bbrc.2013.12.112. and R.C. Coombes. 2004. The nuclear oxysterol
Toniolo, P.G., M. Levitz, A. Zeleniuch-Jacquotte, receptor LXRalpha is expressed in the normal human
S. Banerjee, K.L. Koenig, R.E. Shore, P. Strax, and breast and in breast cancer. Medical Oncology
B.S. Pasternack. 1995. A prospective study of endog- (Northwood, London, England) 21: 123–131.
enous estrogens and breast cancer in postmenopausal doi:10.1385/MO:21:2:123.
women. Journal of the National Cancer Institute 87: Vona-Davis, L., and D.P. Rose. 2013. The obesity-­
190–197. inflammation-­eicosanoid axis in breast cancer. Journal
Toropainen, E.M., P.K. Lipponen, and K.J. Syrjanen. of Mammary Gland Biology and Neoplasia 18: 291–
1995. Expression of insulin-like growth factor II in 307. doi:10.1007/s10911-013-9299-z.
female breast cancer as related to established prognos- Vozarova, B., C. Weyer, K. Hanson, P.A. Tataranni,
tic factors and long-term prognosis. Anticancer C. Bogardus, and R.E. Pratley. 2001. Circulating inter-
Research 15: 2669–2674. leukin-­6 in relation to adiposity, insulin action, and
Tsuchida, A., T. Yamauchi, Y. Ito, Y. Hada, T. Maki, insulin secretion. Obesity Research 9: 414–417.
S. Takekawa, J. Kamon, M. Kobayashi, R. Suzuki, doi:10.1038/oby.2001.54.
K. Hara, N. Kubota, Y. Terauchi, P. Froguel, J. Nakae, Vrieling, A., K. Buck, R. Kaaks, and J. Chang-Claude.
M. Kasuga, D. Accili, K. Tobe, K. Ueki, R. Nagai, and 2010. Adult weight gain in relation to breast cancer
T. Kadowaki. 2004. Insulin/Foxo1 pathway regulates risk by estrogen and progesterone receptor status: A
expression levels of adiponectin receptors and adipo- meta-analysis. Breast Cancer Research and Treatment
nectin sensitivity. The Journal of Biological Chemistry 123: 641–649. doi:10.1007/s10549-010-1116-4.
279: 30817–30822. doi:10.1074/jbc.M402367200. Wang, Z., X. Zhang, P. Shen, B.W. Loggie, Y. Chang, and
Turner, B.C., B.G. Haffty, L. Narayanan, J. Yuan, T.F. Deuel. 2005. Identification, cloning, and expres-
P.A. Havre, A.A. Gumbs, L. Kaplan, J.L. Burgaud, sion of human estrogen receptor-alpha36, a novel vari-
606 A. Engin

ant of human estrogen receptor-alpha66. Biochemical ponectin levels and the risk of breast cancer: A meta-­
and Biophysical Research Communications 336: analysis. European Journal of Cancer Prevention 23:
1023–1027. doi:10.1016/j.bbrc.2005.08.226. 158–165. doi:10.1097/CEJ.0b013e328364f293.
Wang, J., D.-L. Yang, Z.-Z. Chen, and B.-F. Gou. 2016. Yerushalmi, R., K.A. Gelmon, S. Leung, D. Gao,
Associations of body mass index with cancer inci- M. Cheang, M. Pollak, G. Turashvili, B.C. Gilks, and
dence among populations, genders, and menopausal H. Kennecke. 2012. Insulin-like growth factor recep-
status: A systematic review and meta-analysis. Cancer tor (IGF-1R) in breast cancer subtypes. Breast Cancer
Epidemiology 42: 1–8. doi:10.1016/j.canep.2016. Research and Treatment 132: 131–142. doi:10.1007/
02.010. s10549-011-1529-8.
Weisberg, S.P., D. McCann, M. Desai, M. Rosenbaum, Yonezawa, T., K. Katoh, and Y. Obara. 2004. Existence of
R.L. Leibel, and A.W. Ferrante. 2003. Obesity is asso- GPR40 functioning in a human breast cancer cell line,
ciated with macrophage accumulation in adipose tissue. MCF-7. Biochemical and Biophysical Research
The Journal of Clinical Investigation 112: 1796–1808. Communications 314: 805–809.
doi:10.1172/JCI19246. Yoon, J.-H., J.K. Park, S.S. Oh, K.-H. Lee, S.-K. Kim,
Wolowczuk, I., C. Verwaerde, O. Viltart, A. Delanoye, I.-J. Cho, J.-K. Kim, H.-T. Kang, S.G. Ahn, J.-W. Lee,
M. Delacre, B. Pot, and C. Grangette. 2008. Feeding S.-H. Lee, A. Eom, J.-Y. Kim, S.V. Ahn, and S.B. Koh.
our immune system: Impact on metabolism. Clinical 2011. The ratio of serum leptin to adiponectin pro-
& Developmental Immunology 2008: 639803. vides adjunctive information to the risk of metabolic
doi:10.1155/2008/639803. syndrome beyond the homeostasis model assessment
Won, H.S., Y.A. Kim, J.S. Lee, E.K. Jeon, H.J. An, insulin resistance: The Korean Genomic Rural Cohort
D.S. Sun, Y.H. Ko, and J.S. Kim. 2013. Soluble inter- Study. Clinica Chimica Acta: International Journal of
leukin-­6 receptor is a prognostic marker for relapse-­ Clinical Chemistry 412: 2199–2205. doi:10.1016/j.
free survival in estrogen receptor-positive breast cca.2011.08.003.
cancer. Cancer Investigation 31: 516–521. doi:10.310 Yue, W., J.-P. Wang, Y. Li, W.P. Bocchinfuso, K.S. Korach,
9/07357907.2013.826239. P.D. Devanesan, E. Rogan, E. Cavalieri, and
Wu, Q., T. Ishikawa, R. Sirianni, H. Tang, J.G. McDonald, R.J. Santen. 2005. Tamoxifen versus aromatase inhibi-
I.S. Yuhanna, B. Thompson, L. Girard, C. Mineo, tors for breast cancer prevention. Clinical Cancer
R.A. Brekken, M. Umetani, D.M. Euhus, Y. Xie, and Research 11: 925s–930s.
P.W. Shaul. 2013. 27-Hydroxycholesterol promotes Zhang, G.J., and I. Adachi. 1999. Serum interleukin-6 lev-
cell-autonomous, ER-positive breast cancer growth. els correlate to tumor progression and prognosis in
Cell Reports 5: 637–645. doi:10.1016/j. metastatic breast carcinoma. Anticancer Research 19:
celrep.2013.10.006. 1427–1432.
Yao-Borengasser, A., B. Monzavi-Karbassi, R.A. Hedges, Zhang, X.-T., L. Ding, L.-G. Kang, and Z.-Y. Wang. 2012.
L.J. Rogers, S.A. Kadlubar, and T. Kieber-Emmons. Involvement of ER-α36, Src, EGFR and STAT5 in the
2015. Adipocyte hypoxia promotes epithelial-­ biphasic estrogen signaling of ER-negative breast can-
mesenchymal transition-related gene expression and cer cells. Oncology Reports 27: 2057–2065.
estrogen receptor-negative phenotype in breast cancer doi:10.3892/or.2012.1722.
cells. Oncology Reports 33: 2689–2694. doi:10.3892/ Zhou, W., S. Guo, and R.R. Gonzalez-Perez. 2011. Leptin
or.2015.3880. pro-angiogenic signature in breast cancer is linked to
Ye, J., J. Jia, S. Dong, C. Zhang, S. Yu, L. Li, C. Mao, IL-1 signalling. British Journal of Cancer 104: 128–
D. Wang, J. Chen, and G. Yuan. 2014. Circulating adi- 137. doi:10.1038/sj.bjc.6606013.

You might also like