当前位置: 首页 > 期刊 > 《临床肿瘤学》 > 2005年第10期 > 正文
编号:11332200
Biology of Progesterone Receptor Loss in Breast Cancer and Its Implications for Endocrine Therapy
http://www.100md.com 《临床肿瘤学》
     the Breast Center, Baylor College of Medicine

    the Methodist Hospital, Houston, TX

    ABSTRACT

    The response to endocrine therapy in breast cancer correlates with estrogen receptor (ER) and progesterone receptor (PR) status. ER-positive/PR-negative breast cancers respond less well to selective ER modulator (SERM) therapy than ER-positive/PR-positive tumors. The predictive value of PR has long been attributed to the dependence of PR expression on ER activity, with the absence of PR reflecting a nonfunctional ER and resistance to hormonal therapy. However, recent clinical and laboratory evidence suggests that ER-positive/PR-negative breast cancers may be specifically resistant to SERMs, whereas they may be less resistant to estrogen withdrawal therapy with aromatase inhibitors, which is a result inconsistent with the nonfunctional ER theory. Novel alternative molecular mechanisms potentially explaining SERM resistance in ER-positive/PR-negative tumors have been suggested by recent experimental indications that growth factors may downregulate PR levels. Thus, the absence of PR may not simply indicate a lack of ER activity, but rather may reflect hyperactive cross talk between ER and growth factor signaling pathways that downregulate PR even as they activate other ER functions. Therefore, ER-positive/PR-negative breast tumors might best be treated by completely blocking ER action via estrogen withdrawal with aromatase inhibitors, by targeted ER degradation, or by combined therapy targeting both ER and growth factor signaling pathways. In this review, we will discuss the biology and etiology of ER-positive/PR-negative breast cancer, highlighting recent data on molecular cross talk between ER and growth factor signaling pathways and demonstrating how PR might be a useful marker of these activities. Finally, we will consider the clinical implications of these observations.

    INTRODUCTION

    Estrogen and the estrogen receptor (ER) play key roles in both normal breast development and breast cancer progression. Therapeutic strategies directed at inhibiting the action of ER using selective ER modulators (SERMs), targeting ER for degradation with selective ER downregulators (SERDs), or withdrawing estrogen by surgical (oophorectomy) or medical (luteinizing hormone agonists and antagonists) ovarian ablation or by aromatase inhibitors represent highly successful examples of targeted therapy for clinical breast cancer.1-4

    In this review, we will highlight the predictive value of ER and progesterone receptor (PR) in breast cancer and provide hypotheses to explain why a significant subset of breast cancers (ER-positive/PR-negative) may be selectively more resistant to SERM therapy. The difference in clinical benefit from SERMs versus aromatase inhibitors in ER-positive/PR-negative tumors may be explained by new insights into the regulatory mechanisms of ER and PR that could simultaneously cause the ER-positive/PR-negative phenotype and the resistance to SERM therapy. These mechanisms invoke molecular cross talk between ER, PR, and growth factor–receptor signaling pathways and the relationship between the classical and nonclassical effects of ER in breast cancer cells. Here, we summarize the molecular signals regulating PR levels in breast cancer and highlight recent clinical data that support a strong rationale for the design and implementation of clinical studies using alternative endocrine therapies in ER-positive/PR-negative tumors.

    ER AND PR IN BREAST CANCER

    Expression and Prognostic Value of ER and PR

    The importance of steroid hormones in breast cancer has been known for decades, and many studies have described the ontogeny of steroid receptor expression in both rodent model systems and the human mammary gland.5,6 The ER exists as two isoforms, ER and ER?, which are encoded by two different genes (Fig 1). Like other steroid receptors, the ER has a structure consisting of a DNA-binding domain (DBD) flanked by two transcriptional activation domains (AF-1 and AF-2). The receptor binds its ligand estradiol in the ligand-binding domain (LBD). Studies in rodents have shown that ER and ER? are expressed in the normal mammary gland7 and that expression of ER, but not ER?, is critical for normal mammary gland ductal development.8 In humans, ER is also expressed in the normal breast, and a dramatic increase in ER expression is seen in early hyperproliferative premalignant lesions.6

    One of the most studied ER-regulated genes is PR, which mediates progesterone's effects in the development of the mammary gland and breast cancer.9 PR is expressed as two isoforms (PR-A and PR-B) from a single gene (Fig 1).10 Like ER, PR contains a DBD, LBD, and multiple AFs.11 Studies in the rodent mammary gland have shown that PR is critical for lobuloalveolar development of the gland12 and that the ratio of PR-A to PR-B is critical for proper mammary gland development.13 Interestingly, an overabundance of PR-A in human breast cancers has recently been reported to be associated with resistance to tamoxifen,14 whereas a functional promoter polymorphism that results in increased production of PR-B is associated with an increased risk of breast cancer.15 The dramatic increase in breast cancer incidence in women taking both estrogen and progesterone for hormone replacement therapy, compared with estrogen alone, emphasizes the importance of progesterone and the PR in breast cancer.16

    Approximately 75% of primary breast cancers express ER, and more than half of these cancers also express PR.17 Both ER and PR are prognostic factors, although both are weak and lose their prognostic value after long-term follow-up.18 PR is an estrogen-regulated gene, and its synthesis in normal and cancer cells requires estrogen and ER. Therefore, it is not surprising that ER-positive/PR-positive tumors are more common than ER-positive/PR-negative tumors. The etiology of ER-positive/PR-negative tumors is currently unclear. Some studies have shown that ER and PR status can change over the natural history of the disease or during treatment.19 For instance, sequential breast cancer biopsies have shown that ER levels are reduced slightly with intervening endocrine therapy, although complete loss is uncommon. In contrast, PR levels decrease more dramatically during tamoxifen therapy, with up to half of tumors completely losing PR expression when resistance develops.20 These ER-positive/PR-negative metastatic tumors then display a much more aggressive course after loss of PR compared with tumors retaining PR, and patients then have a worse overall survival, indicating a change in tumor cell–regulatory mechanisms.20,21 Whether and how the loss of PR affects the poor clinical course of these tumors is at present unclear.

    Although ER-positive/PR-negative primary untreated tumors may simply evolve by loss of PR from subclinical ER-positive/PR-negative tumors, the differences in the biology and outcome of ER-positive/PR-negative tumors suggest that some of these tumors may initially evolve separately as ER-positive/PR-negative tumors, representing their own individual stable phenotype from the outset. Indeed, recent prospective studies have shown that ER-positive/PR-negative tumors have their own unique epidemiologic risk factors.22 For instance, the incidence of each of the four receptor tumor subtypes (ER-positive/PR-positive, ER-positive/PR-negative, ER-negative/PR-positive, and ER-negative/PR-negative) differs with age,23 pregnancy history, postmenopausal hormone use, and body mass index after menopause.24 We have also found similar associations24A between ER-positive/PR-negative tumors and older age compared with ER-positive/PR-positive tumors. Importantly, many of these associations remain for PR by itself, even when corrected for quantitative ER levels, highlighting the significance of both receptors not only in breast cancer treatment, but also in their etiology.

    ER and PR Predict Response to Endocrine Therapy

    ER status is a strong predictor of response to endocrine therapy. In the adjuvant setting, a meta-analysis showed that tamoxifen significantly reduces recurrence and death only in patients with ER-positive tumors.25 Similar results have been shown in two trials prospectively designed to test the value of tamoxifen in ER-negative tumors26,27 and also in retrospective analyses of several clinical trials.28 In several studies, it has been shown that response to tamoxifen is directly related to ER levels,29 but there are responses in tumors that have as little as 4 to 10 fmol/mg of ER protein or as few as 1% to 10% of cells positive for ER by immunohistochemistry.30

    Although ER is an accepted predictor of response to endocrine therapy, the role of PR has been more controversial. For instance, the Oxford overview of all trials of tamoxifen therapy in early breast cancer found that PR status did not predict benefit.25 However, variability in assay methodology and lack of quality control in many laboratories raise questions about this data. In contrast to this, we have recently published a retrospective analysis of two of the largest data sets of patients (n = 15,871) with early breast cancer treated with endocrine therapy (nearly all received tamoxifen), with ER and PR levels measured in two standardized quality-controlled clinical laboratories.18 In this study, we found that patients with ER-positive/PR-positive tumors benefited much more from adjuvant tamoxifen therapy than patients with ER-positive/PR-negative tumors (Fig 2). Multivariate analyses showed that both ER and PR were independent predictors of overall survival, with the reduction in relative risk of death being significantly greater in ER-positive/PR-positive compared with ER-positive/PR-negative tumors. Importantly, PR still had predictive value even when ER was considered as a continuous variable, indicating that the predictive information is independent of quantitative ER levels18 and that PR adds predictive information to ER. The predictive value of PR in the adjuvant tamoxifen setting has also been shown in two smaller studies.31,32

    Consistent with the predictive power of PR in the adjuvant setting, several studies have shown that elevated PR levels significantly and independently correlate with increased probability of response to tamoxifen, longer time to treatment failure, and longer overall survival in patients with metastatic disease.33 As with ER, response is directly and positively related to PR levels.28 Similarly, two recent neoadjuvant studies showed a better response (clinical outcome and proliferation rate) to endocrine therapy in PR-positive tumors compared with PR-negative tumors.34,35

    ER-positive/PR-negative Breast Tumors Are Selectively Resistant to SERMs

    Supporting the studies indicating a role for PR in predicting response to adjuvant antihormone therapy are recent data from the Arimidex, Tamoxifen, Alone or in Combination (ATAC) trial, which randomly assigned postmenopausal women with early breast cancer to 5 years of treatment with the aromatase inhibitor anastrazole, tamoxifen, or the combination.36 A recent preliminary analysis of this trial with respect to hormone receptor status showed that, when all treatments are combined together (termed endocrine therapy), 7.6% of patients with ER-positive/PR-positive tumors had a recurrence (breast cancer event) over a 47-month follow-up period, whereas 14.8% of patients with ER-positive/PR-negative tumors experienced recurrence (Fig 3A). Importantly, however, a careful analysis of the individual treatments (anastrazole v tamoxifen v the combination) reveals tantalizing hints as to the biology of this reduced efficacy. The inferior outcome in the ER-positive/PR-negative subgroup was restricted largely to the tamoxifen and combination treatment arms (Fig 3B). In contrast, the anastrazole only arm showed only a trend for anastrazole to be superior to tamoxifen in the ER-positive/PR-positive subset. Indeed, this preliminary hormone receptor data suggest that the overall benefit of anastrazole over tamoxifen in the ATAC trial is largely a result of the reduced efficacy of tamoxifen in the patients with ER-positive/PR-negative tumors.

    Collectively, these data indicate that ER-positive/PR-negative breast tumors are less responsive to SERM therapy compared with ER-positive/PR-positive tumors. It is noteworthy that in the 1970s it was hypothesized that PR might provide additional information to more accurately predict which patients will respond to hormonal therapy.38 This theory was based on the rationale that PR is induced by estrogen in ER-positive breast cancer cell lines and, therefore, that PR would serve as an indicator of a functionally intact ER pathway. However, this simple theory fails to fully explain differences such as those seen between anastrazole and tamoxifen in the ATAC trial, which presumably are mediated via the ER.

    Increased Growth Factor Signaling Is Associated Both With the ER-Positive/PR-Negative Phenotype and With SERM Resistance

    Several reports suggest that high growth factor activity in breast cancers may be associated with decreased PR levels. Konecny et al39 recently reported that patients with higher levels of human epidermal growth factor receptor 2 (HER2) in their tumors had statistically significantly lower levels of ER/PR than patients with lower levels of HER2. Intriguingly, their study also revealed that relatively low levels of HER2 amplification/overexpression were associated with marked decreases of PR but not ER, thus causing tumors to be ER-positive/PR-negative. Recently, Dowsett et al35 found that 25% of ER-positive/PR-negative tumors overexpressed HER2 compared with 10% of ER-positive/PR-positive tumors. Several other studies also suggest that ER-positive/PR-negative tumors are more likely to be higher grade and to be amplified for HER2 than ER-positive/PR-positive tumors.21,40 We have also recently found that ER-positive/PR-negative tumors have higher levels of epidermal growth factor receptor (EGFR) and HER2 compared with ER-positive/PR-positive tumors (Arpino et al24A).

    High HER2 signaling has been associated with reduced tamoxifen efficacy in many, although not all, clinical studies.41 However, this may be in part a result of the fact that tumors that overexpress HER2 are more likely to be ER negative and PR negative. In fact, when analyzed as a continuous variable, HER2 is inversely related to ER and PR even in the subset of ER-positive patients.39 Despite this, Dowsett et al35 showed that tamoxifen resistance in ER-positive/HER2-positive tumors was unrelated to ER levels, and Arpino et al24A found that EGFR levels predicted resistance to tamoxifen in multivariate analysis considering ER levels as a continuous variable.42 Therefore these studies suggest that reduced ER expression in EGFR- or HER2-positive tumors may not fully account for the tamoxifen resistance. Other studies suggest that HER2 status alone is not a strong predictor of tamoxifen resistance in ER-positive tumors, although studies like this in metastatic breast cancer rely on the hormone receptor status of the primary tumor and not the metastasis.43 Levels of ER, PR, or HER2 can change over time or with treatment as tumors progress to overt metastasis.44,45

    Two recent neoadjuvant studies confirmed the observation that PR-negative tumors respond less well to hormonal therapy than PR-positive tumors.34,35 Furthermore, Ellis et al34 went on to show that tamoxifen was less effective than an aromatase inhibitor in patients whose tumors were EGFR-positive and/or HER2-positive, again reiterating the potent cross talk that can occur between this pathway and ER, perhaps contributing to relative resistance. A recent retrospective study demonstrating a poor disease-free survival for patients receiving adjuvant tamoxifen whose tumors expressed high levels of both HER2 and the ER coactivator AIB146 further implicates growth factor activity interacting with the ER pathway in resistance to SERMs.

    The HER family of receptors is upstream to the PI3K/Akt/mTOR signaling cascade.41 Phosphatase and tensin homolog (PTEN) is a negative regulator of this pathway, and loss of PTEN, which is associated with activation of this survival pathway, has recently been correlated with loss of PR in clinical breast cancer specimens.47 In another smaller study, loss of PTEN correlated with loss of both ER and PR; however, this study did not specifically examine ER-positive/PR-negative tumors.48 Importantly, it has been reported that loss of heterozygosity (LOH) at chromosome 10q23, which harbors the PTEN gene, occurs in approximately 30% to 40% of sporadic breast cancers and that this LOH is associated with higher histologic grade and specific loss of PR but not ER expression.49

    Finally, Akt, a key downstream kinase, not only downregulates PR levels, but it also induces ligand-independent activation of ER.50 ER signaling might then activate other cell survival pathways, thereby protecting breast cancer cells from tamoxifen-induced apoptosis51-53 or reducing tamoxifen's antagonist properties, resulting in tamoxifen-stimulated growth and treatment resistance. Clinical studies have reported that high levels of phosphorylated Akt predict worse outcome among endocrine-treated patients.54

    MOLECULAR BASIS FOR ER-POSITIVE/PR-NEGATIVE BREAST CANCERS AND RESISTANCE TO SERMS

    ER Structure and Function

    Overview. The ER is a member of a large family of nuclear transcriptional regulators.55 Most research on the biology of ER has focused on its function as a nuclear transcription factor; however, ER can also bind to other transcription factors and proteins and function as a coregulator to enhance genes not traditionally thought of as estrogen target genes. Finally, recent evidence also suggests that estrogen can bind ER located in or near the plasma membrane and rapidly activate other signaling pathways. To better define these mechanisms of action, a recent consensus nomenclature was proposed56 whereby estrogen action in the nucleus is termed nuclear-initiated steroid signaling (NISS) and estrogen action at the plasma membrane is termed membrane-initiated steroid signaling (MISS). Both of these modes of ER action will now be briefly discussed.

    NISS. ER is a hormone-regulated nuclear transcription factor that can induce expression of a number of genes (eg, PR).57 On ligand activation, ER binds to estrogen response elements in target genes, recruits a coregulator complex, and regulates transcription of specific genes (Fig 4, pathway 1).58,59 ER has also been shown to act in a nonclassical manner without a need for DNA binding by modulating gene expression at alternative regulatory sequences such as AP-1,60 SP-1,61 and USF sites62 (Fig 4, pathway 2).

    ER action is controlled by a growing number of coregulatory proteins termed coactivators and corepressors that recruit enzymes that modulate chromatin structure to facilitate or repress gene transcription.63,64 It has been postulated that the ability of SERMs to either activate or repress ER action is in part controlled by the cellular environment of coregulators. Indeed, the first experiments to test this showed that altering coregulator levels could change SERMs from antagonists to agonists but that this was cell specific.65 Further studies showed that there were cell type– and promoter-specific differences in coregulator recruitment that determine the cellular response to SERMs.66 Given these observations, studies have been performed that examined coregulator levels in breast cancer and associations with prognosis and hormone resistance,63,67 although more studies are needed, in particular concerning a fundamental understanding of the determinants of coregulator levels.

    ER also interacts with several other nuclear proteins that themselves are critical in breast cancer. For instance, ER can bind BRCA1, and this interaction results in a reduction of ER activity.68 Mutated BRCA1 protein derived from genetic carriers fails to repress ER, providing a hypothesis for the hormone-dependent cancer formation that is suppressed by ovarian ablation in these patients.69 However, once breast tumors appear in BRCA1 carriers, the tumors are invariably ER negative, indicating that, at this stage, BRCA1 has no ER-dependent function. ER also directly interacts with and is activated by cyclin D1.70,71 Cyclin D1 can also interact with two coactivators of the ER, steroid receptor coactivator (SRC)-172 and p300.73 Thus, cyclin D1 could function as a bridge between ER and coactivators to augment transcriptional activation.

    In addition to estrogen, several other stimuli can enhance NISS in a ligand-independent manner (Fig 4, pathway 3). Growth factors, such as insulin-like growth factor-I (IGF-I), epidermal growth factor (EGF), heregulin, and transforming growth factor alpha,74 neurotransmitters such as dopamine,75 signaling molecules such as cyclic adenosine monophosphate,76 and membrane-permeable phosphatase inhibitors77 can all potentiate NISS. One of the mechanisms for ligand-independent activation may be phosphorylation of ER or its coregulators at specific sites induced by growth factors and stress-related kinases.78 ER is phosphorylated at several sites by multiple kinases including the extracellular regulated kinase (ERK) 1/2 and p38 mitogen-activated protein kinases (MAPKs),79,80 cyclin-dependent kinase (CDK) –2,81 CDK-7,82 c-SRC,83 protein kinase A,84 pp90rsk1,85 and AKT. 51 Phosphorylation of ER at serine residues clustered in its amino terminus (Ser 104/106, 118, and 167) enhances transcriptional transactivating activity arising from the ligand-independent AF-1 domain.84,86 ER dimerization and DNA binding are regulated by phosphorylation at serine 236 in the DBD87 and by tyrosine phosphorylation at residue 537 within the LBD, with the latter modification also involved in estrogen binding.88 Finally, phosphorylation on threonine 311, which is induced by estrogen activation of the p38 MAPK, promotes ER nuclear localization via nuclear export regulation and enhances ER interaction with coactivators.89

    However, phosphorylation of ER coregulators is probably as important as phosphorylation of ER itself in communicating growth factor and other signaling effects on the ER pathway.90 Several key signaling kinases, including the mitogenic ERK1/2, the stress-related kinases I kappa B kinase,91 c-Jun NH2-terminal kinase (JNK), and the p38 MAPK,92 have all been suggested to phosphorylate members of the p160 SRC family of coactivators, and especially AIB1, which is often gene amplified or overexpressed in breast tumor cells. Phosphorylation of coactivators can augment their activity on ER-dependent transcription, even in the absence of ligand or in the presence of antiestrogens, by increasing their subcellular nuclear localization,91 their interaction with the ER, and their ability to recruit other transcriptional coregulators, such as the CBP/p300 coactivator, to the receptor-promoter complex.90 In addition, phosphorylation may also directly activate intrinsic enzymatic activities of the coactivators such as acetyltransferase activity.93 A recent identification of the phosphorylation sites on SRC-392 has shown that multiple upstream signaling cascades phosphorylate different combinations of sites and that this may serve as a mechanism for SRC-3 to integrate and coordinate multiple kinase signaling cascades.

    Like ER's coactivators, corepressor action on the ERs is also regulated and potentially negated by multiple signal transduction pathways.53,94 Direct phosphorylation of corepressors or of interacting components in the corepressor complex and consequent changes in their affinity for the cognate transcription factors and in their cellular distribution are mechanisms that might be responsible for the observed inhibition of corepressor action by growth factors and kinases.95,96

    MISS. In addition to ER acting as a nuclear transcription factor, growing evidence showing that estrogen can have rapid cellular effects that occur within minutes, long before its effects on gene transcription, suggests other mechanisms of action.97 ER can be detected in or near the plasma membrane, where it can directly interact with and modulate several signaling pathways. This type of ER signaling is now referred to as MISS (Fig 4, pathway 4).

    Several lines of evidence suggest that ER does function outside the nucleus, perhaps in the membrane or cytoplasm, although the identity of the receptors mediating the MISS actions of steroid hormones still remains controversial. Non-nuclear localization of ER has been shown by biochemical fractionation of the membrane in cells that overexpress ER98 and by direct visualization of membrane ER using immunocytochemistry.99 In addition, an alternative splicing/translational variant of ER100,101 that binds to the membrane may transduce the rapid estrogen signaling in some cells. Despite this evidence, membrane receptors distinct from the classical ERs may also be involved102 because rapid estrogen effects have been documented in several ER-negative cell lines,103 although the relevance of this data to clinical breast cancer is suspect given the lack of effect of endocrine therapy in ER-negative tumors.

    Cell culture studies have shown that ER can interact at many levels of the IGF signaling pathway (Fig 5). It has previously been shown that estrogen can increase expression of IGF-I receptor (IGF-IR)104 and insulin receptor substrate-1 (IRS-1)105 and sensitize cells to IGFs and, conversely, that IGFs can phosphorylate and activate ER in a ligand-independent manner.106 However, recent studies have shown numerous interactions between proximal IGF signaling intermediates and putative membrane or cytoplasmic ER (Fig 5). Estrogen stimulates association between ER and IGF-IR, and this results in activation of IGF-I signaling via ERK1/2.107 Estrogen also stimulates association between ER and the p85 subunit of phosphatidylinositol 3'-kinase (PI3K), resulting in subsequent PI3K activation.108 Both of these interactions are inhibited by antiestrogens and do not occur with ER?. ER has been shown to interact with the IGF-IR downstream signaling intermediates (IRS-1109), and on estrogen stimulation, IRS-1 translocates to the nucleus and can be found bound to ER on the pS2 promoter.110 ER is also able to interact with shc, which seems to be critical for the translocation of ER to the membrane. Importantly, reduction of shc expression inhibits formation of an ER/IGF-IR complex, and reduction of IGF-IR inhibits ER/shc association. Therefore, it is interesting to note that ER can bind all elements of the proximal IGF-IR signaling cascade, establishing a complex that may be activated by estrogen. The exact mechanism of action of this complex and the resulting functional significance are still to be deciphered.

    Membrane ERs have also been shown to predominantly localize to discrete caveolar domains where they physically interact with the scaffold caveolin proteins111 and act as G protein–coupled receptors.112 G protein activation, via c-Src, leads to activation of matrix metalloproteinases, which in turn, cleave and liberate heparin-binding EGF. This free heparin-binding EGF, acting in an autocrine manner, is responsible for the activation and phosphorylation of EGFR and, subsequently, its downstream kinases ERK1/2 and PI3K. This molecular scenario is ER-dependent and is induced by estrogen and SERMs like tamoxifen, although it is blocked by the pure ER antagonist fulvestrant.112 Interestingly, overexpression of the EGFR family member HER2/neu (HER2) has also been shown in experimental systems to potentiate MISS in response to both estrogen and tamoxifen.113,114 It is this ER activity that may be a critical element in blocking tamoxifen-induced apoptosis in HER2-overexpressing breast cancer cells,115 and it may contribute to increased estrogen activity of this SERM and then to tamoxifen-stimulated growth as a mechanism of resistance.113

    A novel ER-interacting protein called MNAR/PELP1 (standing for modulator of nongenomic activity of estrogen receptor116), which modulates both MISS and NISS activities of ER,117,118 was recently cloned. Similarly, the metastasis-associated genes (MTA) family of ER coregulators may also regulate MISS and NISS. MTA1 is a corepressor of genomic ER activity via traditional recruitment of histone deacetylases to ER transcriptional complexes.119 A naturally occurring variant, MTA1s, not only downregulates nuclear ER activity but, by sequestering ER in the cytoplasm, also simultaneously increases ER cytoplasmic or membrane signaling.114 Interestingly, both forms of MTA1 have been recently identified as targets of growth factor signal transduction.120

    PR Is a Marker of a Functional ER

    There are many theories to explain the generation of ER-positive/PR-negative breast tumors (Table 1). The simplest is that the ER is nonfunctional and unable to stimulate PR production and that the tumor is, therefore, no longer dependent on estrogen for growth and survival. However, some ER-positive/PR-negative breast cancers may simply result from low circulating levels of endogenous estrogens in postmenopausal women that are insufficient to induce PR expression even though the ER pathway is intact. Thus, brief treatment of patients with ER-positive/PR-negative tumors with estrogen can restore PR levels in some of the patients.121 Additionally, reduced response to tamoxifen may be a result of lower ER levels in ER-positive/PR-negative tumors because response has been shown to be directly related to ER levels.28 However, as a group, patients with ER-positive/PR-negative tumors responded significantly better to estrogen withdrawal than to tamoxifen in the ATAC trial (Fig 3B), indicating that these tumors are also still estrogen dependent and that simple loss of ER function via low levels of estrogen or ER is not the whole explanation for resistance to tamoxifen.

    Indeed, an increasing literature suggests that the nonfunctional ER theory is too simple on its own to completely explain the loss of PR in ER-positive breast cancer and that there must be multiple molecular mechanisms for this phenotype. For instance, a recent study of the transcriptional activity of ER in fresh breast tumor lysates showed that ER is transcriptionally active in some ER-positive/PR-negative breast tumors.125 Furthermore, no mutations have been found in the ER DBD of a series of ER-positive/PR-negative tumors.126,127 A subset of ER-positive/PR-negative breast tumors do show an inability of the ER to bind DNA, and this inability to bind DNA may be related to oxidative stress within the tumor.128 Although it is possible that there are mutations in other regions of ER or in ER coregulators important for transcriptional activation that could cause loss of ER function, studies to date have failed to find any that can fully explain loss of PR expression in the tumor.

    Taken together, these data suggest that a nonfunctional ER is not sufficient to totally explain either the ER-positive/PR-negative phenotype or the reduced efficacy of tamoxifen in these tumors. Other regulatory mechanisms must be considered.

    Hypermethylation of the PR Promoter or LOH at the PR Gene Locus

    Ferguson et al129 have shown that loss of ER in breast cancer is in part a result of hypermethylation and repression of the ER promoter, which prevents ER production. Similarly, the same group found that this suppressive form of epigenetic change may account for the loss of PR in some tumors.130 Although no methylation of the PR promoter was found in ER-positive/PR-positive breast tumors, 21% to 40% of ER-positive/PR-negative tumors showed methylation at the PR promoter. Although a definitive cause and effect relationship has not been demonstrated between PR levels and the extent of methylation of the PR gene promoter, the finding in the clinical samples may provide another mechanism that could silence PR expression in ER-positive/PR-negative breast cancer.

    Another potential mechanism for loss of PR in breast cancer emerged when two groups reported genetic loss (LOH) at the PR gene locus (chromosome 11q23).122, 123,131 LOH results in a reduction of PR expression as a result of the loss of one copy of the gene and may, in part, account for the PR-negative status of the tumor. The studies indicated that this LOH occurs in approximately 40% of primary breast cancers and is associated with loss of PR protein expression. Importantly, however, both hypermethylation and LOH fail to explain why PR-negative or PR-low tumors are more resistant to SERM therapy than PR-positive tumors because the therapy targets ER and not PR function.

    Growth Factor Pathway Downregulation of PR and SERM Resistance

    Recent studies by our group and others suggest novel mechanisms for loss of PR expression in breast cancer.50 These alternative mechanisms invoke molecular cross talk between ER and growth factor signaling pathways and the relationship between the classical and nonclassical effects of ER in breast cancer cells. On the basis of all emerging evidence, we propose here two alternative, although probably complementary, mechanisms to explain how tumors may become ER-positive/PR-negative and why this results in poor response to SERM therapy.

    Growth factor–mediated downregulation of PR. Previous studies of breast cancer cell lines have implicated growth factor signaling in repression of PR expression.132 We have shown that replenishment of ER in a specifically selected ER-negative/PR-negative MCF-7 breast cancer cell subline (C4-12) did not restore endogenous PR expression, although ER did restore estrogen induction of cyclin D1, IRS-1, and IGF-IR levels.133 Interestingly, the C4-12 cells exhibit increased HER2 levels and heregulin signaling (Lee AV, unpublished data). Similarly, in antiestrogen-resistant MCF-7 cells generated by continuous culture of the ER-positive/PR-positive parental cells in antiestrogen-supplemented medium, EGF receptor signaling was enhanced, whereas PR levels were diminished.134 Replacement of antiestrogen by estradiol failed to induce PR, whereas expression of other estrogen-responsive genes like pS2 was significantly elevated. Coincidentally, these antiestrogen-resistant MCF-7 cells, like the C4-12 cells, were grown continuously in medium that was depleted of steroid hormones like estrogen but still retained polypeptide growth factors. Therefore, it is possible that long-term cell culture in growth factor–containing medium in the absence of estrogen might suppress PR expression to undetectable levels.

    In a recent report, Konecny et al39 performed in vitro studies comparing ER and PR levels in breast cancer cell lines transfected with HER2. In ZR-75 breast cancer cells, overexpression of HER2 caused PR levels to decrease by 500-fold to those levels normally observed in ER-negative cell lines, whereas ER levels only decreased by half. Moreover, PR expression also decreased significantly in T47D cells overexpressing HER2. Although ER was also reduced, the PR downregulation is probably directly mediated by HER2 signaling because PR expression in T47D breast cancer cells is reportedly independent of ER.135 Taken together, these data suggest that growth factors downregulate PR levels independent of ER levels or activity in breast cancer cells, which is a mechanism further substantiated by results from clinical studies (discussed later).

    These unresolved phenomena raise the question of how growth factors modulate PR expression. In exploring this issue, we found that short-term treatment (hours) with IGF-I, EGF, and heregulin all sharply lowered PR levels and progestin-induced PR activity in breast cancer cells.50 This is in contrast to other estrogen-regulated genes, such as pS2, whose expression is increased by IGF-I.106 The PI3K/Akt pathway was specifically involved in this growth factor downregulation of PR levels because inhibitors of this pathway could reverse the PR downregulation. In addition, we found that this downregulation of PR is not mediated via a reduction of ER levels or ER activity, suggesting that growth factor regulation of PR is independent of ER. The downregulation of PR protein levels occurred at the level of PR mRNA and was dependent on an intact PR promoter. A recent report suggests that AP-1 may be involved in repression of the PR promoter.136 Further studies are needed to fully understand how growth factors inhibit PR promoter activity (Fig 6).

    Thus, PR levels may reflect growth factor activity within a tumor. Low or absent PR expression in some tumors indicates high IGF-IR, EGFR, and HER2 activity, and this growth factor action on PR is independent of ER function or levels. As mentioned previously, this increased growth factor signaling might even potentiate NISS on some gene promoters, while PR, at the same time, is repressed. In addition, high growth factor signaling may reduce the ability of tamoxifen to act as an antagonist, resulting in SERM resistance.

    Growth factors potentiate nonclassical ER signaling. In addition to direct downregulation of PR mRNA levels by growth factors (see previous section), experimental evidence also suggests that hyperactive growth factor signaling can perturb the classical estrogen-dependent ER function by either stimulating ligand-independent activation or shifting the ER activity in the tumor cells to act predominantly via MISS or other nonclassical modes of signaling (Fig 6). In both of these instances, growth factors may alter the set of genes that are normally regulated by estrogen. PR may be a classic example of such a gene.

    Several studies have shown that growth factors can directly modulate ER activity independent of estrogen via phosphorylation of ER itself or via phosphorylation of coregulators. This has been shown to modulate the ability of SERMs to act as agonists and antagonists. The effect of ligand-independent activation of ER on target gene expression has not been extensively studied. However, it was shown that mutation of ser305 in ER to a glutamic acid (to mimic a phosphorylation site) did result in selective induction of estrogen-responsive finger protein and cyclin D1, suggesting that ligand-independent activation of ER may induce different patterns of gene expression compared with estrogen induction of ER activity.137

    Data from several preclinical models suggest that both de novo and acquired resistance to tamoxifen is often associated with increased signaling from the EGFR/HER2 pathway138 and a switch from NISS to MISS ER activity.78,114 Importantly, although human breast cancer xenografts that overexpress HER2 are stimulated by tamoxifen as a mechanism of de novo resistance, these tumors are still completely inhibited by estrogen withdrawal.113 We recently showed that, in these tumors, both estrogen and tamoxifen can induce a rapid (but not transient) phosphorylation and thereby activation of both EGFR and HER2.113 The increased MISS ER activity and the resultant increase in tamoxifen agonist activity in these HER2-overexpressing cells is entirely dependent on ER cross talk with the growth factor signaling pathway because both phenomena were completely reversed in the presence of specific EGFR/HER2 inhibitors.113,139 Thus, in this model, estrogen and tamoxifen may actually decrease PR levels by MISS activation of growth factor signaling (Fig 6). Kumar et al114 and Mazumdar et al114,140 have reported that increased HER2 activity results in more ER outside the nucleus, and they have other evidence that classical ER-dependent transcription is inhibited by a mechanism that is HER2 dependent and involves ER sequestration in the cytoplasm via the ER corepressor MTA1s. High levels of membrane and cytoplasmic ER have also been found in breast cancer cells amplified for the HER2 gene.115 In these cells, a direct interaction between ER and HER2 in the cell membrane is associated with resistance to tamoxifen-induced apoptosis.115 Whether this change in ER localization in the cell also occurs in human tumors requires more study.

    Acquired resistance to both long-term estrogen deprivation and tamoxifen may also be associated with a shift from NISS to MISS ER activity. We and others have shown that tumors that acquire resistance to estrogen withdrawal become hypersensitive to estrogen as a mechanism of resistance and that acquired tamoxifen resistance is a result of a switch to tamoxifen-stimulated growth.141-143 The resistant tumors continue to express high levels of ER and can still be inhibited by the SERD fulvestrant (ICI 182,780). However, the enhanced response to the agonistic effects of both estrogen and tamoxifen does not occur at the level of ER-mediated transcription of classical ER-dependent genes.134,144 Expression of classically estrogen-dependent genes, including PR, is in fact reduced in the tamoxifen-resistant tumor cells. Resistance in these models is associated with upregulation of growth factor receptors (EGFR and/or HER2) and/or their downstream signaling pathways.141,142,145 Together, the experimental data suggest that at least some scenarios of acquired resistance to estrogen deprivation and tamoxifen are predominantly driven by the MISS activity of the ER.

    The underlying mechanisms responsible for the change in the balance between MISS and NISS ER activities are largely unknown. An increase in levels of cytoplasmic and/or membrane ER can lead to hyperactivation of MISS ER activity.110,146 This increased non-nuclear ER may result from an increase in total cellular ER protein, as found for example in tumors resistant to estrogen deprivation,147 or from an active cytoplasmic ER-sequestering mechanism.140 Cytoplasmic sequestration of ER would then reduce its genomic nuclear activity. In addition, increased levels of growth factor receptors, as well as other signaling molecules that interact with the ER and mediate its nongenomic signaling, might also directly enhance the nongenomic ER activity. ER and many of its coregulatory proteins, such as AIB1, are phosphorylated by a number of kinases at distinct amino acid residues in the proteins. It is also possible that phosphorylation at these specific sites could modulate the MISS and NISS activity of the ER.

    Therefore, in tumors in which MISS activity is the dominant mode of ER signaling, estrogen regulation of tumor cell proliferation and survival may no longer rely primarily on ER acting as a transcription factor for the classical set of estrogen-dependent genes. Instead, estrogen or tamoxifen binds ER and, via its MISS functions, activates growth factor receptors and their downstream signaling molecules. Such tumors, if analyzed based on their classical ER-dependent gene patterns (eg, PR status), may portray an expression profile that will misclassify them as tumors with a nonfunctional ER (eg, ER-positive/PR-negative). Furthermore, because MISS signaling may originate from ER molecules confined mainly to the membrane and the cytoplasm, some of these tumors could even be mistakenly classified as ER negative by immunostaining.114 More sensitive methods for detecting and quantifying this non-nuclear pool of ER are needed to address these questions.

    Mechanistically, however, these tumors with high ER MISS activity should still be highly dependent on estrogen and, therefore, should still be highly sensitive to estrogen withdrawal therapy. In contrast, SERMs like tamoxifen stimulate at least some of the MISS actions of ER including activation of growth factor receptors and downstream effectors such as Akt and ERK1/2.148 The sensitivity of these tumors to SERMs may differ substantially from tumors with low growth factor activity in which ER functions primarily as a nuclear transcription factor and in which tamoxifen would be expected to function as an antagonist to block growth. The switch from predominant NISS ER activity to MISS ER activity could also explain the ER-positive/PR-negative phenotype in tumors with active growth factor signaling pathways and why aromatase inhibitors are superior to tamoxifen in such patients.37

    CLINICAL IMPLICATIONS OF GROWTH FACTOR REGULATION OF PR LEVELS

    PR As a Predictive Marker of Both ER and Growth Factor Action

    In theory, PR may be a better indicator than ER for predicting response to SERM therapy because levels of PR reflect the combined and integrated effects of ER and growth factor activity. As stated earlier, several growth factor pathways can activate ER, and ER, in turn, can potentiate these same growth factor signaling pathways.74 Interestingly, the trimeric G protein and integrin pathways, which are also capable of triggering PI3K/Akt activation, are also involved in breast cancer development and can cross talk with ER and growth factor signaling. Thus, more precisely speaking, PR may serve as an indicator for intracellular kinase networking elicited by overall extracellular stimulations.

    Although there are many ER-regulated genes, PR is unique from the perspective of its distinctive mode of regulation by estrogen and growth factors. Cyclin D1, for example, is induced by estrogen, IGF, and EGF.112,149,150 Another well-known estrogen-induced gene, IRS-1, is decreased by IGF-I151 but increased by EGF.152 Hence, PR may be an ideal marker as a readout of both ER and growth factor action and, thus, may aid in the prediction of response to various hormone or growth factor receptor–inhibitor therapies, with the caveat that other mechanisms for PR loss reviewed earlier and false-negative PR assays as a result of low circulating estrogen levels in some older women need to be considered. A 24- to 48-hour treatment with estrogen before biopsy to induce PR levels may help to identify this group.

    PR measurement is certainly useful, but it is clear that new methodologies and assays need to be developed that measure the new activities of ER, such as MISS. Currently, ER positivity is scored by pathologists based on nuclear staining (using antibodies optimized to detect only nuclear ER), a situation that may result in misclassification of tumors that have low levels of cytoplasmic or membrane ER staining that can be detected with other antibodies. In addition, the realization that ER controls a major network of genes suggests that a more comprehensive analysis of the ER-controlled transcriptome may yield more beneficial information than simply ER and PR alone. Expression array profiling of these tumors may help to subclassify patients for treatment purposes.

    Treatment of ER-Positive/PR-Negative Tumors by Estrogen Withdrawal, SERDs, or Combined SERM and Antigrowth Factor Therapy

    Laboratory and clinical evidence indicate that SERMs may be less effective in the presence of high growth factor signaling. In this situation, growth factor–mediated phosphorylation of ER and its coregulators can cause SERMs to show full agonist activity, stimulating tumor growth via NISS, MISS, or both forms of ER activity. In this situation, the most effective way to block ER activity is to fully withdraw any ligand from the ER, rather than by providing a SERM that is seen as an agonist. Therefore, ER-positive/PR-negative tumors, which may show high growth factor signaling, may best be treated by estrogen withdrawal using an aromatase inhibitor, as seen in the ATAC trial (Fig 3). Another effective treatment for these tumors may be the targeted degradation of ER via SERDs. Finally, the combined use of SERMs and antigrowth factor therapies may be beneficial. Although this has yet to be proven in clinical trials, recent preclinical data suggest that the combination of a SERM and an EGFR/HER2 inhibitor is beneficial in ER-positive HER2-overexpressing breast cancer cells.113,138 In this situation, ER and growth factor cross talk results in tamoxifen resistance, with tamoxifen exhibiting estrogenic agonist activity and stimulation of tumor growth. However, treatment with an EGFR or HER2 inhibitor restores the antagonist action of tamoxifen. After the excitement of these preclinical studies, several clinical trials are now underway combining growth factor receptor tyrosine kinase inhibitors and antibody antagonists with SERMs or aromatase inhibitors.

    One extension of our studies suggests that ER-positive/PR-positive tumors must not exhibit strong growth factor signaling. This would imply that inhibitors of growth factor signaling cascades will show little benefit in patients with these tumors and will only work effectively in tumors that are PR negative (and maybe only a subset of these tumors) and if administered in combination with endocrine therapies to block ER-related growth signals. Combinations of a SERM plus a growth factor pathway inhibitor might still be considered for ER-positive/PR-positive tumors because, during treatment, there may be an increase in growth factor activity as a mechanism of resistance, and subsequent loss of PR may occur in some of these tumors that are not completely eradicated. It is tempting to speculate that adjuvant tamoxifen for several years followed by an aromatase inhibitor to treat resistant cells or initial therapy with tamoxifen combined with a growth factor inhibitor might be better than an initial use of an aromatase inhibitor in ER-positive/PR-positive tumors. In contrast, for ER-positive/PR-negative tumors, initial treatment with an aromatase inhibitor may be most favorable. Careful clinical trial design and tissue collections for future molecular studies are needed to address these questions.

    SUMMARY

    Tremendous progress has been made in the treatment of breast cancer. A wealth of knowledge acquired in the last decade in the field of molecular endocrinology and tumor biology has advanced our understanding of the mechanisms of ER action and has led to substantial improvements in the efficacy of cancer treatment and prevention. However, there remains the challenge of developing treatments that are as effective against metastatic breast cancer as they are now for early-stage breast cancer and that will prevent or overcome resistance to endocrine therapy. Loss of PR may be a marker of aberrant growth factor signaling and, consequently, one mechanism for antiestrogen resistance. We anticipate that ER-positive/PR-negative breast cancer probably relies on more than one mechanism for its aggressive phenotype, and understanding why PR-negative tumors respond poorly to endocrine treatment with SERMs could pave the way for the development of better therapeutic strategies.

    Authors' Disclosures of Potential Conflicts of Interest

    Although all authors completed the disclosure declaration, the following author or immediate family members indicated a financial interest. No conflict exists for drugs or devices used in a study if they are not being evaluated as part of the investigation. For a detailed description of the disclosure categories, or for more information about ASCO's conflict of interest policy, please refer to the Author Disclosure Declaration and the Disclosures of Potential Conflicts of Interest section in Information for Contributors.

    Acknowledgment

    We thank Gary Chamness, PhD, and others in the Breast Center for helpful suggestions and critical reading of the manuscript. We apologize to those authors whose outstanding work in this area we were not able to cite.

    NOTES

    Supported by National Institutes of Health grant Nos. CA94118 (A.V.L.), P50 CA58183 (C.K.O.), and P01 CA30195 (C.K.O.), and the Department of Defense Postdoctoral Fellowship Award DAMD17-01-1-0133 (X.C.).

    Authors' disclosures of potential conflicts of interest are found at the end of this article.

    REFERENCES

    Campos SM, Winer EP: Hormonal therapy in postmenopausal women with breast cancer. Oncology 64:289-299, 2003

    Osborne CK, Zhao H, Fuqua SA: Selective estrogen receptor modulators: Structure, function, and clinical use. J Clin Oncol 18:3172-3186, 2000

    Baum M, Buzdar A: The current status of aromatase inhibitors in the management of breast cancer. Surg Clin North Am 83:973-994, 2003

    Buzdar AU: Advances in endocrine treatments for postmenopausal women with metastatic and early breast cancer. Oncologist 8:335-341, 2003

    Fendrick JL, Raafat AM, Haslam SZ: Mammary gland growth and development from the postnatal period to postmenopause: Ovarian steroid receptor ontogeny and regulation in the mouse. J Mammary Gland Biol Neoplasia 3:7-22, 1998

    Allred DC, Mohsin SK, Fuqua SA: Histological and biological evolution of human premalignant breast disease. Endocr Relat Cancer 8:47-61, 2001

    Saji S, Jensen EV, Nilsson S, et al: Estrogen receptors alpha and beta in the rodent mammary gland. Proc Natl Acad Sci U S A 97:337-342, 2000

    Bocchinfuso WP, Korach KS: Mammary gland development and tumorigenesis in estrogen receptor knockout mice. J Mammary Gland Biol Neoplasia 2:323-334, 1997

    Conneely OM, Jericevic BM, Lydon JP: Progesterone receptors in mammary gland development and tumorigenesis. J Mammary Gland Biol Neoplasia 8:205-214, 2003

    Kraus WL, Montano MM, Katzenellenbogen BS: Cloning of the rat progesterone receptor gene 5'-region and identification of two functionally distinct promoters. Mol Endocrinol 7:1603-1616, 1993

    Li X, Lonard DM, O'Malley BW: A contemporary understanding of progesterone receptor function. Mech Ageing Dev 125:669-678, 2004

    Lydon JP, DeMayo FJ, Conneely OM, et al: Reproductive phenotpes of the progesterone receptor null mutant mouse. J Steroid Biochem Mol Biol 56:67-77, 1996

    Shyamala G, Yang X, Cardiff RD, et al: Impact of progesterone receptor on cell-fate decisions during mammary gland development. Proc Natl Acad Sci U S A 97:3044-3049, 2000

    Hopp TA, Weiss HL, Hilsenbeck SG, et al: Breast cancer patients with progesterone receptor PR-A-rich tumors have poorer disease-free survival rates. Clin Cancer Res 10:2751-2760, 2004

    De Vivo I, Hankinson SE, Colditz GA, et al: A functional polymorphism in the progesterone receptor gene is associated with an increase in breast cancer risk. Cancer Res 63:5236-5238, 2003

    Rossouw JE, Anderson GL, Prentice RL, et al: Risks and benefits of estrogen plus progestin in healthy postmenopausal women: Principal results from the Women's Health Initiative randomized controlled trial. JAMA 288:321-333, 2002

    McGuire WL: Hormone receptors: Their role in predicting prognosis and response to endocrine therapy. Semin Oncol 5:428-433, 1978

    Bardou VJ, Arpino G, Elledge RM, et al: Progesterone receptor status significantly improves outcome prediction over estrogen receptor status alone for adjuvant endocrine therapy in two large breast cancer databases. J Clin Oncol 21:1973-1979, 2003

    Hull DF, Clark GM, Osborne CK, et al: Multiple estrogen receptor assays in human breast cancer. Cancer Res 43:413-416, 1983

    Gross GE, Clark GM, Chamness GC, et al: Multiple progesterone receptor assays in human breast cancer. Cancer Res 44:836-840, 1984

    Balleine RL, Earl MJ, Greenberg ML, et al: Absence of progesterone receptor associated with secondary breast cancer in postmenopausal women. Br J Cancer 79:1564-1571, 1999

    Potter JD, Cerhan JR, Sellers TA, et al: Progesterone and estrogen receptors and mammary neoplasia in the Iowa Women's Health Study: How many kinds of breast cancer are there? Cancer Epidemiol Biomarkers Prev 4:319-326, 1995

    Anderson WF, Chatterjee N, Ershler WB, et al: Estrogen receptor breast cancer phenotypes in the Surveillance, Epidemiology, and End Results database. Breast Cancer Res Treat 76:27-36, 2002

    Colditz GA, Rosner BA, Chen WY, et al: Risk factors for breast cancer according to estrogen and progesterone receptor status. J Natl Cancer Inst 96:218-228, 2004

    Arpino G, Weiss H, Lee AV, et al: Estrogen receptor-positive, progesterone receptor-negative breast cancer: Association with growth factor receptor expression and tamoxifen resistance. J Natl Cancer Inst 97:1254-1261, 2005

    Early Breast Cancer Trialists' Collaborative Group: Systemic treatment of early breast cancer by hormonal, cytotoxic, or immune therapy: 133 randomised trials involving 31,000 recurrences and 24,000 deaths among 75,000 women. Lancet 339:1-15, 71-85, 1992

    Fisher B, Anderson S, Tan-Chiu E, et al: Tamoxifen and chemotherapy for axillary node-negative, estrogen receptor-negative breast cancer: Findings from National Surgical Adjuvant Breast and Bowel Project B-23. J Clin Oncol 19:931-942, 2001

    Hutchins L, Green S, Ravdin L: CMF versus CAF with and without tamoxifen in high-risk node-negative breast cancer patients and a natural history follow-up study in low-risk node-negative patients: First results of Intergroup trial INT0102. Proc Am Soc Clin Oncol 17:1a, 1998 (abstr 2)

    Elledge RM, Green S, Pugh R, et al: Estrogen receptor (ER) and progesterone receptor (PgR), by ligand-binding assay compared with ER, PgR and pS2, by immuno-histochemistry in predicting response to tamoxifen in metastatic breast cancer: A Southwest Oncology Group Study. Int J Cancer 89:111-117, 2000

    Osborne CK, Yochmowitz MG, Knight WA III, et al: The value of estrogen and progesterone receptors in the treatment of breast cancer. Cancer 46:2884-2888, 1980

    Harvey JM, Clark GM, Osborne CK, et al: Estrogen receptor status by immunohistochemistry is superior to the ligand-binding assay for predicting response to adjuvant endocrine therapy in breast cancer. J Clin Oncol 17:1474-1481, 1999

    Ferno M, Stal O, Baldetorp B, et al: Results of two or five years of adjuvant tamoxifen correlated to steroid receptor and S-phase levels: South Sweden Breast Cancer Group, and South-East Sweden Breast Cancer Group. Breast Cancer Res Treat 59:69-76, 2000

    Lamy PJ, Pujol P, Thezenas S, et al: Progesterone receptor quantification as a strong prognostic determinant in postmenopausal breast cancer women under tamoxifen therapy. Breast Cancer Res Treat 76:65-71, 2002

    Ravdin PM, Green S, Dorr TM, et al: Prognostic significance of progesterone receptor levels in estrogen receptor-positive patients with metastatic breast cancer treated with tamoxifen: Results of a prospective Southwest Oncology Group study. J Clin Oncol 10:1284-1291, 1992

    Ellis MJ, Coop A, Singh B, et al: Letrozole is more effective neoadjuvant endocrine therapy than tamoxifen for ErbB-1- and/or ErbB-2-positive, estrogen receptor-positive primary breast cancer: Evidence from a phase III randomized trial. J Clin Oncol 19:3808-3816, 2001

    Dowsett M, Harper-Wynne C, Boeddinghaus I, et al: HER-2 amplification impedes the antiproliferative effects of hormone therapy in estrogen receptor-positive primary breast cancer. Cancer Res 61:8452-8458, 2001

    Baum M, Buzdar A, Cuzick J, et al: Anastrozole alone or in combination with tamoxifen versus tamoxifen alone for adjuvant treatment of postmenopausal women with early-stage breast cancer: Results of the ATAC (Arimidex, Tamoxifen Alone or in Combination) trial efficacy and safety update analyses. Cancer 98:1802-1810, 2003

    Dowsett M, on behalf of the ATAC trialists group: Analysis of time to recurrence in the ATAC (Arimidex, Tamoxifen, Alone or in Combination) trial according to estrogen receptor and progesterone receptor status. Breast Cancer Res Treat 82:S7, 2003 (suppl 1)

    Horwitz KB, McGuire WL, Pearson OH, et al: Predicting response to endocrine therapy in human breast cancer. Science 189:726-727, 1975

    Konecny G, Pauletti G, Pegram M, et al: Quantitative association between HER-2/neu and steroid hormone receptors in hormone receptor-positive primary breast cancer. J Natl Cancer Inst 95:142-153, 2003

    Bamberger AM, Milde-Langosch K, Schulte HM, et al: Progesterone receptor isoforms, PR-B and PR-A, in breast cancer: Correlations with clinicopathologic tumor parameters and expression of AP-1 factors. Horm Res 54:32-37, 2000

    Ross JS, Fletcher JA, Bloom KJ, et al: Targeted therapy in breast cancer: The HER-2/neu gene and protein. Mol Cell Proteomics 3:370-398, 2004

    Arpino G, Green S, Allred DC, et al: HER-2 amplification, HER-1 expression, and tamoxifen response in ER-positive metastatic breast cancer: A Southwest Oncology Group Study. Clin Cancer Res 10:5670-5676, 2004

    Elledge RM, Green S, Ciocca D, et al: HER-2 expression and response to tamoxifen in estrogen receptor-positive breast cancer: A Southwest Oncology Group Study. Clin Cancer Res 4:7-12, 1998

    Gutierrez MC, Detre S, Johnston S, et al: Molecular changes in tamoxifen-resistant breast cancer: Relationship between estrogen receptor, HER-2, and p38 mitogen-activated protein kinase. J Clin Oncol 23:2469-2476, 2005

    Meng S, Tripathy D, Shete S, et al: HER-2 gene amplification can be acquired as breast cancer progresses. Proc Natl Acad Sci U S A 101:9393-9398, 2004

    Osborne CK, Bardou V, Hopp TA, et al: Role of the estrogen receptor coactivator AIB1 (SRC-3) and HER-2/neu in tamoxifen resistance in breast cancer. J Natl Cancer Inst 95:353-361, 2003

    Shi W, Zhang X, Pintilie M, et al: Dysregulated PTEN-PKB and negative receptor status in human breast cancer. Int J Cancer 104:195-203, 2003

    Perren A, Weng LP, Boag AH, et al: Immunohistochemical evidence of loss of PTEN expression in primary ductal adenocarcinomas of the breast. Am J Pathol 155:1253-1260, 1999

    Garcia JM, Silva JM, Dominguez G, et al: Allelic loss of the PTEN region (10q23) in breast carcinomas of poor pathophenotype. Breast Cancer Res Treat 57:237-243, 1999

    Cui X, Zhang P, Deng W, et al: Insulin-like growth factor-I inhibits progesterone receptor expression in breast cancer cells via the phosphatidylinositol 3-kinase/Akt/mammalian target of rapamycin pathway: Progesterone receptor as a potential indicator of growth factor activity in breast cancer. Mol Endocrinol 17:575-588, 2003

    Campbell RA, Bhat-Nakshatri P, Patel NM, et al: Phosphatidylinositol 3-kinase/AKT-mediated activation of estrogen receptor alpha: A new model for anti-estrogen resistance. J Biol Chem 276:9817-9824, 2001

    Faridi J, Wang L, Endemann G, et al: Expression of constitutively active Akt-3 in MCF-7 breast cancer cells reverses the estrogen and tamoxifen responsivity of these cells in vivo. Clin Cancer Res 9:2933-2939, 2003

    Kurokawa H, Arteaga CL: ErbB (HER) receptors can abrogate antiestrogen action in human breast cancer by multiple signaling mechanisms. Clin Cancer Res 9:511S-515S, 2003

    Perez-Tenorio G, Stal O: Activation of AKT/PKB in breast cancer predicts a worse outcome among endocrine treated patients. Br J Cancer 86:540-545, 2002

    Parker MG: Steroid and related receptors. Curr Opin Cell Biol 5:499-504, 1993

    Nemere I, Pietras RJ, Blackmore PF: Membrane receptors for steroid hormones: Signal transduction and physiological significance. J Cell Biochem 88:438-445, 2003

    Horwitz KB, McGuire WL: Estrogen control of progesterone receptor in human breast cancer. J Biol Chem 253:2223-2228, 1978

    Beato M: Gene regulation by steroid hormones. Cell 56:335-344, 1989

    Klein-Hitpass L, Ryffel GU, Heitlinger E, et al: A 13bp palindrome is a functional estrogen responsive element and interacts specifically with estrogen receptor. Nucleic Acids Res 16:647-664, 1988

    Kushner PJ, Agard DA, Greene GL, et al: Estrogen receptor pathways to AP-1. J Steroid Biochem Mol Biol 74:311-317, 2000

    Safe S: Transcriptional activation of genes by 17 beta-estradiol through estrogen receptor-Sp1 interactions. Vitam Horm 62:231-252, 2001

    Xing W, Archer TK: Upstream stimulatory factors mediate estrogen receptor activation of the cathepsin D promoter. Mol Endocrinol 12:1310-1321, 1998

    Dobrzycka KM, Townson SM, Jiang S, et al: Estrogen receptor corepressors: A role in human breast cancer? Endocr Relat Cancer 10:517-536, 2003

    Smith CL, O'Malley BW: Coregulator function: A key to understanding tissue specificity of selective receptor modulators. Endocr Rev 25:45-71, 2004

    Smith CL, Nawaz Z, O'Malley BW: Coactivator and corepressor regulation of the agonist/antagonist activity of the mixed antiestrogen, 4-hydroxytamoxifen. Mol Endocrinol 11:657-666, 1997

    Shang Y, Brown M: Molecular determinants for the tissue specificity of SERMs. Science 295:2465-2468, 2002

    Xu J, Li Q: Review of the in vivo functions of the p160 steroid receptor coactivator family. Mol Endocrinol 17:1681-1692, 2003

    Fan S, Wang J, Yuan R, et al: BRCA1 inhibition of estrogen receptor signaling in transfected cells. Science 284:1354-1356, 1999

    Fan S, Ma YX, Wang C, et al: Role of direct interaction in BRCA1 inhibition of estrogen receptor activity. Oncogene 20:77-87, 2001

    Zwijsen RM, Wientjens E, Klompmaker R, et al: CDK-independent activation of estrogen receptor by cyclin D1. Cell 88:405-415, 1997

    Neuman E, Ladha MH, Lin N, et al: Cyclin D1 stimulation of estrogen receptor transcriptional activity independent of cdk4. Mol Cell Biol 17:5338-5347, 1997

    Zwijsen RM, Buckle RS, Hijmans EM, et al: Ligand-independent recruitment of steroid receptor coactivators to estrogen receptor by cyclin D1. Genes Dev 12:3488-3498, 1998

    McMahon C, Suthiphongchai T, DiRenzo J, et al: P/CAF associates with cyclin D1 and potentiates its activation of the estrogen receptor. Proc Natl Acad Sci U S A 96:5382-5387, 1999

    Lee AV, Cui X, Oesterreich S: Cross-talk among estrogen receptor, epidermal growth factor, and insulin-like growth factor signaling in breast cancer. Clin Cancer Res 7:4429s-4435s, 2001

    Smith CL, Conneely OM, O'Malley BW: Modulation of the ligand-independent activation of the human estrogen receptor by hormone and antihormone. Proc Natl Acad Sci U S A 90:6120-6124, 1993

    Aronica SM, Katzenellenbogen BS: Stimulation of estrogen receptor-mediated transcription and alteration in the phosphorylation state of the rat uterine estrogen receptor by estrogen, cyclic adenosine monophosphate, and insulin-like growth factor-I. Mol Endocrinol 7:743-752, 1993

    Auricchio F, Di Domenico M, Migliaccio A, et al: The role of estradiol receptor in the proliferative activity of vanadate on MCF-7 cells. Cell Growth Differ 6:105-113, 1995

    Schiff R, Massarweh SA, Shou J, et al: Cross-talk between estrogen receptor and growth factor pathways as a molecular target for overcoming endocrine resistance. Clin Cancer Res 10:331S-336S, 2004

    Kato S, Endoh H, Masuhiro Y, et al: Activation of the estrogen receptor through phosphorylation by mitogen-activated protein kinase. Science 270:1491-1494, 1995

    Lee H, Jiang F, Wang Q, et al: MEKK1 activation of human estrogen receptor alpha and stimulation of the agonistic activity of 4-hydroxytamoxifen in endometrial and ovarian cancer cells. Mol Endocrinol 14:1882-1896, 2000

    Rogatsky I, Trowbridge JM, Garabedian MJ: Potentiation of human estrogen receptor alpha transcriptional activation through phosphorylation of serines 104 and 106 by the cyclin A-CDK2 complex. J Biol Chem 274:22296-22302, 1999

    Chen D, Washbrook E, Sarwar N, et al: Phosphorylation of human estrogen receptor alpha at serine 118 by two distinct signal transduction pathways revealed by phosphorylation-specific antisera. Oncogene 21:4921-4931, 2002

    Arnold SF, Obourn JD, Jaffe H, et al: Phosphorylation of the human estrogen receptor on tyrosine 537 in vivo and by src family tyrosine kinases in vitro. Mol Endocrinol 9:24-33, 1995

    Le Goff P, Montano MM, Schodin DJ, et al: Phosphorylation of the human estrogen receptor: Identification of hormone-regulated sites and examination of their influence on transcriptional activity. J Biol Chem 269:4458-4466, 1994

    Joel PB, Smith J, Sturgill TW, et al: Pp90rsk1 regulates estrogen receptor-mediated transcription through phosphorylation of Ser-167. Mol Cell Biol 18:1978-1984, 1998

    Ali S, Metzger D, Bornert JM, et al: Modulation of transcriptional activation by ligand-dependent phosphorylation of the human oestrogen receptor A/B region. Embo J 12:1153-1160, 1993

    Chen D, Pace PE, Coombes RC, et al: Phosphorylation of human estrogen receptor alpha by protein kinase A regulates dimerization. Mol Cell Biol 19:1002-1015, 1999

    Arnold SF, Vorojeikina DP, Notides AC: Phosphorylation of tyrosine 537 on the human estrogen receptor is required for binding to an estrogen response element. J Biol Chem 270:30205-30212, 1995

    Lee H, Bai W: Regulation of estrogen receptor nuclear export by ligand-induced and p38-mediated receptor phosphorylation. Mol Cell Biol 22:5835-5845, 2002

    Font de Mora J, Brown M: AIB1 is a conduit for kinase-mediated growth factor signaling to the estrogen receptor. Mol Cell Biol 20:5041-5047, 2000

    Wu RC, Qin J, Hashimoto Y, et al: Regulation of SRC-3 (pCIP/ACTR/AIB-1/RAC-3/TRAM-1) coactivator activity by I kappa B kinase. Mol Cell Biol 22:3549-3561, 2002

    Wu RC, Qin J, Yi P, et al: Selective phosphorylations of the SRC-3/AIB1 coactivator integrate genomic responses to multiple cellular signaling pathways. Mol Cell 15:937-949, 2004

    Lopez GN, Turck CW, Schaufele F, et al: Growth factors signal to steroid receptors through mitogen-activated protein kinase regulation of p160 coactivator activity. J Biol Chem 276:22177-22182, 2001

    Lavinsky RM, Jepsen K, Heinzel T, et al: Diverse signaling pathways modulate nuclear receptor recruitment of N-CoR and SMRT complexes. Proc Natl Acad Sci U S A 95:2920-2925, 1998

    Hong SH, Privalsky ML: The SMRT corepressor is regulated by a MEK-1 kinase pathway: Inhibition of corepressor function is associated with SMRT phosphorylation and nuclear export. Mol Cell Biol 20:6612-6625, 2000

    Baek SH, Ohgi KA, Rose DW, et al: Exchange of N-CoR corepressor and Tip60 coactivator complexes links gene expression by NF-kappaB and beta-amyloid precursor protein. Cell 110:55-67, 2002

    Levin ER: Cellular functions of the plasma membrane estrogen receptor. Trends Endocrinol Metab 10:374-377, 1999

    Razandi M, Pedram A, Greene GL, et al: Cell membrane and nuclear estrogen receptors (ERs) originate from a single transcript: Studies of ERalpha and ERbeta expressed in Chinese hamster ovary cells. Mol Endocrinol 13:307-319, 1999

    Pappas TC, Gametchu B, Watson CS: Membrane estrogen receptors identified by multiple antibody labeling and impeded-ligand binding. FASEB J 9:404-410, 1995

    Li L, Haynes MP, Bender JR: Plasma membrane localization and function of the estrogen receptor alpha variant (ER46) in human endothelial cells. Proc Natl Acad Sci U S A 100:4807-4812, 2003

    Figtree GA, McDonald D, Watkins H, et al: Truncated estrogen receptor alpha 46-kDa isoform in human endothelial cells: Relationship to acute activation of nitric oxide synthase. Circulation 107:120-126, 2003

    Filardo EJ: Epidermal growth factor receptor (EGFR) transactivation by estrogen via the G-protein-coupled receptor, GPR30: A novel signaling pathway with potential significance for breast cancer. J Steroid Biochem Mol Biol 80:231-238, 2002

    Tsai EM, Wang SC, Lee JN, et al: Akt activation by estrogen in estrogen receptor-negative breast cancer cells. Cancer Res 61:8390-8392, 2001

    Stewart AJ, Johnson MD, May FE, et al: Role of insulin-like growth factors and the type I insulin-like growth factor receptor in the estrogen-stimulated proliferation of human breast cancer cells. J Biol Chem 265:21172-21178, 1990

    Lee AV, Jackson JG, Gooch JL, et al: Enhancement of insulin-like growth factor signaling in human breast cancer: Estrogen regulation of insulin receptor substrate-1 expression in vitro and in vivo. Mol Endocrinol 13:787-796, 1999

    Lee AV, Weng CN, Jackson JG, et al: Activation of estrogen receptor-mediated gene transcription by IGF-I in human breast cancer cells. J Endocrinol 152:39-47, 1997

    Kahlert S, Nuedling S, van Eickels M, et al: Estrogen receptor alpha rapidly activates the IGF-1 receptor pathway. J Biol Chem 275:18447-18453, 2000

    Simoncini T, Hafezi-Moghadam A, Brazil DP, et al: Interaction of oestrogen receptor with the regulatory subunit of phosphatidylinositol-3-OH kinase. Nature 407:538-541, 2000

    Morelli C, Garofalo C, Bartucci M, et al: Estrogen receptor-alpha regulates the degradation of insulin receptor substrates 1 and 2 in breast cancer cells. Oncogene 22:4007-4016, 2003

    Song RX, McPherson RA, Adam L, et al: Linkage of rapid estrogen action to MAPK activation by ERalpha-Shc association and Shc pathway activation. Mol Endocrinol 16:116-127, 2002

    Anderson RG: The caveolae membrane system. Annu Rev Biochem 67:199-225, 1998

    Razandi M, Alton G, Pedram A, et al: Identification of a structural determinant necessary for the localization and function of estrogen receptor alpha at the plasma membrane. Mol Cell Biol 23:1633-1646, 2003

    Shou J, Massarweh S, Osborne CK, et al: Mechanisms of tamoxifen resistance: Increased estrogen receptor-HER2/neu cross-talk in ER/HER2-positive breast cancer. J Natl Cancer Inst 96:926-935, 2004

    Kumar R, Wang RA, Mazumdar A, et al: A naturally occurring MTA1 variant sequesters oestrogen receptor-alpha in the cytoplasm. Nature 418:654-657, 2002

    Chung YL, Sheu ML, Yang SC, et al: Resistance to tamoxifen-induced apoptosis is associated with direct interaction between Her2/neu and cell membrane estrogen receptor in breast cancer. Int J Cancer 97:306-312, 2002

    Wong CW, McNally C, Nickbarg E, et al: Estrogen receptor-interacting protein that modulates its nongenomic activity-crosstalk with Src/Erk phosphorylation cascade. Proc Natl Acad Sci U S A 99:14783-14788, 2002

    Vadlamudi RK, Wang RA, Mazumdar A, et al: Molecular cloning and characterization of PELP1, a novel human coregulator of estrogen receptor alpha. J Biol Chem 276:38272-38279, 2001

    Balasenthil S, Vadlamudi RK: Functional interactions between the estrogen receptor coactivator PELP1/MNAR and retinoblastoma protein. J Biol Chem 278:22119-22127, 2003

    Xue Y, Wong J, Moreno GT, et al: NURD, a novel complex with both ATP-dependent chromatin-remodeling and histone deacetylase activities. Mol Cell 2:851-861, 1998

    Kumar R, Wang RA, Bagheri-Yarmand R: Emerging roles of MTA family members in human cancers. Semin Oncol 30:30-37, 2003

    Bloom ND, Tobin EH, Schreibman B, et al: The role of progesterone receptors in the management of advanced breast cancer. Cancer 45:2992-2997, 1980

    Tomlinson IP, Nicolai H, Solomon E, et al: The frequency and mechanism of loss of heterozygosity on chromosome 11q in breast cancer. J Pathol 180:38-43, 1996

    Winqvist R, Hampton GM, Mannermaa A, et al: Loss of heterozygosity for chromosome 11 in primary human breast tumors is associated with poor survival after metastasis. Cancer Res 55:2660-2664, 1995

    Lapidus RG, Ferguson AT, Ottaviano YL, et al: Methylation of estrogen and progesterone receptor gene 5' CpG islands correlates with lack of estrogen and progesterone receptor gene expression in breast tumors. Clin Cancer Res 2:805-810, 1996

    Singh A, Ali S, Kothari MS, et al: Reporter gene assay demonstrates functional differences in estrogen receptor activity in purified breast cancer cells: A pilot study. Int J Cancer 107:700-706, 2003

    Fuqua SAW, Allred DC, Elledge RM, et al: The ER-positive/PgR-negative breast cancer phenotype is not associated with mutations within the DNA binding domain. Breast Cancer Res Treat 26:191-202, 1993

    Mies C, Voigt W: Sequence analysis of the DNA binding domain of the estrogen receptor gene in ER (+)/PR (–) breast cancer. Diagn Mol Pathol 5:39-44, 1996

    Liang X, Lu B, Scott GK, et al: Oxidant stress impaired DNA-binding of estrogen receptor from human breast cancer. Mol Cell Endocrinol 146:151-161, 1998

    Ferguson AT, Lapidus RG, Baylin SB, et al: Demethylation of the estrogen receptor gene in estrogen receptor-negative breast cancer cells can reactivate estrogen receptor gene expression. Cancer Res 55:2279-2283, 1995

    Lapidus RG, Nass SJ, Davidson NE: The loss of estrogen and progesterone receptor gene expression in human breast cancer. J Mammary Gland Biol Neoplasia 3:85-94, 1998

    Rousseau-Merck MF, Misrahi M, Loosfelt H, et al: Localization of the human progesterone receptor gene to chromosome 11q22-q23. Hum Genet 77:280-282, 1987

    Cormier EM, Wolf MF, Jordan VC: Decrease in estradiol-stimulated progesterone receptor production in MCF-7 cells by epidermal growth factor and possible clinical implication for paracrine-regulated breast cancer growth. Cancer Res 49:576-580, 1989

    Oesterreich S, Zhang P, Guler RL, et al: Re-expression of estrogen receptor alpha in estrogen receptor alpha-negative MCF-7 cells restores both estrogen and insulin-like growth factor-mediated signaling and growth. Cancer Res 61:5771-5777, 2001

    McClelland RA, Barrow D, Madden TA, et al: Enhanced epidermal growth factor receptor signaling in MCF7 breast cancer cells after long-term culture in the presence of the pure antiestrogen ICI 182,780 (Faslodex). Endocrinology 142:2776-2788, 2001

    Horwitz KB, Mockus MB, Lessey BA: Variant T47D human breast cancer cells with high progesterone-receptor levels despite estrogen and antiestrogen resistance. Cell 28:633-642, 1982

    Petz LN, Ziegler YS, Schultz JR, et al: Fos and Jun inhibit estrogen-induced transcription of the human progesterone receptor gene through an activator protein-1 site. Mol Endocrinol 18:521-532, 2004

    Balasenthil S, Barnes CJ, Rayala SK, et al: Estrogen receptor activation at serine 305 is sufficient to upregulate cyclin D1 in breast cancer cells. FEBS Lett 567:243-247, 2004

    Kurokawa H, Lenferink AE, Simpson JF, et al: Inhibition of HER2/neu (erbB-2) and mitogen-activated protein kinases enhances tamoxifen action against HER2-overexpressing, tamoxifen-resistant breast cancer cells. Cancer Res 60:5887-5894, 2000

    Barhwani L, Schiff R, Mohsin SK, et al: Inhibiting the EGFR/HER2 pathway with gefitinib and/or trastuzumab restores tamoxifen sensitivity in HER-2 overexpressing tumors. Breast Cancer Res Treat 82:S13, 2003 (suppl)

    Mazumdar A, Wang RA, Mishra SK, et al: Transcriptional repression of oestrogen receptor by metastasis-associated protein 1 corepressor. Nat Cell Biol 3:30-37, 2001

    Song RX, Santen RJ, Kumar R, et al: Adaptive mechanisms induced by long-term estrogen deprivation in breast cancer cells. Mol Cell Endocrinol 193:29-42, 2002

    Schafer JM, Bentrem DJ, Takei H, et al: A mechanism of drug resistance to tamoxifen in breast cancer. J Steroid Biochem Mol Biol 83:75-83, 2002

    Osborne CK, Coronado-Heinsohn EB, Hilsenbeck SH, et al: Comparison of the effects of a pure steroidal antiestrogen with those of tamoxifen in a model of human breast cancer. J Natl Cancer Inst 87:746-750, 1995

    Osborne CK, Coronado E, Allred DC, et al: Acquired tamoxifen resistance: Correlation with reduced breast tumor levels of tamoxifen and isomerization of trans-4-hydroxytamoxifen. J Natl Cancer Inst 83:1477-1482, 1991

    Knowlden JM, Hutcheson IR, Jones HE, et al: Elevated levels of epidermal growth factor receptor/c-erbB2 heterodimers mediate an autocrine growth regulatory pathway in tamoxifen-resistant MCF-7 cells. Endocrinology 144:1032-1044, 2003

    Zhang Z, Maier B, Santen RJ, et al: Membrane association of estrogen receptor alpha mediates estrogen effect on MAPK activation. Biochem Biophys Res Commun 294:926-933, 2002

    Jeng MH, Shupnik MA, Bender TP, et al: Estrogen receptor expression and function in long-term estrogen-deprived human breast cancer cells. Endocrinology 139:4164-4174, 1998

    Osborne CK, Schiff R: Growth factor receptor cross-talk with estrogen receptor as a mechanism for tamoxifen resistance in breast cancer. Breast 12:362-367, 2003

    Liu MM, Albanese C, Anderson CM, et al: Opposing action of estrogen receptors alpha and beta on cyclin D1 gene expression. J Biol Chem 277:24353-24360, 2002

    Dufourny B, van Teeffelen HA, Hamelers IH, et al: Stabilization of cyclin D1 mRNA via the phosphatidylinositol 3-kinase pathway in MCF-7 human breast cancer cells. J Endocrinol 166:329-338, 2000

    Lee AV, Gooch JL, Oesterreich S, et al: Insulin-like growth factor I-induced degradation of insulin receptor substrate 1 is mediated by the 26S proteasome and blocked by phosphatidylinositol 3'-kinase inhibition. Mol Cell Biol 20:1489-1496, 2000

    Zhang H, Hoff H, Sell C: Insulin-like growth factor I-mediated degradation of insulin receptor substrate-1 is inhibited by epidermal growth factor in prostate epithelial cells. J Biol Chem 275:22558-22562, 2000

    Torres-Arzayus MI, De Mora JF, Yuan J, et al: High tumor incidence and activation of the PI3K/AKT pathway in transgenic mice define AIB1 as an oncogene. Cancer Cell 6:263-274, 2004(Xiaojiang Cui, Rachel Sch)