Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-26T02:30:52.263Z Has data issue: false hasContentIssue false

Role of plasma membrane ER protein in breast cancer

Published online by Cambridge University Press:  17 February 2006

E. M. Rosen
Affiliation:
Department of Oncology, Lombardi Comprehensive Cancer Center, Georgetown University, NW, Washington, DC, USA.
S. Fan
Affiliation:
Department of Oncology, Lombardi Comprehensive Cancer Center, Georgetown University, NW, Washington, DC, USA.
Y. Ma
Affiliation:
Department of Oncology, Lombardi Comprehensive Cancer Center, Georgetown University, NW, Washington, DC, USA.
E. R. Levin
Affiliation:
Division of Endocrinology, Veterans Affairs Medical Center, California, CA, USA.

Abstract

The role of estradiol and the estrogen receptor (ER-α) in the etiology of breast cancer have long been appreciated. This understanding has been complicated by two discoveries in the 1990s: (1) a second estrogen receptor (ER-β) whose expression pattern and activity overlap with but are distinct from those of ER-α; and (2) a pool of ERs located at the plasma membrane. This plasma membrane-localized ER constitutes a distinct pool of receptors whose protein interactions, signaling mechanisms, and cellular functions are not the same as that of the cytoplasmic- and nuclear-localized ER and are not as well understood. Here, we will consider the structure and function of the membrane-localized ER protein. We will then discuss what is known about the role of the membrane ER in the development and its implications for breast cancer treatment.

Type
Focus On
Copyright
2006 Cambridge University Press

Introduction

ER-α (ESRA), a member of the nuclear receptor superfamily, is a ligand-activated transcription factor that contains domains for DNA binding, transcriptional activation, and hormone (17β-estradiol, E2) binding. The full-length ER-α is a 595 amino acid, 66-kDa protein. ER-β (ESRB), which is encoded by a separate gene, was identified and characterized in 1996 [1,2]. The DNA-binding and ligand-binding domains of ER-β show a high degree of identity to ER-α; while the N-terminal activation function (AF-1), hinge, and F (C-terminal) domains are not conserved. A third receptor (ER-γ) was identified in teleost fish [3]; but a mammalian homolog has not been found. Early evidence of the existence of membrane-associated E2 receptors that can transduce rapid signaling events through G-protein activation has been reviewed elsewhere [4]. In the last few years, new data have emerged on the structure, function, and potential physiologic importance of membrane ERs.

Characterization of plasma membrane ER

While the presence of high-affinity cell-surface binding sites for E2 was known in the 1970s [5], structural characterization of these sites has only recently been achieved. Thus, expression of exogenous ER-α or ER-β in Chinese hamster ovary cells resulted in expression of ER in the nucleus and cell membrane [6]. The abundance of membrane ER was 2–3% of that of nuclear ER, but the dissociation constants (Kds) were similar. Both membrane ERs were able to activate G proteins (Gαq and Gαs), generate cAMP (via Gαs), stimulate inositol triphosphate (IP3) production and calcium influx (via Gαq), and induce extracellular signal-regulated kinase (ERK) signaling and cell proliferation [6]. About 5% of the endogenous ER-α localizes to the cell membrane in MCF-7, an E2-responsive breast cancer cell line, suggesting a role for the membrane ER in breast cancer. Recently, a G protein-coupled receptor, GPR30, was found to localize to the endoplasmic reticulum, bind E2, and induce calcium mobilization and IP3 synthesis in response to E2 [7].

The mechanism(s) by which ER-α localizes to the cell membrane is uncertain, since it does not contain a classical membrane localization signal. Serine-522 in the E-domain (ligand-binding domain) is required for membrane insertion, as is caveolin, a scaffold protein [8]. While full-length ER-α probably comprises most membrane ER, the E-domain is sufficient for insertion into the membrane. Membrane localization of ER-α also requires palmitoylation of cysteine-447. Mutation of this site blocked palmitoylation of ER-α, association with caveolin, membrane localization, rapid signaling, and proliferation in response to E2 [9]. Dimerization is required for membrane ER-α to mediate the rapid non-genomic actions of E2, but not for its insertion into the membrane [10].

Recent studies have identified an N-terminally truncated ER-α (46 kDa) in vascular endothelial cells that is generated by alternative splicing and recruited to the plasma membrane by palmitoylation [11,12]. ER46 transduces membrane-initiated E2 responses, including activation of eNOS (NOS3, endothelial nitric oxide synthetase), consistent with an earlier study showing that endogenous plasma membrane ER-α activates GαI, leading to synthesis of nitric oxide [13]. Whether ER46 is expressed on the membranes of other cell types is unclear at present.

Signaling from membrane ER

One signaling mechanism of membrane located G protein-coupled ERs involves cross-activation of growth factor receptor signaling pathways, including those of the epidermal growth factor receptor (EGFR) and the insulin-like growth factor 1 receptor (IGF1R) [1416]. In a recent study, E2-stimulated ERK activation in breast cancer and vascular endothelial cells required both ER-α and EGFR activation, which was mediated by the rapid release of the heparin-binding EGF-like growth factor (HBEGF) in MCF-7 cells [17]. Other events in this pathway were identified, including its dependence upon several G proteins (Gαq, Gαi, and Gβγ) and the role of SRC-mediated activation of several matrix metalloproteinases (MMP2 and MMP9) in E2-induced HBEGF release and activation of several protein kinases (ERK, c-Akt, and p38β MAP kinase (SAPK2B)) [17].

Shc is a ubiquitous signaling adapter protein containing Src homology 2 and 3 (SH2 and SH3) domains. In MCF-7 cells, E2 causes the rapid association of Shc with ER-α and their translocation to the plasma membrane. This process involves E2-induced phosphorylation of IGF1R and formation of a ternary protein complex of ER-α, Shc, and IGF1R at the membrane [18,19]. The E2-induced association of ER-α with IGF1R and membrane localization of ER-α required Shc; and all three components were necessary for E2-induced ERK phosphorylation. The ability of membrane ER-α to transduce EGFR and IGF1R signaling (leading to ERK activation and cell proliferation) has other implications. Thus, EGF and IGF1 can activate unliganded ER-α via ERK-mediated phosphorylation of serine-118 on the AF-1 domain [19,20], although other kinases may mediate the E2-induced phosphorylation of serine-118 [21]. c-Akt, another target of growth factor signaling, phosphorylates ER-α on serine-167 of AF-1; and this event may contribute to Tamoxifen resistance [22]. Coregulator proteins are also phosphorylation targets. Thus, E2 rapidly induces phosphorylation of the coactivator AIB1 (amplified in breast cancer 1) in an ER-dependent manner [23]. The p160 family coactivator GRIP1 is phosphorylated by ERK at serine-736, an event that is required for growth factor induction of GRIP1 coactivator function [24]. These findings suggest that growth factor signaling initiated by membrane ER may activate the nuclear ER-α via phosphorylation of nuclear ER-α or its coactivators [25], allowing the membrane ER to regulate both transcriptional and non-genomic actions of E2.

Role of membrane ER in breast cancer

Previous studies revealed that the tumor suppressor protein encoded by the breast cancer susceptibility gene 1 (BRCA1) interacts directly with ER-α and inhibits its transcriptional activity and estrogen-responsive gene expression [2629]. In ER-α positive breast cancer cells (MCF-7 and ZR-75-1), E2 caused activation of ERK and cell proliferation that was blocked by exogenous BRCA1 or inhibition of ERK signaling [30]. BRCA1 also inhibited EGF-induced ERK activity and cell proliferation that was mediated, in part, through the mitogen-activated kinase phosphatase 1 (MKP1). These findings suggest that the ability of BRCA1 to inhibit E2-stimulated breast cancer cell proliferation is due, in part, to inhibition of membrane ER-α signaling.

A role for the membrane ER in breast cancer response to hormonal manipulation has been postulated. Thus, it was suggested that membrane ER-α contributes to estrogen hypersensitivity in women who relapse following oophorectomy [31]. These patients often respond to aromatase inhibitors, which block peripheral conversion of androgens to estrogen. In the setting of long-term E2 deprivation, ER-α is up-regulated, as are growth factor pathways that mediate the rapid non-genomic effects of E2 (including those involving ERK, phosphatidylinositol-3-kinase (PI3K), mammalian target of rapamycin (MTOR), and several signaling adaptor proteins (Shc, Grb2, and Sos)) [31]. These are some of the same components that mediate membrane ER-α signaling, suggesting that the membrane ER contributes to the E2 hypersensitivity. A similar mechanism may mediate relapse in patients treated with Tamoxifen. In this respect, it has been reported that resistance to Tamoxifen-induced apoptosis is mediated by HER2/Neu and the plasma membrane ER [32].

Just as membrane ERs can mediate cardioprotective and neuroprotective effects of E2, they may mediate survival of breast cancer cells. Thus, E2-blocked chemotherapy and radiation-induced apoptosis in ER-positive breast cancer cells through stimulation of ERK and inhibition of c-Jun N-terminal kinase (JNK) activity by the membrane ER-α [33]. Expression of the metastasis-associated gene 1 (MTA1) is associated with aggressive breast cancer. Interestingly, a short form of MTA1 (MTA1s) sequesters ER-α in the cytoplasm and represses E2-induced transcriptional activity, while promoting its non-genomic responses [34]. These findings suggest that cell membrane (or at least non-nuclear) ER-α signaling may promote the malignant phenotype of breast cancer.

The presence of ER-α and ER-β in plasma membrane caveolae from cultured lung carcinoma cells was described recently; and ER-α immunostaining was detected at the cell membrane in archival human breast and lung tumor samples, based on confocal microscopy [35]. Consistent with a plasma membrane ER-α in lung cancer cells, E2 induced ERK signaling and lung cancer cell proliferation that was blocked by Faslodex (ICI 182,780), a pure anti-estrogen [35]. A recent study compared the properties of stably integrated wild-type nuclear ER-α vs. a modified membrane-targeted ER-α in originally ER-negative MDA-MB-231 breast cancer cells. Unlike the nuclear ER-α, the membrane receptor expression was not decreased by E2 or Faslodex; and the ability to regulate ERK activity, c-Akt activity, and cell proliferation differed between the nuclear vs. membrane receptors [36]. Interestingly, an inverse correlation between EGFR and ER-α levels in human breast cancers has been described [37,38]. This correlation may be mediated through an E2-sensitive negative regulatory element within first intron of the EGFR gene. It has been postulated that increased EGFR expression may contribute to the aggressive behavior and poor prognosis of ER-negative breast cancers.

Perspectives

Figure 1 shows a model for cross-talk between plasma membrane ER-α, EGFR, IGF1R, and nuclear ER-α that may occur in human breast cancer cells and result in stimulation of proliferation and inhibition of apoptosis. The studies described herein suggest potential roles for the cell membrane ER-α in the development and progression of breast cancer, including a role in relapse following hormonal therapy. Studies of membrane ER structure and function are hindered by the low abundance of ER protein localized at the membrane and the difficulty in isolating effects due to membrane vs. nuclear ER pools in the same cell. The development of more specific reagents to investigate the cell membrane ER in vitro and in vivo and to identify functions of other intracellular receptor pools (e.g. mitochondrial and endoplasmic reticulum) should better position us to: (1) understand the physiologic roles of these receptors; and (2) selectively target extranuclear ER for cancer treatment.

Figure 1. Schematic illustration of signaling pathways for cell membrane-localized ER-α in breast cancer cells. The ligated plasma membrane ER-α (shown here as a homodimer) signals through several pathways, involving IGF1R, G proteins, and EGFR. Activation of growth factor signaling pathways through several kinases (including c-Akt, ERK, and others (e.g. JNK)) lead to the following consequences: (1) stimulation of cell proliferation (which is modulated by BRCA1 and MKP1); (2) inhibition of apoptosis; and (3) stimulation of transcription by the nuclear ER-α. Activation of nuclear ER-α is mediated by phosphorylation of nuclear receptor coactivators (e.g. AIB1, GRIP1, and CBP (CREB-binding protein)) or, more directly, by phosphorylation of ER-α itself. Abbreviations: ERE: estrogen response element; MEK: MAPK/ERK kinase; Pi: inorganic phosphate; PKA: protein kinase A. Other abbreviations, see text.

Acknowledgements

Dr Rosen has been supported, in part, by USPHS grants R01-CA82599, R01-CA80000, R01-CA104546, and RO1-ES09169 and by a grant from the Susan G. Komen Breast Cancer Foundation (BCTR0503799).

References

Mosselman S, Polman J, Dijkema R. ER-beta: identification and characterization of a novel human estrogen receptor. FEBS Lett 1996; 392: 4953.Google Scholar
Kuiper GG, Enmark E, Pelto-Huikko M, et al. Cloning of a novel estrogen receptor expressed in rat prostate and ovary. Proc Natl Acad Sci USA 1996; 93: 59255930.Google Scholar
Hawkins MB, Thornton JW, Crews D, et al. Identification of a third distinct estrogen receptor and reclassification of estrogen receptors in teleosts. Proc Natl Acad Sci USA 2000; 97: 10 751–10 756.Google Scholar
Levin ER. Cell localization, physiology, and nongenomic actions of estrogen receptors. J Appl Physiol 2001; 91: 18601867.Google Scholar
Pietras R, Szego CM. Specific binding sites for oestrogen at the outer surfaces of isolated endometrial cells. Nature 1977; 265: 6972.Google Scholar
Razandi M, Pedram A, Greene GL, Levin ER. Cell membrane and nuclear estrogen receptors derive from a single transcript: studies of ERα and ERβ expressed in CHO cells. Mol Endocrinol 1999; 13: 307319.Google Scholar
Revankar CM, Cimino DF, Sklar LA, et al. A transmembrane intracellular estrogen receptor mediates rapid cell signaling. Science 2005; 307: 16251630.Google Scholar
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 2003; 23: 16331646.Google Scholar
Acconcia F, Ascenzi P, Bocedi A, et al. Palmitoylation-dependent estrogen receptor alpha membrane localization: regulation by 17beta-estradiol. Mol Biol Cell 2005; 16: 231237.Google Scholar
Razandi M, Pedram A, Merchenthaler I, et al. Plasma membrane estrogen receptors exist and functions as dimmers. Mol Endocrinol 2004; 18: 28542865.Google Scholar
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 USA 2003; 100: 48074812.Google Scholar
Flouriot G, Brand H, Denger S, et al. Identification of a new isoform of the human estrogen receptor-alpha (hER-alpha) that is encoded by distinct transcripts and that is able to repress hER-alpha activation function 1. EMBO J 2000; 19: 46884700.Google Scholar
Wyckoff MH, Chambliss KL, Mineo C, et al. Plasma membrane estrogen receptors are coupled to endothelial nitric-oxide synthase through G alpha (i). J Biol Chem 2001; 276: 27 07127 076.Google Scholar
Filardo EJ, Quinn JA, Frackelton Jr AR, Bland KI. Estrogen action via the G protein-coupled receptor, GPR30: stimulation of adenylyl cyclase and cAMP-mediated attenuation of the epidermal growth factor receptor-to-MAPK signaling axis. Mol Endocrinol 2002; 16: 7084.Google Scholar
Filardo EJ, Quinn JA, Bland KI, Frackelton Jr AR. Estrogen-induced activation of Erk-1 and Erk-2 requires the G protein-coupled receptor homolog, GPR30, and occurs via trans-activation of the epidermal growth factor receptor through release of HB-EGF. Mol Endocrinol 2000; 14: 16491660.Google Scholar
Song RX, Barnes CJ, Zhang Z, et al. The role of Shc and insulin-like growth factor 1 receptor in mediating the translocation of estrogen receptor alpha to the plasma membrane. Proc Natl Acad Sci USA 2004; 101: 20762081.Google Scholar
Razandi M, Pedram A, Park ST, Levin ER. Proximal events in signaling by plasma membrane estrogen receptors. J Biol Chem 2003; 278: 27012712.Google Scholar
Zhang Z, Kumar R, Santen RJ, Song RX. The role of adapter protein Shc in estrogen non-genomic action. Steroids 2004; 69: 523529.Google Scholar
Kato S, Endoh H, Masuhiro Y, et al. Activation of the estrogen receptor through phosphorylation by mitogen-activated protein kinase. Science 1995; 270: 14911494.Google Scholar
Bunone G, Briand PA, Miksicek RJ, Picard D. Activation of the unliganded estrogen receptor by EGF involves the MAP kinase pathway and direct phosphorylation. EMBO J 1996; 15: 21742183.Google Scholar
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 2002; 21: 49214931.Google Scholar
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 2001; 276: 98179824.Google Scholar
Zheng FF, Wu RC, Smith CL, O'Malley BW. Rapid estrogen-induced phosphorylation of the SRC-3 coactivator occurs in an extranuclear complex containing estrogen receptor. Mol Cell Biol 2005; 25: 82738284.Google Scholar
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 2001; 276: 22 17722 182.Google Scholar
Levin ER. Bidirectional signaling between the estrogen receptor and the epidermal growth factor receptor. Mol Endocrinol 2003; 17: 309317.Google Scholar
Fan S, Wang J, Yuan R, et al. BRCA1 inhibition of estrogen receptor signaling in transfected cells. Science 1999; 284: 13541356.Google Scholar
Fan S, Ma YX, Wang C, et al. p300 Modulates the BRCA1 inhibition of estrogen receptor activity. Cancer Res 2002; 62: 141151.Google Scholar
Xu J, Fan S, Rosen EM. Regulation of the estrogen-inducible gene expression profile by the breast cancer susceptibility gene BRCA1. Endocrinology 2005; 146: 20312047.Google Scholar
Ma YX, Tomita Y, Fan S, et al. Structural determinants of the BRCA1: estrogen receptor interaction. Oncogene 2005; 24: 18311846.Google Scholar
Razandi M, Pedram A, Rosen EM, Levin ER. BRCA1 inhibits membrane estrogen and growth factor receptor signaling to cell proliferation in breast cancer. Mol Cell Biol 2004; 24: 59005913.Google Scholar
Santen RJ, Song RX, Zhang Z, et al. Long-term estradiol deprivation in breast cancer cells up-regulates growth factor signaling and enhances estrogen sensitivity. Endocr Relat Cancer 2005; 12 (Suppl 1): S61S73.Google Scholar
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 2002; 97: 306312.Google Scholar
Razandi M, Pedram A, Levin ER. Plasma membrane estrogen receptors signal to antiapoptosis in breast cancer. Mol Endocrinol 2000; 14: 14341447.Google Scholar
Kumar R, Wang E-A, Mazumdar A, et al. A naturally occurring MTA1 variant sequesters oestrogen receptor-alpha in the cytoplasm. Nature 2002; 418: 654657.Google Scholar
Pietras RJ, Marquez DC, Chen HW, et al. Estrogen and growth factor receptor interactions in human breast and non-small cell lung cancer cells. Steroids 2005; 70: 372381.Google Scholar
Rai D, Frolova A, Frasor J, et al. Distinctive actions of membrane-targeted versus nuclear localized estrogen receptors in breast cancer cells. Mol Endocrinol 2005; 19: 16061617.Google Scholar
deFazio A, Chiew YE, Sini RL, et al. Expression of c-erbB receptors, heregulin and oestrogen receptor in human breast cell lines. Int J Cancer 2000; 87: 487498.Google Scholar
Wilson MA, Chrysogelos SA. Identification and characterization of a negative regulatory element within the epidermal growth factor receptor gene first intron in hormone-dependent breast cancer cells. J Cell Biochem 2002; 85: 601614.Google Scholar
Figure 0

Schematic illustration of signaling pathways for cell membrane-localized ER-α in breast cancer cells. The ligated plasma membrane ER-α (shown here as a homodimer) signals through several pathways, involving IGF1R, G proteins, and EGFR. Activation of growth factor signaling pathways through several kinases (including c-Akt, ERK, and others (e.g. JNK)) lead to the following consequences: (1) stimulation of cell proliferation (which is modulated by BRCA1 and MKP1); (2) inhibition of apoptosis; and (3) stimulation of transcription by the nuclear ER-α. Activation of nuclear ER-α is mediated by phosphorylation of nuclear receptor coactivators (e.g. AIB1, GRIP1, and CBP (CREB-binding protein)) or, more directly, by phosphorylation of ER-α itself. Abbreviations: ERE: estrogen response element; MEK: MAPK/ERK kinase; Pi: inorganic phosphate; PKA: protein kinase A. Other abbreviations, see text.