Skip to main content
Erschienen in: Breast Cancer Research and Treatment 3/2020

Open Access 04.03.2020 | Preclinical study

G1T48, an oral selective estrogen receptor degrader, and the CDK4/6 inhibitor lerociclib inhibit tumor growth in animal models of endocrine-resistant breast cancer

verfasst von: Kaitlyn J. Andreano, Suzanne E. Wardell, Jennifer G. Baker, Taylor K. Desautels, Robert Baldi, Christina A. Chao, Kendall A. Heetderks, Yeeun Bae, Rui Xiong, Debra A. Tonetti, Lauren M. Gutgesell, Jiong Zhao, Jessica A. Sorrentino, Delita A. Thompson, John E. Bisi, Jay C. Strum, Gregory R. J. Thatcher, John D. Norris

Erschienen in: Breast Cancer Research and Treatment | Ausgabe 3/2020

Abstract

Purpose

The combination of targeting the CDK4/6 and estrogen receptor (ER) signaling pathways with palbociclib and fulvestrant is a proven therapeutic strategy for the treatment of ER-positive breast cancer. However, the poor physicochemical properties of fulvestrant require monthly intramuscular injections to patients, which limit the pharmacokinetic and pharmacodynamic activity of the compound. Therefore, an orally available compound that more rapidly reaches steady state may lead to a better clinical response in patients. Here, we report the identification of G1T48, a novel orally bioavailable, non-steroidal small molecule antagonist of ER.

Methods

The pharmacological effects and the antineoplastic mechanism of action of G1T48 on tumors was evaluated using human breast cancer cells (in vitro) and xenograft efficacy models (in vivo).

Results

G1T48 is a potent and efficacious inhibitor of estrogen-mediated transcription and proliferation in ER-positive breast cancer cells, similar to the pure antiestrogen fulvestrant. In addition, G1T48 can effectively suppress ER activity in multiple models of endocrine therapy resistance including those harboring ER mutations and growth factor activation. In vivo, G1T48 has robust antitumor activity in a model of estrogen-dependent breast cancer (MCF7) and significantly inhibited the growth of tamoxifen-resistant (TamR), long-term estrogen-deprived (LTED) and patient-derived xenograft tumors with an increased response being observed with the combination of G1T48 and the CDK4/6 inhibitor lerociclib.

Conclusions

These data show that G1T48 has the potential to be an efficacious oral antineoplastic agent in ER-positive breast cancer.
Hinweise

Electronic supplementary material

The online version of this article (https://​doi.​org/​10.​1007/​s10549-020-05575-9) contains supplementary material, which is available to authorized users.
Kaitlyn J. Andreano and Suzanne E. Wardell contributed equally to this work.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

The estrogen receptor (ER/ESR1) is expressed in a majority of breast cancers, and drugs that inhibit ER signaling are the cornerstone of breast cancer pharmacotherapy for ER-positive/HER2-negative disease [1]. These targeted approaches include the Selective Estrogen Receptor Modulator (SERM) tamoxifen that acts as a competitive ER antagonist in the breast, and aromatase inhibitors (AIs) that inhibit aromatase, the enzyme responsible for estrogen production [2, 3]. However, the development of resistance limits the duration of meaningful therapeutic responses. Mechanisms of resistance to these endocrine therapies include cell cycle dysregulation and activation of alternative growth factor signaling pathways [1]. For example, activation of MAPK, PI3K, and GSK-3 can result in increased phosphorylation of ER or its attendant coregulatory proteins leading to ligand-independent ER activity and resistance [47]. Recently, genomic alterations in the ER gene itself, including amplification, translocation, and ligand binding domain mutations (most frequently ER-D538G and ER-Y537S) have emerged with AI therapy [1, 810].
After progression during tamoxifen and AI therapy, other endocrine treatments including the steroidal selective estrogen receptor downregulator (SERD) fulvestrant (Faslodex®) are generally used [11]. SERDs are a class of ER antagonists that in addition to competitively displacing estrogens, also trigger ER downregulation [12]. Although initially successful, the onset of resistance limits durable responses when used as a monotherapy. Therefore, in an effort to improve the therapeutic lifespan of endocrine treatments for metastatic breast cancer, combination regimens have been extensively studied. Clinical trials using a combination of AI or fulvestrant with pan-PI3K or mTOR inhibitors have been promising but inconclusive, and toxicity often remains an impediment to dose escalation [1316]. Therefore, CDK4/6 inhibitors have emerged as a favored option when considering combination endocrine therapies [1720]. However, the poor bioavailability of fulvestrant, coupled with its intramuscular route of administration and the long time to steady state blood levels, compromises its clinical use [21, 22]. Indeed, even at the higher clinical dose (500 mg) of fulvestrant, pharmacodynamic imaging suggests incomplete receptor saturation [23].
Collectively, these data highlight an unmet need for a safe, orally bioavailable SERD with appropriate pharmaceutical properties. Herein we describe the preclinical development of G1T48 (rintodestrant), an orally bioavailable, potent, and selective non-steroidal ER antagonist and downregulator [24]. G1T48 was found to robustly inhibit ER activity in multiple in vitro models of endocrine therapy resistance, including those harboring ER mutations or growth factor activation. Importantly, G1T48 demonstrated robust antitumor activity in an animal model of early stage estrogen-dependent breast cancer and suppressed the growth of tamoxifen- and estrogen deprivation-resistant xenograft tumors with increased efficacy observed for the combination of G1T48 and lerociclib, a newly developed CDK4/6 inhibitor [25, 26].

Methods

Reagents

Fulvestrant (CAS No: 129453–61-8, > 99% purity) was purchased from MedChem Express. Estradiol (E2) (E8875), lasofoxifene (SML1026), 4-hydroxytamoxifen (H7904), and tamoxifen (T5648) were purchased from Sigma. Raloxifene (2280) was purchased from Tocris. GDC-0810 (S7855), bazedoxifene (S2128), and AZD9496 (S8372) were purchased from Selleckchem. GW5638 (5638), GW7604 (7604), and RU 58,668 (RU) were provided by Donald McDonnell (Duke University). G1T48 was provided by G1 Therapeutics, Inc., as analytical grade compound.

RNA analysis

MCF7 cells were authenticated by short tandem repeat profiling, were tested for Mycoplasma and were not cultured for more than three months at a time [27]. MCF7 cells were plated in DMEM/F12 supplemented with 8% charcoal dextran treated FBS for 48 h. Cells were then treated for 24 h with ligand and RNA was isolated using the Aurum™ total RNA isolation kit (Bio-Rad, Hercules, CA). After cDNA synthesis (iScript kit, Bio-Rad) real-time PCR was performed using the Bio-Rad CFX384 real-time system. GAPDH mRNA expression was used to normalize all real-time data using the 2-ΔΔCT method [28]. For more detailed description of this method, please see Online Resource 1.

Proliferation

MCF7 cells were plated in DMEM/F12 supplemented with 8% charcoal dextran treated FBS in 96-well plates (5 K cells/well) for 48 h. Cells were treated with estradiol (0.1 nM) or insulin (20 μM) with or without test compound (dose response; 1.0–11 to 1.0–05 M) for 6 days. Plates were decanted and frozen at – 80°°C overnight prior to quantitation of DNA by fluorescence using Hoechst 33258.

Supplementary material

Detailed methods are available in Online Resource 1 for the following protocols: In-Cell Western, Radioactive Binding Assay, Chromatin Immunoprecipitation, Transcriptional Reporter Assays.

Murine studies

All procedures were approved by the Institutional Animal Care and Use Committee (IACUC) of Duke University or South Texas Accelerated Research Therapeutics (START, San Antonio, Texas) prior to initiating the experiment. For complete details, see Online Resource 1.

Results

G1T48 is similar to fulvestrant in its ability to downregulate the estrogen receptor and inhibit estrogen signaling in breast cancer cells

Novel ER antagonists with SERD activity have recently been described, but clinical development of these compounds has thus far been limited due to unanticipated side effects or for undisclosed reasons [2936]. We sought to identify an orally bioavailable SERD using the chemical backbone of raloxifene as a starting point, since this SERM has demonstrated a favorable safety profile in the clinical setting of breast cancer prevention and osteoporosis treatment [37, 38]. G1T48 incorporates an acrylic acid side chain (Fig. 1a) [29, 31, 32, 34, 39, 40], and was the product of structure-guided investigations, driven by activity in breast cancer cell lines [24]. G1T48 was first assessed for its ability to downregulate ER when compared to several benchmark SERMs and SERDs including fulvestrant [12, 41]. Using In-Cell Western assays, G1T48 was found to downregulate ER with an efficacy modestly more potent than steroidal and other SERDs (e.g., fulvestrant, AZD9496; approximately 10% ER remaining after treatment) (Fig. 1b, online resource 2). Bazedoxifene (BZA), raloxifene (RAL), tamoxifen, 4-hydroxytamoxifen (4OHT), and lasofoxifene (laso) were also found to partially downregulate ER. These data demonstrate that in vitro G1T48 is a pure antiestrogen and selective estrogen receptor degrader (PA-SERD).
We next evaluated the ability of G1T48 to inhibit endogenous ER target gene transcription in MCF7 cells. As shown in Fig. 2a, G1T48 suppressed estrogen-mediated activation of the Trefoil Factor-1 (TFF1) mRNA similarly to fulvestrant and additional antiestrogens (4OHT, GDC-0810, AZD9496, RAL). The biochemical basis of G1T48-mediated ER antagonism was further evaluated using 3H-estradiol whole-cell competition assay. Results showed that G1T48 displaced radiolabeled agonist binding with potency greater than fulvestrant and similar to BZA (Fig. 2b). Radioligand binding assay (RBA) IC50 shows that G1T48 is a competitive ER antagonist (Online Resource 3).
The inability of some ER antagonists, notably SERMs, to completely oppose the actions of estradiol is seen as a liability when being considered for the treatment of advanced therapy-resistant breast cancer. While data from Fig. 1b confirm G1T48 is a SERD, the potential remains for SERM activity, as G1T48 was developed based on compounds that exhibit both SERM and SERD activity. Therefore, we next considered the impact of G1T48 treatment on ER target genes that are differentially regulated by SERMs and SERDs [42]. As shown in Fig. 2c, compounds with SERM activity regulate these genes in a manner similar to the agonist estradiol, a reflection of their intrinsic agonist potential (red, green, and blue clusters). In contrast, G1T48 regulates these genes in a pattern that is consistent with compounds previously shown to downregulate ER (e.g., GW7604, fulvestrant, GW5638, RU 565899, GDC-0810, and AZD9496; orange, teal and purple clusters).
When bound by estrogen, ER is recruited to target gene promoters to activate or repress target gene transcription through recruitment of coregulator (coactivator or corepressor) proteins that modify chromatin structure [43]. We assessed the ability of G1T48 and benchmark SERMs or SERDs to inhibit estrogen-mediated recruitment of ER to the TFF1 promoter using chromatin immunoprecipitation (ChIP) assays. While estrogen and 4OHT treatment significantly increased ER binding to the TFF1 promoter (Fig. 2d), G1T48 inhibited the binding of ER to this promoter, with or without estrogen, with efficacy similar to fulvestrant, supporting the idea that G1T48 is an efficient ER antagonist in vitro.
We next evaluated the ER selectivity of G1T48 by assessing its ability to inhibit the transcriptional activity of related steroid hormone receptors androgen receptor (AR), progesterone receptor (PR), glucocorticoid receptor (GR), and mineralocorticoid receptor (MR) using a cell-based reporter assay. When administered at doses up to 10 µM, G1T48 did not affect the transcriptional activities of AR, GR, MR, or PR (Online Resource 4), indicating that G1T48 is a highly selective antagonist of ER.

G1T48 inhibits the growth of ER-positive breast cancer cells

To examine the therapeutic potential of G1T48, we performed cell proliferation assays using multiple ER-positive breast cancer cell lines (Fig. 3). G1T48 significantly inhibited estrogen-mediated growth of MCF7 cells demonstrating approximately threefold higher potency when compared to fulvestrant (Fig. 3a, Online Resource 5). Additionally, G1T48 and benchmark antiestrogens also inhibited the estrogen-mediated growth of ER-positive BT474 and ZR-75-1 breast cancer cells, while no growth inhibition was observed in ER-negative MDA-MB-436 breast cancer cells (Fig. 3, Online Resource 5). Furthermore, G1T48 does not impact apoptosis in MCF7 breast cancer cells (Online Resource 6). Thus, G1T48 selectively inhibits the growth of ER-positive, but not ER-negative, breast cancer cells.

G1T48 inhibits estrogen signaling in endocrine-resistant breast cancer models

In addition to the upregulation of growth factor signaling, a second key mechanism of resistance to aromatase inhibition is newly described mutations in the ligand binding domain of ER [8, 9], mutations that result in reduced potency for 4OHT and fulvestrant as compared to wild-type (wt) receptor [4450]. To assess the activity of G1T48 on endocrine refractory ER mutants, we utilized a reporter gene assay in ER-negative SKBR3 breast cancer cells transfected with ER expression vectors (wtER or the two most common ER mutants, ER-Y537S or ER-D538G) (Fig. 4a). G1T48 was found to be a potent and effective inhibitor of both wtER and ER-D538G transcription. As has been previously reported [49], all antiestrogens tested, including G1T48, demonstrated reduced potency against ER mutant transcriptional activity when compared to wtER (Online Resource 7). To further understand the significance of these results, MCF7 cells expressing doxycycline inducible wtER, ER-D538G, or ER-Y537S were engineered and G1T48 was evaluated for its ability to inhibit the ER-dependent growth of these cells. As has been previously reported for 4OHT and fulvestrant (and also confirmed here), G1T48 exhibited an increased GI50 in cells expressing the ER-Y537S and ER-D538G mutations when compared to wtER (Fig. 4b, Online Resource 7) [44, 45, 4751]. Collectively, these data highlights that G1T48, like other SERDs, may be useful at targeting some mutant receptors, a potential that can be further evaluated clinically.
Dysregulated growth factor signaling has emerged as a primary mechanism of resistance to tamoxifen and AI therapy [10]. Activation of these signaling pathways can alter the pharmacology of compounds like tamoxifen, converting them from antagonists to agonists through phosphorylation of ER or its attendant coregulator proteins [47]. G1T48 and comparator SERMs and SERDs were evaluated for their ability to inhibit insulin-mediated MCF7 cell proliferation, a model for endocrine therapy resistance. Compounds with SERD activity, including G1T48, effectively blocked growth factor-mediated cell growth, while compounds with SERM activity (e.g., 4OHT) were less effective (Fig. 4c). These data together support the potential for G1T48 to have efficacy in the treatment of AI or tamoxifen-resistant breast cancers having growth factor pathway activation.

Evaluation of the in vivo therapeutic efficacy of the SERD G1T48 and the CDK4/6 inhibitor lerociclib using estrogen-dependent and tamoxifen-resistant (TamR) breast cancer xenograft models

We next evaluated the therapeutic potential of G1T48 in ER-positive primary and endocrine refractory breast cancer models in vivo. G1T48 was first assessed, as a monotherapy or in combination with the CDK4/6 inhibitor lerociclib, for its impact on the growth of naïve MCF7 xenograft tumors (Fig. 5a). Ovariectomized estrogen-treated female nu/nu mice bearing MCF7 xenograft tumors were randomized to treatment with vehicle, lerociclib (50 mg/kg), and/or G1T48 (30 or 100 mg/kg). G1T48 treatment demonstrated dose-dependent repression of tumor growth. Combination of lerociclib and G1T48 was more effective than either monotherapy, demonstrating an added benefit to using these agents together. End of study tumor volumes are presented in Online Resource 8.
The TamR xenograft model exhibits tamoxifen-stimulated growth that can be inhibited by compounds with SERD activity with added benefit observed upon combination with CDK4/6 inhibitors [52]. Therefore, ovariectomized tamoxifen-treated mice bearing TamR xenografts were randomized to treatment with lerociclib (50 mg/kg or 100 mg/kg), G1T48 (30 mg/kg or 100 mg/kg), fulvestrant (200 mg/kg), or CDK4/6 inhibitor palbociclib (100 mg/kg) as monotherapies or a combination of lerociclib (50 mg/kg) and G1T48 (30 or 100 mg/kg). In this model system, lerociclib demonstrated efficacy equivalent to that of the mechanistic clinical comparator palbociclib (Fig. 5b). G1T48 was found to demonstrate dose-dependent inhibition of TamR tumor growth (Fig. 5c) albeit with less efficacy than fulvestrant. Interestingly, G1T48 treatment resulted in greater downregulation of intratumoral ER levels than fulvestrant despite less efficient inhibition of tumor growth (Online Resource 8). Finally, combination of G1T48 with lerociclib, using suboptimal doses of each inhibitor, resulted in tumor growth inhibition significantly greater than that observed for either compound as monotherapy (Fig. 5d). End of study tumor volumes are presented in Online Resource 10. Mouse body weight was largely unaffected by the treatment regimens in this study as presented in Online Resource 11.

Evaluation of the combined efficacy of lerociclib and G1T48 in a xenograft tumor model of resistance to estrogen deprivation in vivo

Although AIs have seen rapid adoption in the adjuvant setting, de novo and acquired resistance remains a persistent impediment to sustained clinical responses. We have developed an ER-positive model of aromatase resistance, termed long-term estrogen-deprived (LTED), to model this clinical situation [53]. In order to evaluate the combined efficacy of lerociclib and G1T48 in this model system, LTED xenograft tumors were orthotopically established in ovariectomized female nu/nu mice. When tumors measured 0.1–0.15 cm3 volume, G1T48 (5 mg/kg or 100 mg/kg) and lerociclib (50 mg/kg or 100 mg/kg) were administered alone and in combination, with fulvestrant (25 mg/kg) and palbociclib (100 mg/kg) included for comparison. As previously observed with the MCF7 parental and TamR models, G1T48 demonstrated dose-dependent inhibition of tumor growth (Fig. 6a). Additionally, the tumor growth inhibition after treatment with G1T48 and lerociclib was comparable to their clinical comparators (fulvestrant and palbociclib, respectively) (Fig. 6b). Combination of G1T48 with lerociclib suppressed tumor growth significantly compared to monotherapy (Fig. 6c) and resulted in tumor regression for a majority of tumors receiving the combined therapy regimen. End of study tumor volumes are presented in Online Resource 12.

Evaluation of the combined efficacy of lerociclib and G1T48 in a patient-derived xenograft model harboring the ER-Y537S Mutation

As described above, mutations in the LBD of ESR1 are an emerging mechanism of resistance to AIs. The efficacy of G1T48, as a mono- and combination therapy with lerociclib, was evaluated using a Patient-Derived Xenograft (PDX model) harboring the ER-Y537S mutation (Fig. 7a). Female athymic nu/nu mice were implanted with the ST2177 LUMB PDX tumor [33] and following treatment with G1T48, a dose-dependent decrease in tumor growth was observed. Intriguingly, G1T48 alone (30 and 100 mg/kg) or in combination with lerociclib was more efficacious than fulvestrant (Fig. 7a). Survival curve analysis demonstrated that the combination of 30 mg/kg G1T48 with lerociclib was more effective than monotherapy using either drug alone (Fig. 7b, c). Taken together these data highlight that G1T48 is either comparable or superior to fulvestrant in several models of endocrine therapy resistance, demonstrating its potential as a therapeutic agent.

Discussion

Targeting ER activity using therapies that directly oppose the mitogenic action of estrogen or that block estrogen synthesis is a proven strategy for the treatment and prevention of breast cancer. In locally advanced or metastatic disease, resistance to these therapies frequently emerges within two years, at which time treatment options are severely limited. Fulvestrant, a potent ER antagonist and downregulator, was initially approved for the treatment of endocrine therapy-resistant disease and more recently as first-line therapy for advanced ER-positive, HER2-negative breast cancer not previously treated with endocrine therapy. However, despite promising preclinical activity, the poorly controlled pharmacokinetics of fulvestrant remains a significant barrier to prolonged clinical efficacy. Clinical trials comparing high-dose (500 mg) to low-dose (250 mg) fulvestrant demonstrated superiority for the 500 mg dose in both first- and second-line settings, suggesting that increased target engagement can improve the outcome of ER degradation therapy [54, 55]. However, given its intramuscular route of administration, continued improvements in the clinical response to fulvestrant by further dose escalation appear unlikely. Therefore, development campaigns in this area have focused on the identification of orally bioavailable SERDs. The most active SERDs share common chemical features: either (a) a steroidal backbone (e.g., fulvestrant, RU58,668) or (b) an acrylic acid side chain (GW7604, GDC-0810, AZD9496) [29, 31, 32, 40, 56, 57]. Additional ER antagonists with novel chemical structures have also been reported to exhibit SERD properties [35, 36, 58], but none has yet gained FDA approval and some have been discontinued due to adverse effects or for undisclosed reasons [2936, 40, 56]. We have identified a novel, orally bioavailable SERD, G1T48, that contains both a steroidal backbone and an acrylate side chain. G1T48 binds ER with low nanomolar affinity, inhibits estrogen-mediated target gene expression and breast cancer cell growth, and importantly blocks the tumor promoting effects of ER in both naïve and endocrine therapy-resistant animal models of breast cancer. Similar to AZD9496 and GDC-0810, G1T48 has good pharmacokinetic properties and maintains a more favorable side-effect profile compared to those reported for AZD9496 [56, 59, 60].
A hallmark feature of fulvestrant differentiating it from compounds like tamoxifen is that fulvestrant is a true antagonist with no agonist activity regardless of tissue context. By contrast, tamoxifen is a Selective Estrogen Receptor Modulator (SERM), demonstrating robust antagonist activity in the breast, but mimicking the agonist effect of estrogen in bone, the endometrium, and serum lipid profiles [61, 62]. This mechanistic difference between tamoxifen and fulvestrant can also be observed in breast cancer cells, where transcriptional profiling studies revealed that tamoxifen can regulate a subset of genes in a similar manner to estradiol. Our ER target gene regulation studies confirm the agonist activity of tamoxifen, with stimulation of SDK2, AGR2, and RAPGEL1 expression similar to the effect of estrogen treatment [42]. Compounds with SERD activity such as fulvestrant and AZD9496 did not increase these transcripts, consistent with a lack of agonist potential (i.e., pure antagonism). The transcriptional profile in breast cancer cells of G1T48 is most similar to fulvestrant and other SERDs. Interestingly, our studies revealed that there were modest differences in the transcriptional profiles even among the pure antagonist class of compounds, suggesting that they might engender different receptor conformations. The impact of these differences in ER target gene activation remain to be explored but could suggest that cross-resistance between different classes of SERDs can be avoided. Recent studies have indicated that in addition to receptor degradation, ER mobility is differentially impacted by sub-classes of SERMs and SERDs, and that compounds impeding mobility are more efficacious antagonists [35]. The impact of G1T48 on ER mobility is not currently known; however, our studies establish that G1T48 has very low intrinsic ER agonist activity.
Acquired resistance to endocrine therapy is complex and multifactorial; however, mutations in the ESR1 gene that result in ligand-independent receptor activity have emerged as a potential mechanism to account for approximately 30–40% of resistant disease following AI treatment [8, 4446, 4851]. It is significant, therefore, that G1T48 was found to suppress both the ligand-independent cell growth and transcriptional activity attributed to the two most prevalent endocrine refractory ER mutants, ER-Y537S and ER-D538G. Intriguingly, in contrast to the reconstituted transactivation assay in SKBR3, G1T48 was found to efficiently inhibit the growth of MCF7 cells engineered to overexpress the ER-Y537S variant. Cell context may contribute to this discrepancy; differential cofactor expression patterns in the two cell lines and/or the presence of endogenous wtER in MCF7 cells may influence G1T48 efficacy.
Long-term estrogen deprivation leading to a state of estrogen hypersensitivity is another means to model aromatase inhibitor therapy resistance. We have developed a new model of resistance to estrogen deprivation without ER mutation [53]. Using this model system, treatment with low-dose G1T48 (5 mg) resulted in incomplete tumor growth inhibition, while high-dose G1T48 (100 mg) as monotherapy resulted in tumor regression in the majority of animals, similar to fulvestrant, demonstrating the effectiveness of SERD therapy in this setting of resistant disease.
The combination of SERDs with CDK4/6 inhibitors has been evaluated clinically, most recently in the PALOMA-3 trial comparing the co-administration of the CDK4/6 inhibitor palbociclib (Ibrance®) with fulvestrant to fulvestrant alone. The results of this study demonstrated an overall survival benefit (median survival 34.9 months compared to 28 months) and a significant progression free survival rate (9.5 months vs 4.6 months) for the combination arm [17, 20]. These noteworthy improvements led to the 2016 FDA approval of palbociclib and fulvestrant combination therapy for ER-positive, HER-2- negative breast cancers progressing on other endocrine therapies. Further trials (MONALESSA-3 (NCT02422615) and MONARCH-2 (NCT02107703) have also demonstrated the utility of administering other CDK4/6 inhibitors with fulvestrant to improve patient outcomes [1720]. The increased efficacy observed for the combination of G1T48 and lerociclib, as compared to monotherapy administration, in multiple in vivo breast cancer models sensitive or refractory to endocrine therapy treatment supports the potential utility of this regimen as an intervention in multiple stages of breast cancer treatment. Furthermore, lerociclib has been shown to promote less myelosuppression than palbociclib [25, 26]. Collectively, these data indicate that G1T48 has the potential to be an efficacious oral antineoplastic agent in ER+ breast cancer.

Acknowledgements

This work was supported by a sponsored research grant from G1 Therapeutics. The authors would like to thank the McDonnell laboratory for providing the space to perform the experiments and for topical and intellectual discussions about the research. The authors would also like to think Alexander Yllanes for his technical assistance.

Compliance with ethical standards

Conflicts of interest

Jessica A. Sorrentino, Delita A. Thompson, John E. Bisi, and Jay C. Strum are employees of G1 Therapeutics, Inc. Delita A. Thompson, John E. Bisi, Jessica A. Sorrentino, and Jay C. Strum own stock in G1 Therapeutics, Inc. John D. Norris is a paid consultant for G1 Therapeutics, Inc. All other authors have no conflicts of interest.

Ethical approval

This article does not contain any studies with human participants performed by any of the authors. All procedures were approved by the Duke University Institutional Animal Care and Use Committee (IACUC) prior to initiating the experiment (protocol A011-16-01 and A272-18-12).
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Gynäkologie

Kombi-Abonnement

Mit e.Med Gynäkologie erhalten Sie Zugang zu CME-Fortbildungen der beiden Fachgebiete, den Premium-Inhalten der Fachzeitschriften, inklusive einer gedruckten gynäkologischen oder urologischen Zeitschrift Ihrer Wahl.

Anhänge

Electronic supplementary material

Below is the link to the electronic supplementary material.
Literatur
1.
Zurück zum Zitat Ignatiadis M, Sotiriou C (2013) Luminal breast cancer: from biology to treatment. Nat Rev Clin Oncol 10(9):494–506CrossRefPubMed Ignatiadis M, Sotiriou C (2013) Luminal breast cancer: from biology to treatment. Nat Rev Clin Oncol 10(9):494–506CrossRefPubMed
2.
Zurück zum Zitat Maximov PY, Lee TM, Jordan VC (2013) The discovery and development of Selective Estrogen Receptor Modulators (SERMs) for clinical practice. Curr Clin Pharmacol 8(2):135–155CrossRefPubMedPubMedCentral Maximov PY, Lee TM, Jordan VC (2013) The discovery and development of Selective Estrogen Receptor Modulators (SERMs) for clinical practice. Curr Clin Pharmacol 8(2):135–155CrossRefPubMedPubMedCentral
3.
Zurück zum Zitat Rydén L et al (2016) Aromatase inhibitors alone or sequentially combined with tamoxifen in postmenopausal early breast cancer compared with tamoxifen or placebo - Meta-analyses on efficacy and adverse events based on randomized clinical trials. Breast 26:106–114CrossRefPubMed Rydén L et al (2016) Aromatase inhibitors alone or sequentially combined with tamoxifen in postmenopausal early breast cancer compared with tamoxifen or placebo - Meta-analyses on efficacy and adverse events based on randomized clinical trials. Breast 26:106–114CrossRefPubMed
4.
Zurück zum Zitat Font de Mora J, Brown M (2000) AIB1 is a conduit for kinase-mediated growth factor signaling to the estrogen receptor. Mol Cell Biol 20(14):5041–5047CrossRefPubMedPubMedCentral Font de Mora J, Brown M (2000) AIB1 is a conduit for kinase-mediated growth factor signaling to the estrogen receptor. Mol Cell Biol 20(14):5041–5047CrossRefPubMedPubMedCentral
5.
Zurück zum Zitat Campbell RA et al (2001) Phosphatidylinositol 3-kinase/AKT-mediated activation of estrogen receptor alpha: a new model for anti-estrogen resistance. J Biol Chem 276(13):9817–9824CrossRefPubMed Campbell RA et al (2001) Phosphatidylinositol 3-kinase/AKT-mediated activation of estrogen receptor alpha: a new model for anti-estrogen resistance. J Biol Chem 276(13):9817–9824CrossRefPubMed
6.
Zurück zum Zitat Yue W et al (2002) Activation of the MAPK pathway enhances sensitivity of MCF-7 breast cancer cells to the mitogenic effect of estradiol. Endocrinology 143(9):3221–3229CrossRefPubMed Yue W et al (2002) Activation of the MAPK pathway enhances sensitivity of MCF-7 breast cancer cells to the mitogenic effect of estradiol. Endocrinology 143(9):3221–3229CrossRefPubMed
7.
Zurück zum Zitat Medunjanin S et al (2005) Glycogen synthase kinase-3 interacts with and phosphorylates estrogen receptor alpha and is involved in the regulation of receptor activity. J Biol Chem 280(38):33006–33014CrossRefPubMed Medunjanin S et al (2005) Glycogen synthase kinase-3 interacts with and phosphorylates estrogen receptor alpha and is involved in the regulation of receptor activity. J Biol Chem 280(38):33006–33014CrossRefPubMed
8.
9.
Zurück zum Zitat Ma CX et al (2015) Mechanisms of aromatase inhibitor resistance. Nat Rev Cancer 15(5):261–275CrossRefPubMed Ma CX et al (2015) Mechanisms of aromatase inhibitor resistance. Nat Rev Cancer 15(5):261–275CrossRefPubMed
10.
Zurück zum Zitat Tryfonidis K et al (2016) Endocrine treatment in breast cancer: cure, resistance and beyond. Cancer Treat Rev 50:68–81CrossRefPubMed Tryfonidis K et al (2016) Endocrine treatment in breast cancer: cure, resistance and beyond. Cancer Treat Rev 50:68–81CrossRefPubMed
11.
12.
Zurück zum Zitat McDonnell DP, Wardell SE, Norris JD (2015) Oral selective estrogen receptor downregulators (SERDs), a breakthrough endocrine therapy for breast cancer. J Med Chem 58(12):4883–4887CrossRefPubMedPubMedCentral McDonnell DP, Wardell SE, Norris JD (2015) Oral selective estrogen receptor downregulators (SERDs), a breakthrough endocrine therapy for breast cancer. J Med Chem 58(12):4883–4887CrossRefPubMedPubMedCentral
13.
Zurück zum Zitat Krop IE et al (2016) Pictilisib for oestrogen receptor-positive, aromatase inhibitor-resistant, advanced or metastatic breast cancer (FERGI): a randomised, double-blind, placebo-controlled, phase 2 trial. Lancet Oncol 17(6):811–821CrossRefPubMedPubMedCentral Krop IE et al (2016) Pictilisib for oestrogen receptor-positive, aromatase inhibitor-resistant, advanced or metastatic breast cancer (FERGI): a randomised, double-blind, placebo-controlled, phase 2 trial. Lancet Oncol 17(6):811–821CrossRefPubMedPubMedCentral
14.
Zurück zum Zitat Baselga J et al (2017) Buparlisib plus fulvestrant versus placebo plus fulvestrant in postmenopausal, hormone receptor-positive, HER2-negative, advanced breast cancer (BELLE-2): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol 18(7):904–916CrossRefPubMedPubMedCentral Baselga J et al (2017) Buparlisib plus fulvestrant versus placebo plus fulvestrant in postmenopausal, hormone receptor-positive, HER2-negative, advanced breast cancer (BELLE-2): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol 18(7):904–916CrossRefPubMedPubMedCentral
15.
Zurück zum Zitat Campone M et al (2018) Buparlisib plus fulvestrant versus placebo plus fulvestrant for postmenopausal, hormone receptor-positive, human epidermal growth factor receptor 2-negative, advanced breast cancer: overall survival results from BELLE-2. Eur J Cancer 103:147–154CrossRefPubMed Campone M et al (2018) Buparlisib plus fulvestrant versus placebo plus fulvestrant for postmenopausal, hormone receptor-positive, human epidermal growth factor receptor 2-negative, advanced breast cancer: overall survival results from BELLE-2. Eur J Cancer 103:147–154CrossRefPubMed
16.
Zurück zum Zitat Di Leo A et al (2018) Buparlisib plus fulvestrant in postmenopausal women with hormone-receptor-positive, HER2-negative, advanced breast cancer progressing on or after mTOR inhibition (BELLE-3): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol 19(1):87–100CrossRefPubMed Di Leo A et al (2018) Buparlisib plus fulvestrant in postmenopausal women with hormone-receptor-positive, HER2-negative, advanced breast cancer progressing on or after mTOR inhibition (BELLE-3): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol 19(1):87–100CrossRefPubMed
17.
Zurück zum Zitat Cristofanilli M et al (2016) Fulvestrant plus palbociclib versus fulvestrant plus placebo for treatment of hormone-receptor-positive, HER2-negative metastatic breast cancer that progressed on previous endocrine therapy (PALOMA-3): final analysis of the multicentre, double-blind, phase 3 randomised controlled trial. Lancet Oncol 17(4):425–439CrossRefPubMed Cristofanilli M et al (2016) Fulvestrant plus palbociclib versus fulvestrant plus placebo for treatment of hormone-receptor-positive, HER2-negative metastatic breast cancer that progressed on previous endocrine therapy (PALOMA-3): final analysis of the multicentre, double-blind, phase 3 randomised controlled trial. Lancet Oncol 17(4):425–439CrossRefPubMed
18.
Zurück zum Zitat Sledge GW et al (2017) MONARCH 2: abemaciclib in combination with fulvestrant in women With HR+/HER2- advanced breast cancer who had progressed while receiving endocrine therapy. J Clin Oncol 35(25):2875–2884CrossRefPubMed Sledge GW et al (2017) MONARCH 2: abemaciclib in combination with fulvestrant in women With HR+/HER2- advanced breast cancer who had progressed while receiving endocrine therapy. J Clin Oncol 35(25):2875–2884CrossRefPubMed
19.
Zurück zum Zitat Slamon DJ et al (2018) Phase III Randomized study of ribociclib and fulvestrant in hormone receptor-positive, human epidermal growth factor receptor 2-negative advanced breast cancer: MONALEESA-3. J Clin Oncol 36(24):2465–2472CrossRefPubMed Slamon DJ et al (2018) Phase III Randomized study of ribociclib and fulvestrant in hormone receptor-positive, human epidermal growth factor receptor 2-negative advanced breast cancer: MONALEESA-3. J Clin Oncol 36(24):2465–2472CrossRefPubMed
20.
Zurück zum Zitat Turner NC et al (2018) Overall survival with palbociclib and fulvestrant in advanced breast cancer. N Engl J Med 379(20):1926–1936CrossRefPubMed Turner NC et al (2018) Overall survival with palbociclib and fulvestrant in advanced breast cancer. N Engl J Med 379(20):1926–1936CrossRefPubMed
21.
Zurück zum Zitat Howell A, Sapunar F (2011) Fulvestrant revisited: efficacy and safety of the 500-mg dose. Clin Breast Cancer 11(4):204–210CrossRefPubMed Howell A, Sapunar F (2011) Fulvestrant revisited: efficacy and safety of the 500-mg dose. Clin Breast Cancer 11(4):204–210CrossRefPubMed
22.
Zurück zum Zitat Robertson JF et al (2004) Pharmacokinetic profile of intramuscular fulvestrant in advanced breast cancer. Clin Pharmacokinet 43(8):529–538CrossRefPubMed Robertson JF et al (2004) Pharmacokinetic profile of intramuscular fulvestrant in advanced breast cancer. Clin Pharmacokinet 43(8):529–538CrossRefPubMed
23.
24.
Zurück zum Zitat Xiong R et al (2017) Novel selective estrogen receptor downregulators (SERDs) developed against treatment-resistant breast cancer. J Med Chem 60(4):1325–1342CrossRefPubMedPubMedCentral Xiong R et al (2017) Novel selective estrogen receptor downregulators (SERDs) developed against treatment-resistant breast cancer. J Med Chem 60(4):1325–1342CrossRefPubMedPubMedCentral
25.
Zurück zum Zitat Bisi JE et al (2017) Preclinical development of G1T38: A novel, potent and selective inhibitor of cyclin dependent kinases 4/6 for use as an oral antineoplastic in patients with CDK4/6 sensitive tumors. Oncotarget 8(26):42343–42358CrossRefPubMedPubMedCentral Bisi JE et al (2017) Preclinical development of G1T38: A novel, potent and selective inhibitor of cyclin dependent kinases 4/6 for use as an oral antineoplastic in patients with CDK4/6 sensitive tumors. Oncotarget 8(26):42343–42358CrossRefPubMedPubMedCentral
27.
Zurück zum Zitat Wright TM et al (2014) Delineation of a FOXA1/ERalpha/AGR2 regulatory loop that is dysregulated in endocrine therapy-resistant breast cancer. Mol Cancer Res 12(12):1829–1839CrossRefPubMedPubMedCentral Wright TM et al (2014) Delineation of a FOXA1/ERalpha/AGR2 regulatory loop that is dysregulated in endocrine therapy-resistant breast cancer. Mol Cancer Res 12(12):1829–1839CrossRefPubMedPubMedCentral
28.
Zurück zum Zitat Livak K, Schmittgen T (2001) Analysis of relative gene expression data using real-time quantitative PCR and the 2(ΔΔC(T)) method. Methods 25:402–408CrossRefPubMed Livak K, Schmittgen T (2001) Analysis of relative gene expression data using real-time quantitative PCR and the 2(ΔΔC(T)) method. Methods 25:402–408CrossRefPubMed
29.
Zurück zum Zitat Lai A et al (2015) Identification of GDC-0810 (ARN-810), an orally bioavailable selective estrogen receptor degrader (SERD) that Demonstrates robust activity in tamoxifen-resistant breast cancer xenografts. J Med Chem 58(12):4888–4904CrossRefPubMed Lai A et al (2015) Identification of GDC-0810 (ARN-810), an orally bioavailable selective estrogen receptor degrader (SERD) that Demonstrates robust activity in tamoxifen-resistant breast cancer xenografts. J Med Chem 58(12):4888–4904CrossRefPubMed
30.
Zurück zum Zitat Wardell SE et al (2015) Evaluation of the pharmacological activities of RAD1901, a selective estrogen receptor degrader. Endocr Relat Cancer 22(5):713–724CrossRefPubMedPubMedCentral Wardell SE et al (2015) Evaluation of the pharmacological activities of RAD1901, a selective estrogen receptor degrader. Endocr Relat Cancer 22(5):713–724CrossRefPubMedPubMedCentral
31.
Zurück zum Zitat Joseph JD et al (2016) The selective estrogen receptor downregulator GDC-0810 is efficacious in diverse models of ER+ breast cancer. Elife 5:e15828CrossRefPubMedPubMedCentral Joseph JD et al (2016) The selective estrogen receptor downregulator GDC-0810 is efficacious in diverse models of ER+ breast cancer. Elife 5:e15828CrossRefPubMedPubMedCentral
32.
Zurück zum Zitat Weir HM et al (2016) AZD9496: An oral estrogen receptor inhibitor that blocks the growth of ER-positive and ESR1-mutant breast tumors in preclinical models. Cancer Res 76(11):3307–3318CrossRefPubMed Weir HM et al (2016) AZD9496: An oral estrogen receptor inhibitor that blocks the growth of ER-positive and ESR1-mutant breast tumors in preclinical models. Cancer Res 76(11):3307–3318CrossRefPubMed
33.
Zurück zum Zitat Bihani T et al (2017) Elacestrant (RAD1901), a selective estrogen receptor degrader (SERD), has antitumor activity in multiple ER. Clin Cancer Res 23(16):4793–4804CrossRefPubMed Bihani T et al (2017) Elacestrant (RAD1901), a selective estrogen receptor degrader (SERD), has antitumor activity in multiple ER. Clin Cancer Res 23(16):4793–4804CrossRefPubMed
34.
Zurück zum Zitat Tria GS et al (2018) Discovery of LSZ102, a potent, orally bioavailable selective estrogen receptor degrader (SERD) for the treatment of estrogen receptor positive breast cancer. J Med Chem 61(7):2837–2864CrossRefPubMed Tria GS et al (2018) Discovery of LSZ102, a potent, orally bioavailable selective estrogen receptor degrader (SERD) for the treatment of estrogen receptor positive breast cancer. J Med Chem 61(7):2837–2864CrossRefPubMed
35.
Zurück zum Zitat Guan J et al (2019) Therapeutic ligands antagonize estrogen receptor function by impairing its mobility. Cell 178(4):949–963.e18CrossRefPubMed Guan J et al (2019) Therapeutic ligands antagonize estrogen receptor function by impairing its mobility. Cell 178(4):949–963.e18CrossRefPubMed
36.
Zurück zum Zitat Kahraman M et al (2019) Maximizing ER-α degradation maximizes activity in a tamoxifen-resistant breast cancer model: identification of GDC-0927. ACS Med Chem Lett 10(1):50–55CrossRefPubMed Kahraman M et al (2019) Maximizing ER-α degradation maximizes activity in a tamoxifen-resistant breast cancer model: identification of GDC-0927. ACS Med Chem Lett 10(1):50–55CrossRefPubMed
37.
Zurück zum Zitat Draper MW et al (1996) A controlled trial of raloxifene (LY139481) HCl: impact on bone turnover and serum lipid profile in healthy postmenopausal women. J Bone Miner Res 11(6):835–842CrossRefPubMed Draper MW et al (1996) A controlled trial of raloxifene (LY139481) HCl: impact on bone turnover and serum lipid profile in healthy postmenopausal women. J Bone Miner Res 11(6):835–842CrossRefPubMed
38.
Zurück zum Zitat Vogel VG et al (2006) Effects of tamoxifen vs raloxifene on the risk of developing invasive breast cancer and other disease outcomes: the NSABP Study of Tamoxifen and Raloxifene (STAR) P-2 trial. JAMA 295(23):2727–2741CrossRefPubMed Vogel VG et al (2006) Effects of tamoxifen vs raloxifene on the risk of developing invasive breast cancer and other disease outcomes: the NSABP Study of Tamoxifen and Raloxifene (STAR) P-2 trial. JAMA 295(23):2727–2741CrossRefPubMed
39.
Zurück zum Zitat Wijayaratne AL et al (1999) Comparative analyses of mechanistic differences among antiestrogens. Endocrinology 140(12):5828–5840CrossRefPubMed Wijayaratne AL et al (1999) Comparative analyses of mechanistic differences among antiestrogens. Endocrinology 140(12):5828–5840CrossRefPubMed
40.
Zurück zum Zitat De Savi C et al (2015) Optimization of a novel binding motif to (E)-3-(3,5-difluoro-4-((1R,3R)-2-(2-fluoro-2-methylpropyl)-3-methyl-2,3,4,9-tetrahydro-1H-pyrido[3,4-b]indol-1-yl)phenyl)acrylic acid (AZD9496), a potent and orally bioavailable selective estrogen receptor downregulator and antagonist. J Med Chem 58(20):8128–8140CrossRefPubMed De Savi C et al (2015) Optimization of a novel binding motif to (E)-3-(3,5-difluoro-4-((1R,3R)-2-(2-fluoro-2-methylpropyl)-3-methyl-2,3,4,9-tetrahydro-1H-pyrido[3,4-b]indol-1-yl)phenyl)acrylic acid (AZD9496), a potent and orally bioavailable selective estrogen receptor downregulator and antagonist. J Med Chem 58(20):8128–8140CrossRefPubMed
41.
Zurück zum Zitat Wardell SE, Nelson ER, McDonnell DP (2014) From empirical to mechanism-based discovery of clinically useful Selective Estrogen Receptor Modulators (SERMs). Steroids 90:30–38CrossRefPubMedPubMedCentral Wardell SE, Nelson ER, McDonnell DP (2014) From empirical to mechanism-based discovery of clinically useful Selective Estrogen Receptor Modulators (SERMs). Steroids 90:30–38CrossRefPubMedPubMedCentral
42.
Zurück zum Zitat Wardell SE, Kazmin D, McDonnell DP (2012) Research resource: transcriptional profiling in a cellular model of breast cancer reveals functional and mechanistic differences between clinically relevant SERM and between SERM/estrogen complexes. Mol Endocrinol 26(7):1235–1248CrossRefPubMedPubMedCentral Wardell SE, Kazmin D, McDonnell DP (2012) Research resource: transcriptional profiling in a cellular model of breast cancer reveals functional and mechanistic differences between clinically relevant SERM and between SERM/estrogen complexes. Mol Endocrinol 26(7):1235–1248CrossRefPubMedPubMedCentral
43.
Zurück zum Zitat Nagy L, Schwabe JW (2004) Mechanism of the nuclear receptor molecular switch. Trends Biochem Sci 29(6):317–324CrossRefPubMed Nagy L, Schwabe JW (2004) Mechanism of the nuclear receptor molecular switch. Trends Biochem Sci 29(6):317–324CrossRefPubMed
44.
Zurück zum Zitat Bahreini A et al (2017) Mutation site and context dependent effects of ESR1 mutation in genome-edited breast cancer cell models. Breast Cancer Res 19(1):60CrossRefPubMedPubMedCentral Bahreini A et al (2017) Mutation site and context dependent effects of ESR1 mutation in genome-edited breast cancer cell models. Breast Cancer Res 19(1):60CrossRefPubMedPubMedCentral
45.
Zurück zum Zitat Harrod A et al (2017) Genomic modelling of the ESR1 Y537S mutation for evaluating function and new therapeutic approaches for metastatic breast cancer. Oncogene 36(16):2286–2296CrossRefPubMed Harrod A et al (2017) Genomic modelling of the ESR1 Y537S mutation for evaluating function and new therapeutic approaches for metastatic breast cancer. Oncogene 36(16):2286–2296CrossRefPubMed
46.
Zurück zum Zitat Martin LA et al (2017) Discovery of naturally occurring ESR1 mutations in breast cancer cell lines modelling endocrine resistance. Nat Commun 8(1):1865CrossRefPubMedPubMedCentral Martin LA et al (2017) Discovery of naturally occurring ESR1 mutations in breast cancer cell lines modelling endocrine resistance. Nat Commun 8(1):1865CrossRefPubMedPubMedCentral
47.
Zurück zum Zitat Merenbakh-Lamin K et al (2013) D538G mutation in estrogen receptor-α: A novel mechanism for acquired endocrine resistance in breast cancer. Cancer Res 73(23):6856–6864CrossRefPubMed Merenbakh-Lamin K et al (2013) D538G mutation in estrogen receptor-α: A novel mechanism for acquired endocrine resistance in breast cancer. Cancer Res 73(23):6856–6864CrossRefPubMed
50.
Zurück zum Zitat Toy W et al (2017) Activating ESR1 Mutations Differentially Impact the Efficacy of ER Antagonists. Cancer Discov 7(3):277–287CrossRefPubMed Toy W et al (2017) Activating ESR1 Mutations Differentially Impact the Efficacy of ER Antagonists. Cancer Discov 7(3):277–287CrossRefPubMed
51.
Zurück zum Zitat Jeselsohn R et al (2014) Emergence of constitutively active estrogen receptor-α mutations in pretreated advanced estrogen receptor-positive breast cancer. Clin Cancer Res 20(7):1757–1767CrossRefPubMedPubMedCentral Jeselsohn R et al (2014) Emergence of constitutively active estrogen receptor-α mutations in pretreated advanced estrogen receptor-positive breast cancer. Clin Cancer Res 20(7):1757–1767CrossRefPubMedPubMedCentral
52.
Zurück zum Zitat Wardell SE et al (2015) Efficacy of SERD/SERM hybrid-CDK4/6 inhibitor combinations in models of endocrine therapy-resistant breast cancer. Clin Cancer Res 21(22):5121–5130CrossRefPubMedPubMedCentral Wardell SE et al (2015) Efficacy of SERD/SERM hybrid-CDK4/6 inhibitor combinations in models of endocrine therapy-resistant breast cancer. Clin Cancer Res 21(22):5121–5130CrossRefPubMedPubMedCentral
53.
Zurück zum Zitat Wardell SE et al (2019) Pharmacokinetic and pharmacodynamic analysis of fulvestrant in preclinical models of breast cancer to assess the importance of its estrogen receptor-alpha degrader activity in antitumor efficacy. Breast Cancer Res Treat 179(1):67–77CrossRefPubMedPubMedCentral Wardell SE et al (2019) Pharmacokinetic and pharmacodynamic analysis of fulvestrant in preclinical models of breast cancer to assess the importance of its estrogen receptor-alpha degrader activity in antitumor efficacy. Breast Cancer Res Treat 179(1):67–77CrossRefPubMedPubMedCentral
54.
Zurück zum Zitat Di Leo A et al (2010) Results of the CONFIRM phase III trial comparing fulvestrant 250 mg with fulvestrant 500 mg in postmenopausal women with estrogen receptor-positive advanced breast cancer. J Clin Oncol 28(30):4594–4600CrossRefPubMed Di Leo A et al (2010) Results of the CONFIRM phase III trial comparing fulvestrant 250 mg with fulvestrant 500 mg in postmenopausal women with estrogen receptor-positive advanced breast cancer. J Clin Oncol 28(30):4594–4600CrossRefPubMed
55.
Zurück zum Zitat Di Leo A et al. (2014) Final overall survival: fulvestrant 500 mg vs 250 mg in the randomized CONFIRM trial. J Natl Cancer Inst 106(1):djt337. Di Leo A et al. (2014) Final overall survival: fulvestrant 500 mg vs 250 mg in the randomized CONFIRM trial. J Natl Cancer Inst 106(1):djt337.
56.
Zurück zum Zitat Hamilton EP et al (2018) A first-in-human study of the new oral selective estrogen receptor degrader AZD9496 for ER. Clin Cancer Res 24(15):3510–3518CrossRefPubMed Hamilton EP et al (2018) A first-in-human study of the new oral selective estrogen receptor degrader AZD9496 for ER. Clin Cancer Res 24(15):3510–3518CrossRefPubMed
57.
Zurück zum Zitat Willson TM et al (1997) Dissection of the molecular mechanism of action of GW5638, a novel estrogen receptor ligand, provides insights into the role of estrogen receptor in bone. Endocrinology 138(9):3901–3911CrossRefPubMed Willson TM et al (1997) Dissection of the molecular mechanism of action of GW5638, a novel estrogen receptor ligand, provides insights into the role of estrogen receptor in bone. Endocrinology 138(9):3901–3911CrossRefPubMed
58.
Zurück zum Zitat Wardell SE et al (2013) Bazedoxifene exhibits antiestrogenic activity in animal models of tamoxifen-resistant breast cancer: implications for treatment of advanced disease. Clin Cancer Res 19(9):2420–2431CrossRefPubMedPubMedCentral Wardell SE et al (2013) Bazedoxifene exhibits antiestrogenic activity in animal models of tamoxifen-resistant breast cancer: implications for treatment of advanced disease. Clin Cancer Res 19(9):2420–2431CrossRefPubMedPubMedCentral
59.
Zurück zum Zitat Dees EC et al (2019) First-in-human dose-escalation study of g1t48, an oral selective estrogen receptor degrader (SERD), in postmenopausal women with ER+/Her2− locally advanced or metastatic breast cancer (ABC). European Society for Medical Oncology (ESMO), Barcelona, Spain Dees EC et al (2019) First-in-human dose-escalation study of g1t48, an oral selective estrogen receptor degrader (SERD), in postmenopausal women with ER+/Her2− locally advanced or metastatic breast cancer (ABC). European Society for Medical Oncology (ESMO), Barcelona, Spain
60.
Zurück zum Zitat Cheung KWK et al (2019) GDC-0810 Pharmacokinetics and transporter-mediated drug interaction evaluation with an endogenous biomarker in the first-in-human, dose escalation study. Drug Metab Dispos 47(9):966–973CrossRefPubMed Cheung KWK et al (2019) GDC-0810 Pharmacokinetics and transporter-mediated drug interaction evaluation with an endogenous biomarker in the first-in-human, dose escalation study. Drug Metab Dispos 47(9):966–973CrossRefPubMed
61.
Zurück zum Zitat Love RR et al (1992) Effects of tamoxifen on bone mineral density in postmenopausal women with breast cancer. N Engl J Med 326(13):852–856CrossRefPubMed Love RR et al (1992) Effects of tamoxifen on bone mineral density in postmenopausal women with breast cancer. N Engl J Med 326(13):852–856CrossRefPubMed
62.
Zurück zum Zitat Love RR et al (1991) Effects of tamoxifen on cardiovascular risk factors in postmenopausal women. Ann Intern Med 115(11):860–864CrossRefPubMed Love RR et al (1991) Effects of tamoxifen on cardiovascular risk factors in postmenopausal women. Ann Intern Med 115(11):860–864CrossRefPubMed
Metadaten
Titel
G1T48, an oral selective estrogen receptor degrader, and the CDK4/6 inhibitor lerociclib inhibit tumor growth in animal models of endocrine-resistant breast cancer
verfasst von
Kaitlyn J. Andreano
Suzanne E. Wardell
Jennifer G. Baker
Taylor K. Desautels
Robert Baldi
Christina A. Chao
Kendall A. Heetderks
Yeeun Bae
Rui Xiong
Debra A. Tonetti
Lauren M. Gutgesell
Jiong Zhao
Jessica A. Sorrentino
Delita A. Thompson
John E. Bisi
Jay C. Strum
Gregory R. J. Thatcher
John D. Norris
Publikationsdatum
04.03.2020
Verlag
Springer US
Erschienen in
Breast Cancer Research and Treatment / Ausgabe 3/2020
Print ISSN: 0167-6806
Elektronische ISSN: 1573-7217
DOI
https://doi.org/10.1007/s10549-020-05575-9

Weitere Artikel der Ausgabe 3/2020

Breast Cancer Research and Treatment 3/2020 Zur Ausgabe

Adjuvante Immuntherapie verlängert Leben bei RCC

25.04.2024 Nierenkarzinom Nachrichten

Nun gibt es auch Resultate zum Gesamtüberleben: Eine adjuvante Pembrolizumab-Therapie konnte in einer Phase-3-Studie das Leben von Menschen mit Nierenzellkarzinom deutlich verlängern. Die Sterberate war im Vergleich zu Placebo um 38% geringer.

Alectinib verbessert krankheitsfreies Überleben bei ALK-positivem NSCLC

25.04.2024 NSCLC Nachrichten

Das Risiko für Rezidiv oder Tod von Patienten und Patientinnen mit reseziertem ALK-positivem NSCLC ist unter einer adjuvanten Therapie mit dem Tyrosinkinase-Inhibitor Alectinib signifikant geringer als unter platinbasierter Chemotherapie.

Bei Senioren mit Prostatakarzinom auf Anämie achten!

24.04.2024 DGIM 2024 Nachrichten

Patienten, die zur Behandlung ihres Prostatakarzinoms eine Androgendeprivationstherapie erhalten, entwickeln nicht selten eine Anämie. Wer ältere Patienten internistisch mitbetreut, sollte auf diese Nebenwirkung achten.

ICI-Therapie in der Schwangerschaft wird gut toleriert

Müssen sich Schwangere einer Krebstherapie unterziehen, rufen Immuncheckpointinhibitoren offenbar nicht mehr unerwünschte Wirkungen hervor als andere Mittel gegen Krebs.

Update Onkologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.