Next Article in Journal
Dyslipidemia and Inflammation as Hallmarks of Oxidative Stress in COVID-19: A Follow-Up Study
Next Article in Special Issue
The Role of Irisin/FNDC5 Expression and Its Serum Level in Breast Cancer
Previous Article in Journal
Deciphering the Binding of the Nuclear Localization Sequence of Myc Protein to the Nuclear Carrier Importin α3
Previous Article in Special Issue
Checkpoint Kinase 2 Inhibition Can Reverse Tamoxifen Resistance in ER-Positive Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modulating the Activity of Androgen Receptor for Treating Breast Cancer

Department of Pathology, Rm 014, 7/F, Block T, Queen Mary Hospital, Pokfulam Road, School of Clinical Medicine, Li Ka Shing Faculty of Medicine, The University of Hong Kong, Hong Kong SAR, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(23), 15342; https://doi.org/10.3390/ijms232315342
Submission received: 11 November 2022 / Revised: 1 December 2022 / Accepted: 2 December 2022 / Published: 5 December 2022
(This article belongs to the Special Issue Molecular Research and Treatment of Breast Cancer 2.0)

Abstract

:
The androgen receptor (AR) is a steroid hormone receptor widely detected in breast cancer. Evidence suggests that the AR might be a tumor suppressor in estrogen receptor alpha-positive (ERα+ve) breast cancer but a tumor promoter in estrogen receptor alpha-negative (ERα-ve) breast cancer. Modulating AR activity could be a potential strategy for treating breast cancer. For ERα+ve breast cancer, activation of the AR had been demonstrated to suppress the disease. In contrast, for ERα-ve breast cancer, blocking the AR could confer better prognosis to patients. These studies support the feasibility of utilizing AR modulators as anti-cancer drugs for different subtypes of breast cancer patients. Nevertheless, several issues still need to be addressed, such as the lack of standardization in the determination of AR positivity and the presence of AR splice variants. In future, the inclusion of the AR status in the breast cancer report at the time of diagnosis might help improve disease classification and treatment decision, thereby providing additional treatment strategies for breast cancer.

1. Introduction

The androgen receptor (AR), also known as NR3C4, is one of the sex steroid hormone receptors widely expressed in males and females. The receptor plays a vital role in multiple physiological processes, such as the development of the reproductive system [1]. The AR gene is located at locus Xq11–12 on the X-chromosome. It harbors eight exons encoding a protein of 110 kDa molecular weight. The AR protein is composed of three main functional domains: the N-terminal domain (NTD) (encoded by exon 1), the DNA binding domain (DBD) (encoded by exon 2–3) and the C-terminal ligand binding domain (LBD) (encoded by exon 5–8), together with a hinge region (HR) (encoded by exon 4) [2]. The inactivated AR is predominantly localized to the cytoplasm where it binds to molecular chaperones and co-chaperones which stabilize its protein structure and maintain it in an inactivated state [3]. The AR binds to ligands, such as dihydrotestosterone (DHT), via its LBD. This binding induces a conformational change and dissociation with its chaperones, leading to AR activation. As a transcriptional factor, the active AR is rapidly translocated into the nucleus and utilizes its DBD to recognize and bind to androgen-receptor-responsive elements (AREs) in the promoter or enhancer of its target genes. Such an interaction can modulate the transcriptional activity of the target genes [4]. A cytoplasmic AR can also interact with and trigger several secondary message pathways, leading to non-genomic activity. The AR targets a diverse range of oncogenic pathways, including PI3K/AKT, EGFR, Src and WNT pathways [5,6,7,8]. These pathways cover various cellular processes such as cell proliferation, migration, metastasis, apoptosis, differentiation and DNA damage repair. Steroid hormone receptors, such as estrogen receptor alpha (ERα), are considered a critical factor in the development of breast cancer [9]. Likewise, dysregulation of the AR also has been correlated to carcinogenesis. It is noted that around 80–90% of prostate cancers are detected as AR-positive (AR+ve) at the time of diagnosis [10], and AR signaling is crucial for prostate cancer development and progression [11]. Androgen deprivation therapy by drugs such as AR antagonists or by surgical castration to block the activity of the AR has become a gold standard in controlling disease progression. Furthermore, the AR may contribute to the growth of other malignancies, such as breast cancer, hepatocellular carcinoma and ovarian cancer, with ongoing clinical trials targeting the AR in patients bearing these cancers [12]. Therein, this narrative review will discuss the potential for targeting the AR in breast cancer treatment. The suitable electronic sources used in the current review were identified from the PubMed database by searching the following key words: breast cancer; androgen receptor; targeted therapy; prognosis.

2. Different Roles of AR in Breast Cancer

According to GLOBOCAN 2020, the latest global cancer statistics reported by the International Agency for Research on Cancer (IARC) of the World Health Organization (WHO), female breast cancer was estimated to be the most prevalent cancer worldwide, with approximately 2.26 million newly diagnosed cases (accounting for 11.7% of all cancers) and 0.68 million death cases in 2020 [13]. Breast cancer is a highly heterogeneous tumor that can be further categorized into five subtypes depending on their intrinsic gene expression profiles: Luminal A, Luminal B, human epidermal growth factor receptor 2-enriched (HER2+ve), basal-like and normal-like [14], of which Luminal A, Luminal B and normal-like are ERα-positive (ERα+ve), HER2+ve are ERα negative (ERα-ve) and basal-like are often ERα, progesterone receptor (PR) and HER2 negative (i.e., triple negative breast cancer; hereafter referred as TNBC). Remarkably, it has been reported that around 70–90% of breast cancers are detected as AR+ve cases [15,16], implicating that the AR may play some roles in disease development and progression. However, the role of the AR in breast cancer is controversial.
In ERα+ve breast cancer, ERα-mediated signaling is the driving factor for cancer development. The AR, by interfering with the function of ERα, may have a tumor suppressor function. The importance of cross-talk between AR and ERα has been supported by mechanistic studies which suggest that AR and ERα can interact with each other through the AR’s N-terminal domain (NTD) and ERα’s ligand-binding domain (LBD). This protein–protein interaction can inhibit the activity of both proteins [17]. Studies have also reported that the AR can compete with ERα by binding to Erα-targeted DNA sequences (Erα-responsive elements, ERE) or by interacting with its co-activators to hijack ERα from ERE to ARE (Figure 1) [18,19]. The AR can also up-regulate tumor suppressor genes, such as the tumor suppressors SEC14L2, EAF2 and ZBTB16, to repress tumor growth directly [19]. Furthermore, the AR can up-regulate the expression of the tumor suppressor protein, estrogen receptors beta (ERβ), to indirectly inhibit the biological activity of ERα and down-regulate the expression level of the proto-oncogene CCND1, or may enhance other tumor suppressors, including FOXO1 and FOXO3a [20,21]. Due to these activities against ERα, patients with an ERα+ve/AR+ve phenotype would have relatively better clinical features, e.g., smaller tumor [22,23], and longer survival time indicated by disease-free survival (DFS), overall survival (OS) and recurrence-free survival (RFS) [24,25,26]. Thus, the AR should be a good prognostic factor in ERα+ve breast cancer.
In ERα-ve breast cancer patients, including HER2+ve and TNBC, the AR shows oncogenic effects. In HER2+ve breast cancer, it has been reported that an active AR could induce the expression of HER2, which subsequently activates the MAPK pathway. The activated MAPK pathway induces the AR’s expression, resulting in a positive-feedback loop [27]. Moreover, an activated AR could up-regulate WNT7B. Activation of the WNT/β-catenin pathway would result. In addition, the AR could further form a complex with β-catenin. The complex could promote the transcription of HER3, which is a critical factor for the oncogenic activity of HER2 [28]. In TNBC, the AR has been shown to interact with SRC directly and activate the oncogenic pathway SRC/PI3K/FAK and its downstream genes [29]. These actions contribute to the continuous development of the tumor, leading to poor survival of the patients, rendering the AR a poor prognostic factor in ERα-ve breast cancer [25,30,31,32,33].

3. Modulating the Activity of AR as Breast Cancer Treatment

3.1. Activation of AR Can Suppress ERα+ve Breast Cancer

The possibility of modulating AR activity for breast cancer treatment can be traced back to 1939, when the first record of breast cancer patients who might benefit from receiving testosterone propionate to activate the AR was published [34]. Subsequently, similar observations were reported to support the potential application of androgens in treating breast cancer [35,36,37,38]. Although the expression status of ERα in the tumors was not indicated in these studies, they indicated the therapeutic value of utilizing androgens in breast cancer patients. Nevertheless, this innovative therapeutic strategy for breast cancer has not been widely studied for decades, since people found that androgens could be converted to estrogen via steroid aromatase. Such an effect may enhance the activity of the oncoprotein ERα. As more and more evidence revealed the anti-ERα properties of AR in breast cancers, investigators started to revisit the potential of androgens in treating ERα breast patients.
In a retrospective study, 508 postmenopausal and ovariectomized women who had received testosterone implants in addition to conventional hormone therapy were followed up for a mean duration of 5.8 years. The incidence of breast cancer was recorded, and the results showed that using testosterone might reduce the occurrence of conventional-hormone-therapy-induced breast cancer [39]. A prospective study recruiting 1268 pre-/postmenopausal women who were subcutaneously implanted with testosterone alone or combined with an aromatase inhibitor anastrozole to treat symptoms of hormone deficiency aimed to study the influence of the treatments on the occurrence of breast cancer: the 5-year interim report indicated that the annual incidence rate in the intent to treat patients was estimated to be 142 cases per 100,000 individuals (0.142%) and it could be further reduced to 73 cases per 100,000 people (0.073%) if the subjects received testosterone therapy, which was significantly lower than the control groups (0.293–0.39%) [40]; the 10-year analysis results showed the incidence rate of breast cancer in the testosterone-treated population was also significantly less than the age-specific Surveillance, Epidemiology, and End Results (SEER) expected result (0.165% vs. 0.271%). This study supported the role of testosterone in lowering the risk of hormone-therapy-related breast cancer [41].
The combination of testosterone and anastrozole was also applied to treat menopausal symptoms in 72 breast cancer survivors. The treatment was effective in relieving the symptoms without elevating estradiol; no tumor relapse was found for up to 8 years [42]. Combining testosterone and anastrozole for treating a hormone-receptor-positive breast cancer patient presented a promising result with a 12-fold decrease in the tumor volume and without elevation of estradiol [43]. Similarly, the treatment of testosterone and another aromatase inhibitor, letrozole, in hormone-receptor-positive invasive breast cancer patients led to 43% diminished tumor size with complete pathologic response when given in combination with chemotherapy [44].
Another study provided more direct evidence to support that androgens as a single drug could be an effective anti-cancer agent for treating ERα+ve breast cancer patients. The study recruited 53 ERα+ve/PR+ve metastatic breast cancer patients whose tumors no longer responded to anti-hormonal therapies. These patients were treated with testosterone propionate. Approximately 60% of the subjects showed disease regression or stabilization, representing a positive result [45]. An independent group performed a similar study. Hormone-receptor-positive metastatic breast cancer patients were subjected to receiving another androgen, fluoxymesterone. The results showed that around 43% of the participants achieved clinical benefit (3% complete response, 10% partial response and 30% stable disease) for at least 6 months [46].
As proof of principle, a phase II trial (NCT01616758) on evaluating the efficacy and safety of enobosarm, a synthetic and selective AR activating agent, in ERα+ve metastatic breast cancer demonstrated that around 50% (8/16) of the patients reached the best response with a median duration of 4.5 months. The drug was well tolerated [47]. Due to the promising result, a larger-scale trial (NCT02463032) was subsequently conducted. In the enobosarm-treated ERα+ve metastatic breast cancer patients whose nuclei AR staining is more than 40%, the clinical benefit rate (CBR), the best objective tumor response (BOR) and the median radiographic progression-free survival (rPFS) were 80%, 48% and 5.47 months (mean = 7.15 months), respectively. In comparison, in the patients whose nuclei AR staining is less than 40%, the CBR, BOR and rPFS were 18%, 0% and 2.72 months (mean = 2.7 months), respectively, suggesting the anti-cancer potential of enobosarm in ERα+ve breast cancer [48]. To further confirm these results in a larger cohort, a phase III study (NCT04869943) to evaluate the efficacy of enobosarm in AR+ve/ERα+ve metastatic breast cancer patients has been recently launched and is ongoing.
More recently, a pre-clinical study also demonstrated that activation of the AR by either DHT or enobosarm in ERα+ve endocrine-resistant breast cancer patient-derived xenografts (PDXs) significantly suppressed the estrogen-induced tumor growth. By contrast, inhibition of the AR activity via the AR antagonist enzalutamide enhanced the ability of estrogen to stimulate tumor growth [19]. These studies reflected the safety and efficacy of androgens in decreasing the risk of hormone-induced breast cancer, suppressing the tumor progression and inhibiting tumor growth, indicating the feasibility of utilizing androgens for treating ERα+ve breast cancer patients.

3.2. Blocking of AR Can Suppress ERα-ve Breast Cancer

3.2.1. Evidence from Clinical Studies

Many AR suppressive agents are available for treating prostate cancer. These agents significantly contributed to improving patients’ treatment outcomes. Due to the importance of the AR in ERα-ve breast cancer, studies on evaluating the efficacy of anti-androgens, alone or combined with other anti-breast cancer drugs, have been documented. A phase II clinical trial (NCT02091960) was conducted to evaluate the therapeutic value of the AR antagonist enzalutamide. In the trial, enzalutamide was used on advanced ERα-ve/AR+ve/HER2+ve breast cancer patients. In all, 24% of the patients exhibited a clinical benefit at 24 weeks [49]. A case report described a 55-year-old patient with metastatic TNBC with 100% of nuclei showing a positive signal for AR staining revealed by IHC. The patient received seven lines of cytotoxic chemotherapy and radiotherapy for which only a partial response could be reached, and the disease progressed. The anti-androgen drug bicalutamide was then introduced, to which the patient achieved a complete clinical response at four months and sustained for eight more months [50].
Another phase II study (NCT00468715) demonstrated bicalutamide in AR+ve/ERα-ve/PgR-ve metastatic breast cancer. The results showed a clinical benefit rate of 19% at 6 months and a median progression-free survival of 12 weeks, comparable to other chemotherapies studies in TNBC patients [51]. Similarly, a study (NCT01889238) evaluated the effect of enzalutamide on locally advanced or metastatic AR+ve/TNBC. A clinical benefit rate of 33% was achieved with improved progression-free survival (3.3 months) and overall survival (17.6 months) [52].
An ongoing phase II trial (NCT03383679) combined an androgen-receptor antagonist darolutamide with a chemo-drug capecitabine in advanced AR+ve/TNBC patients recently reported the first-stage result of the study that darolutamide is well tolerated and 26.3% of the patients present a clinical benefit at 16 weeks. The project is now moving to its second stage [53]. Another ongoing phase IIB trial project (NCT02689427) applying enzalutamide in combination with the chemo-drug paclitaxel to treat AR+ve/TNBC patients reported that 33.3% of the patients who were resistant to the doxorubicin-based drugs showed complete response or minimal residual disease to the treatment. This value was lowered to 23% in patients with luminal androgen receptor (LAR). The results suggested a unique pathological feature in this sub-population; noteworthy, around 85% of the LAR TNBC patients were detected to have aberrant PI3K pathway activation in the study. Hence, additional PI3K-targeted drugs for patients who belong to this specific subtype may significantly enhance the treatment outcomes [54].
Other than AR inhibitory drugs, agents targeting androgen synthesis also showed promising potential in clinical application for TNBC patient management. Abiraterone acetate is a CYP17A1-specific inhibitor that suppresses the production of androgens [55]. Combined abiraterone acetate and prednisone for treating locally advanced or metastatic AR+ve/TNBC patients in a phase II trial (NCT01842321) have been tested. The results demonstrated that the clinical benefit rate at 6 months was 20.0%, with one subject having a complete response and five subjects with stable disease for more than 6 months. The median progression-free survival was 2.8 months [56].
A novel antiandrogenic drug seviteronel which functions by a unique dual mechanism of action that lowers the biosynthesis of androgen via inhibiting CYP17 lyase, as well as blocking AR activation via antagonizing AR, has attracted great attention since the drug is more effective than abiraterone acetate in suppressing AR activity [57]. An open-label phase I study proved that seviteronel was well-tolerated at the dose of 450 mg once daily. Of the enrolled TNBC patients, 28.6% (2/7) under this dose reached clinical benefit at 4 months, indicating the drug’s safety and efficacy in treating the patients [58]. In stage I of the phase II study (NCT02580448), the preliminary result showed that 33% of the TNBC subjects could achieve a clinical benefit at 4 months, which was similar to the study mentioned above; moreover, 70% of the subjects presented with decreased circulating tumor cells after the treatment. The results confirmed the clinical activity of seviteronel in breast cancer therapy [59]. Furthermore, multiple clinical trials (NCT03004534, NCT02457910, NCT03207529, NCT03090165, NCT05095207, NCT01990209, NCT04947189) are underway investigating the feasibility of AR-targeting therapy, alone or combined with other therapeutic agents, in AR+ve/ERα-ve breast cancer patients. These results highlight the feasibility of using anti-androgen agents for treating ER-ve breast cancer.

3.2.2. Evidence from Pre-Clinical Studies

In pre-clinical studies, blocking AR signaling in ERα-ve breast cancer models has been shown to suppress cell growth. Inhibition of AR activity by shRNA or enzalutamide could interfere with cell proliferation in AR+ve/ERα-ve/HER2+ve breast cancer cell lines and xenograft models; when combined with trastuzumab, it could further enhance the suppressive capacity. Enzalutamide could reduce the expression of the proliferation marker Ki-67 and induce the expression of active caspase-3 in the xenograft model, demonstrating an anti-proliferative property after AR suppression in HER2+ve breast cancer [60].
A combination of enzalutamide and trastuzumab or mTOR inhibitor everolimus in AR+ve/ERα-ve/HER2+ve and AR+ve/TNBC could synergistically inhibit cell proliferation [61]. Additionally, inhibition of AR in a trastuzumab-resistant HER2+ve breast cancer cell line has been shown to suppress cell proliferation effectively and re-sensitize cells to trastuzumab [61]. It has been shown that concurrent treatment of TNBC cells with bicalutamide and EGFR, PDGFRβ and Erk1/2 inhibitors or PI3K inhibitor could additively inhibit cell proliferation; suppression of the PI3K/mTOR pathway in the cells was demonstrated to decrease the expression of the AR, which indicates a potential direction for anti-androgen drug development for treatment of TNBC patients [62,63].
Blocking the AR could significantly suppress proliferation and increase apoptosis in both in vitro and in vivo AR+ve/ERα-ve models, which might result from the inhibition of Wnt/β-catenin and EGFR signaling pathways [64,65]. ERα-ve cases tend to have more aggressive clinical features, such as higher tumor grade and metastatic nature. Unfortunately, the option for available targeted therapies for this population is limited. Patients who initially responded to treatment quickly relapse later. Therefore, the patients usually have a poor prognosis [66]. The introduction of anti-androgen drugs in these subtypes of patients may help to relieve this predicament.

4. Challenges in Targeting AR for Breast Cancer Treatment

4.1. Lacking Consensus on Defining the AR Positivity in Breast Cancer Hampers the Clinical Diagnosis of the Marker in the Patients

Unlike other prognostic and predictive factors such as ERα that have been standardized for detection in breast cancer, the AR detection protocol to define positivity in patients has not been standardized. The cut-off scores for pathological examination by immunohistochemistry (IHC) of AR positivity varied in different studies. Moreover, the antibodies against the AR used for IHC staining in different studies have been diverse (Table 1). Other studies determined AR positivity by assessing the mRNA of the AR [67,68]. These could very well be the reasons that has led to contradictory conclusions with resulting controversy amongst studies defining the role of the AR in breast cancer patients. Before introducing AR-targeting therapies for breast cancer treatment in the clinical setting, a standardized procedure for AR detection and an agreed definition for AR positivity needs to be established.
Table 1. The definition of AR positivity as determined by IHC, and the antibodies used varied in different studies.
Table 1. The definition of AR positivity as determined by IHC, and the antibodies used varied in different studies.
Anti-AR AntibodyDefinitions of AR PositivityCompany; Catalog NumberReferences
AR441Nuclear stained ≥ 10%Thermo Scientific, Fremont, CA, USA; MA5-13426[25]
Nuclear stained ≥ 1%Dako, Glostrup, Denmark; M3562[24,26,30,32,67,69]
H-score > 10[70]
Nuclear stained ≥ 75%[71]
Nuclear stained ≥ 10%[72]
F39.4.1H-score ≥ 190BioGenex, Fremont, CA, USA; AM256-5ME[23]
≥1% #[73]
AR27Nuclear stained ≥ 10%Leica, Newcastle, UK; AR-318-L-CE[74]
SP107Nuclear stained ≥ 1% and ≥10%Cell Marque, Rocklin, CA, USA; 200R-14[75]
EP120Nuclear stained ≥ 10%ZSGB, Beijing, China; ZA-0554[76]
EPR1535 (2)≥10% #Abcam, Cambridge, UK; ab133273[77]
ER179 (2)Nuclear stained ≥ 1%Epitomics, Burlingame, CA, USA; 3184-1[31]
# Indicates that AR expression was not defined in either nuclear or cytoplasmic compartments in the original study.

4.2. The AR Splice Variants (AR-Vs) Might Affect the Outcomes of AR-Targeting Therapies

Alternative splicing of the AR can give rise to variants, some with truncation of the LBD. These variants may be innately and constitutively active even in the absence of ligands. Based on current studies, at least 18 AR-Vs have been identified (Figure 2), many of which have been correlated to castration resistance, tumor metastasis and subsequently poor OS and DFS in prostate patients, leading to failure of treatment [78,79]. For example, AR-V7, a constitutively active AR variant, has been implicated as the most clinically relevant splice variant in prostate cancer patients. The patients are usually resistant to anti-AR therapy and have poor prognoses [80,81]. Moreover, evidence suggests that targeting AR-V7 may effectively re-sensitize the tumor cell to AR-targeting treatments or inhibit tumor cell proliferation, supporting the potential application of targeting AR-Vs for improving treatment outcome in AR-dominant cancer patients [82,83,84,85,86]. Hence, to address this issue, agents for targeting AR-Vs may need to be considered for the development of AR-targeting strategies in the treatment of breast cancers.
Fortunately, several strategies for targeting AR-Vs have been established [87] for the development of anti-AR-V drugs for which clinical trials have been conducted. For instance, TAS3681, a selective AR antagonist that can reduce the protein expression levels of AR and AR-V7 [88], has been tested in metastatic castration-resistant prostate cancer (NCT02566772). The efficacy of niclosamide, an AR-V7 inhibitor [83], in combination with other anti-prostate cancer agents, has also been tested in advanced prostate cancer (NCT03123978) and hormone-resistant prostate cancer (NCT02807805). However, these agents are not yet available in the clinic.
Figure 2. The wild-type AR (full-length) and its alternative splice variants. The full-length AR (AR-FL) gene harbors eight canonical exons, exons 1–8 and seven cryptic exons (CEs), 1b, CE1-5 and CE9. Different transcriptional manners of the gene generate AR-FL and at least eighteen AR-Vs, including AR-V1 [89], AR-V2 [90], AR-V3 [90], AR-V4 [89], AR-V5 [90], AR-V6 [90], AR-V7 [89], AR-V8 [91], AR-V9 [91], AR-V10 [91], AR-V11 [91], AR-V12 [92], AR-V13 [92], AR-V14 [92], AR8 [93], AR23 [94], AR45 [95] and ARΔ3 [96]. Among them, AR-V5 and AR-V6 have similar structures but differ by an 80 bp contiguous 5′ extension in the CE2 of AR-V5 [90]; AR-V8, AR-V10 and AR-V11 also have similar structures but differ by their downstream 3′ sequence with the proteins truncated after exon 3 with 10–39 amino acids extension before the stop codon [91]. AR8 and AR23 contain a unique sequence (presented by a small black square in the figure) with 23 amino acids inserted in the DBD [93,94].
Figure 2. The wild-type AR (full-length) and its alternative splice variants. The full-length AR (AR-FL) gene harbors eight canonical exons, exons 1–8 and seven cryptic exons (CEs), 1b, CE1-5 and CE9. Different transcriptional manners of the gene generate AR-FL and at least eighteen AR-Vs, including AR-V1 [89], AR-V2 [90], AR-V3 [90], AR-V4 [89], AR-V5 [90], AR-V6 [90], AR-V7 [89], AR-V8 [91], AR-V9 [91], AR-V10 [91], AR-V11 [91], AR-V12 [92], AR-V13 [92], AR-V14 [92], AR8 [93], AR23 [94], AR45 [95] and ARΔ3 [96]. Among them, AR-V5 and AR-V6 have similar structures but differ by an 80 bp contiguous 5′ extension in the CE2 of AR-V5 [90]; AR-V8, AR-V10 and AR-V11 also have similar structures but differ by their downstream 3′ sequence with the proteins truncated after exon 3 with 10–39 amino acids extension before the stop codon [91]. AR8 and AR23 contain a unique sequence (presented by a small black square in the figure) with 23 amino acids inserted in the DBD [93,94].
Ijms 23 15342 g002

4.3. High Expression Levels of AR in ERα+ve Patients May also Lead to Poor Prognosis in Disease Management

The AR tends to be an ERα suppressor in breast cancer. Activating the AR properly in AR+ve/ERα+ve patients to counteract the activity of ERα should be a prioritized option for controlling disease progression. However, without the influence of ERα, such as in ERα-ve breast cancers, the AR can also promote disease progression. It has been reported that breast tumors with an AR-to-ERα (AR/ERα) ratio greater than or equal to 2 have higher expression of proliferation signatures (AURKA, BIRC5, CCNB1, MKI67 and UBE2C) than those with a value less than 2 [97]. Aggressive clinical features such as lymph node metastasis, larger tumor size or higher histopathological grade [98] and poorer response to tamoxifen have been associated with a high AR-to-ERα ratio [99].
Moreover, independent studies focused on different AR/ERα ratios also observed similar results. For instance, patients with a higher AR/ERα ratio value have a poorer prognosis, shorter disease-specific survival (DSS) and a higher risk of tumor aggression in breast cancer clinical studies [100,101]. Although activation of AR has been shown to suppress ERα+ve breast cancer by counteracting the activity of ERα, we should bear in mind that AR could have an oncogenic effect, especially in those patients with higher values of AR/ERα ratio. In such a condition, the tumor-driving ability of AR may become dominant. Thus, maintaining the balance between the signal intensity of AR and ERα pathways to reach an optimal therapeutic effect will need serious consideration for the further application of targeting the AR in breast cancer treatment.

4.4. The Current Molecular Subtyping of Breast Cancer May Not Reflect the Role of AR Accurately in Breast Cancer

The different roles of AR in ERα+ve and ERα-ve breast cancers have been discussed separately. As a heterogeneous cancer, the different intrinsic molecular profiles of breast cancer provide the basis for breast cancer molecular subtyping, which largely determines treatment strategy for each patient. Although the AR is expressed in a considerable proportion of breast cancer, due to its undetermined role, the expression status of AR has not been included in the current molecular subtyping. However, novel molecular subtypes of TNBC involving the AR have identified a subtype called the luminal androgen receptor (LAR) [102]. Interestingly, this subtype demonstrated the worst response to neoadjuvant chemotherapy but the best prognosis amongst the different TNBC subtypes [103], despite its AR being highly expressed (>10-fold). Moreover, as mentioned earlier, high expression of AR in ERα+ve breast cancer patients (high AR/ERα ratio) are associated with worse prognosis. These findings indicate that the current molecular subtyping system for breast cancer might not be sufficient to define the roles of the AR in breast cancer. Hence, to achieve an optimal therapeutic outcome by targeting the AR, a more detailed breast cancer molecular subtyping needs to be proposed, with the expression status of the AR in breast cancer taken into account.

4.5. Side Effects of Modulating AR Activity

Since the AR plays a critical role in multiple normal biological processes, it is not surprising that treating breast cancer patients with AR modulators, either agonist or antagonist, will alter the activity of the AR. This may have adverse effects. AR agonists have been reported to induce acne, hirsutism and alteration in lipid profile [104], while AR antagonists may cause adverse effects such as amenorrhea, hyperkalemia, gynecomastia and thromboembolism [105]. These side effects may decrease the patient’s quality of life or even be life threatening. Before using these agents for treating breast cancer, we must know if the patients will gain a benefit that can compensate for the side-effects.

5. Future Perspectives

Currently, many conventional AR-targeting drugs that modulate the activity of AR signaling can be applied in cancer therapy. Patients who benefit from these drugs have largely had their survival time prolonged, with improved disease treatment outcomes, regardless of their standard treatment for prostate cancer or their clinical trial treatment for breast cancer. In addition, more novel drugs produced by advanced techniques are in development. The proteolysis-targeting chimaera (PROTAC) molecule was designed for utilizing the ubiquitin–proteasome system (UPS) to degrade specific proteins by its bound E3 ubiquitin ligase. The model was proposed by Sakamoto et al. in 2001 [106], and subsequently rapidly and widely deployed for developing multiple new drugs, including anti-AR agents. ARV-110 (also known as bavdegalutamide) is one of them, which has exhibited a satisfactory effect in suppressing tumor growth and overcoming anti-AR drug resistance [107]. An ongoing phase I/II clinical trial has proved the safety of ARV-110 (NCT03888612) and its effect on cancer treatment (NCT05177042). More AR PROTAC degraders such as ARD-2585, ARD-61, ARCC-4 and A031 have also demonstrated anti-tumor potentials in pre-clinical studies [108,109,110,111].
Targeting AR by N-terminal inhibitors is another possible strategy. Some constitutively activated AR-Vs may lack the LBD but retain an intact NTD. Inhibitors targeting the NTD of the AR may inhibit not only the function of the entire length of the AR, but also the function of AR-Vs. One of the AR NTD inhibitors are the EPI compounds. In vitro and in vivo studies demonstrated that these compounds could suppress the growth of different prostate cancer models [83,87,88]. Clinical trials (NCT05075577, NCT04421222) have been proposed for metastatic or castration-resistant prostate cancer patients. Nuclear translocation inhibitors to prevent the AR from genomic transcription is another strategy to inhibit the function of the AR. As a transcriptional factor, nuclear translocation is the prerequisite for the action of the AR. Small molecules such as EPPI, CPPI and JJ-450 can suppress the nuclear translocation of the AR to block its transcriptional activity. These drugs have been evaluated to be effective in suppressing the proliferation of prostate cancer cells [86,112], but the effect on breast cancer remains unaddressed. Testing the effect of these chemicals on breast cancer may bring new insights into breast cancer treatment.

6. Conclusions

The roles of the AR should be clarified in different subtypes of breast cancer. Modulating the activity of the AR will be a promising strategy for treating breast cancer. However, more studies are needed to address the existing issues in the hope of translating these strategies into the clinical setting.

Author Contributions

Conceptualization, C.-P.Y., H.T. and U.-S.K.; Funding acquisition, H.T. and U.-S.K.; Project administration, E.P.S.M. and U.-S.K.; Visualization, C.-P.Y. and U.-S.K.; Writing—original draft, C.-P.Y. and U.-S.K.; Writing—review and editing, C.-P.Y., H.T., E.P.S.M., M.-H.L. and U.-S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Health and Medical Research Fund (07182026; 07182046), Research Grants Council, Hong Kong General Research Fund (17113020), and the Committee on Research and Conference Grants from the University of Hong Kong (202111159194).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The materials and resources in this study are available from the corresponding author upon reasonable request.

Acknowledgments

U.-S.K. is a holder of the Ada Chan Professorship in Oncological Pathology, the University of Hong Kong.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Davey, R.A.; Grossmann, M. Androgen Receptor Structure, Function and Biology: From Bench to Bedside. Clin. Biochem. Rev. 2016, 37, 3–15. [Google Scholar] [PubMed]
  2. Gerratana, L.; Basile, D.; Buono, G.; De Placido, S.; Giuliano, M.; Minichillo, S.; Coinu, A.; Martorana, F.; De Santo, I.; Del Mastro, L.; et al. Androgen receptor in triple negative breast cancer: A potential target for the targetless subtype. Cancer Treat. Rev. 2018, 68, 102–110. [Google Scholar] [CrossRef]
  3. Echeverria, P.C.; Picard, D. Molecular chaperones, essential partners of steroid hormone receptors for activity and mobility. Biochim. Biophys. Acta 2010, 1803, 641–649. [Google Scholar] [CrossRef] [PubMed]
  4. Jin, H.J.; Kim, J.; Yu, J. Androgen receptor genomic regulation. Transl. Androl. Urol. 2013, 2, 157–177. [Google Scholar] [CrossRef] [PubMed]
  5. Mulholland, D.J.; Cheng, H.; Reid, K.; Rennie, P.S.; Nelson, C.C. The androgen receptor can promote beta-catenin nuclear translocation independently of adenomatous polyposis coli. J. Biol. Chem. 2002, 277, 17933–17943. [Google Scholar] [CrossRef] [Green Version]
  6. Sarker, D.; Reid, A.H.; Yap, T.A.; de Bono, J.S. Targeting the PI3K/AKT pathway for the treatment of prostate cancer. Clin. Cancer Res. 2009, 15, 4799–4805. [Google Scholar] [CrossRef] [Green Version]
  7. Traish, A.M.; Morgentaler, A. Epidermal growth factor receptor expression escapes androgen regulation in prostate cancer: A potential molecular switch for tumour growth. Br. J. Cancer 2009, 101, 1949–1956. [Google Scholar] [CrossRef] [Green Version]
  8. Varkaris, A.; Katsiampoura, A.D.; Araujo, J.C.; Gallick, G.E.; Corn, P.G. Src signaling pathways in prostate cancer. Cancer Metast. Rev. 2014, 33, 595–606. [Google Scholar] [CrossRef] [Green Version]
  9. Ahmad, N.; Kumar, R. Steroid hormone receptors in cancer development: A target for cancer therapeutics. Cancer Lett. 2011, 300, 1–9. [Google Scholar] [CrossRef]
  10. Aurilio, G.; Cimadamore, A.; Mazzucchelli, R.; Lopez-Beltran, A.; Verri, E.; Scarpelli, M.; Massari, F.; Cheng, L.; Santoni, M.; Montironi, R. Androgen Receptor Signaling Pathway in Prostate Cancer: From Genetics to Clinical Applications. Cells 2020, 9, 2653. [Google Scholar] [CrossRef]
  11. Crumbaker, M.; Khoja, L.; Joshua, A.M. AR Signaling and the PI3K Pathway in Prostate Cancer. Cancers 2017, 9, 34. [Google Scholar] [CrossRef] [PubMed]
  12. Schweizer, M.T.; Yu, E.Y. AR-Signaling in Human Malignancies: Prostate Cancer and Beyond. Cancers 2017, 9, 7. [Google Scholar] [CrossRef] [Green Version]
  13. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, Z.Q.; Zhang, X.S.; Zhang, S.H. Breast tumor subgroups reveal diverse clinical prognostic power. Sci. Rep. 2014, 4, 4002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Collins, L.C.; Cole, K.S.; Marotti, J.D.; Hu, R.; Schnitt, S.J.; Tamimi, R.M. Androgen receptor expression in breast cancer in relation to molecular phenotype: Results from the Nurses’ Health Study. Mod. Pathol. 2011, 24, 924–931. [Google Scholar] [CrossRef] [Green Version]
  16. Iacopetta, D.; Rechoum, Y.; Fuqua, S.A. The Role of Androgen Receptor in Breast Cancer. Drug Discov. Today Dis. Mech. 2012, 9, e19–e27. [Google Scholar] [CrossRef] [Green Version]
  17. Basile, D.; Cinausero, M.; Iacono, D.; Pelizzari, G.; Bonotto, M.; Vitale, M.G.; Gerratana, L.; Puglisi, F. Androgen receptor in estrogen receptor positive breast cancer: Beyond expression. Cancer Treat. Rev. 2017, 61, 15–22. [Google Scholar] [CrossRef]
  18. Hickey, T.E.; Robinson, J.L.; Carroll, J.S.; Tilley, W.D. Minireview: The androgen receptor in breast tissues: Growth inhibitor, tumor suppressor, oncogene? Mol. Endocrinol. 2012, 26, 1252–1267. [Google Scholar] [CrossRef]
  19. Hickey, T.E.; Selth, L.A.; Chia, K.M.; Laven-Law, G.; Milioli, H.H.; Roden, D.; Jindal, S.; Hui, M.; Finlay-Schultz, J.; Ebrahimie, E.; et al. The androgen receptor is a tumor suppressor in estrogen receptor-positive breast cancer. Nat. Med. 2021, 27, 310–320. [Google Scholar] [CrossRef]
  20. Rizza, P.; Barone, I.; Zito, D.; Giordano, F.; Lanzino, M.; De Amicis, F.; Mauro, L.; Sisci, D.; Catalano, S.; Wright, K.D.; et al. Estrogen receptor beta as a novel target of androgen receptor action in breast cancer cell lines. Breast Cancer Res. 2014, 16, R21. [Google Scholar] [CrossRef]
  21. Datta, J.; Willingham, N.; Manouchehri, J.M.; Schnell, P.; Sheth, M.; David, J.J.; Kassem, M.; Wilson, T.A.; Radomska, H.S.; Coss, C.C.; et al. Activity of Estrogen Receptor beta Agonists in Therapy-Resistant Estrogen Receptor-Positive Breast Cancer. Front. Oncol. 2022, 12, 857590. [Google Scholar] [CrossRef] [PubMed]
  22. Park, S.; Koo, J.; Park, H.S.; Kim, J.H.; Choi, S.Y.; Lee, J.H.; Park, B.W.; Lee, K.S. Expression of androgen receptors in primary breast cancer. Ann. Oncol. 2010, 21, 488–492. [Google Scholar] [CrossRef]
  23. Aleskandarany, M.A.; Abduljabbar, R.; Ashankyty, I.; Elmouna, A.; Jerjees, D.; Ali, S.; Buluwela, L.; Diez-Rodriguez, M.; Caldas, C.; Green, A.R.; et al. Prognostic significance of androgen receptor expression in invasive breast cancer: Transcriptomic and protein expression analysis. Breast Cancer Res. Treat. 2016, 159, 215–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hu, R.; Dawood, S.; Holmes, M.D.; Collins, L.C.; Schnitt, S.J.; Cole, K.; Marotti, J.D.; Hankinson, S.E.; Colditz, G.A.; Tamimi, R.M. Androgen receptor expression and breast cancer survival in postmenopausal women. Clin. Cancer Res. 2011, 17, 1867–1874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Park, S.; Koo, J.S.; Kim, M.S.; Park, H.S.; Lee, J.S.; Lee, J.S.; Kim, S.I.; Park, B.W.; Lee, K.S. Androgen receptor expression is significantly associated with better outcomes in estrogen receptor-positive breast cancers. Ann. Oncol. 2011, 22, 1755–1762. [Google Scholar] [CrossRef]
  26. Kensler, K.H.; Poole, E.M.; Heng, Y.J.; Collins, L.C.; Glass, B.; Beck, A.H.; Hazra, A.; Rosner, B.A.; Eliassen, A.H.; Hankinson, S.E.; et al. Androgen Receptor Expression and Breast Cancer Survival: Results From the Nurses’ Health Studies. JNCI J. Natl. Cancer Inst. 2019, 111, 700–708. [Google Scholar] [CrossRef]
  27. Chia, K.M.; Liu, J.; Francis, G.D.; Naderi, A. A Feedback Loop between Androgen Receptor and ERK Signaling in Estrogen Receptor-Negative Breast Cancer. Neoplasia 2011, 13, 154–166. [Google Scholar] [CrossRef] [Green Version]
  28. Ni, M.; Chen, Y.W.; Lim, E.; Wimberly, H.; Bailey, S.T.; Imai, Y.; Rimm, D.L.; Liu, X.S.; Brown, M. Targeting Androgen Receptor in Estrogen Receptor-Negative Breast Cancer. Cancer Cell 2011, 20, 119–131. [Google Scholar] [CrossRef] [Green Version]
  29. Giovannelli, P.; Di Donato, M.; Auricchio, F.; Castoria, G.; Migliaccio, A. Androgens Induce Invasiveness of Triple Negative Breast Cancer Cells Through AR/Src/PI3-K Complex Assembly. Sci. Rep. 2019, 9, 4490. [Google Scholar] [CrossRef] [Green Version]
  30. Asano, Y.; Kashiwagi, S.; Onoda, N.; Kurata, K.; Morisaki, T.; Noda, S.; Takashima, T.; Ohsawa, M.; Kitagawa, S.; Hirakawa, K. Clinical verification of sensitivity to preoperative chemotherapy in cases of androgen receptor-expressing positive breast cancer. Br. J. Cancer 2016, 114, 14–20. [Google Scholar] [CrossRef]
  31. Choi, J.E.; Kang, S.H.; Lee, S.J.; Bae, Y.K. Androgen Receptor Expression Predicts Decreased Survival in Early Stage Triple-Negative Breast Cancer. Ann. Surg. Oncol. 2015, 22, 82–89. [Google Scholar] [CrossRef] [PubMed]
  32. Dieci, M.V.; Tsvetkova, V.; Griguolo, G.; Miglietta, F.; Mantiero, M.; Tasca, G.; Cumerlato, E.; Giorgi, C.A.; Giarratano, T.; Faggioni, G.; et al. Androgen Receptor Expression and Association with Distant Disease-Free Survival in Triple Negative Breast Cancer: Analysis of 263 Patients Treated with Standard Therapy for Stage I-III Disease. Front. Oncol. 2019, 9, 452. [Google Scholar] [CrossRef]
  33. Venema, C.M.; Bense, R.D.; Steenbruggen, T.G.; Nienhuis, H.H.; Qiu, S.Q.; van Kruchten, M.; Brown, M.; Tamimi, R.M.; Hospers, G.A.P.; Schroder, C.P.; et al. Consideration of breast cancer subtype in targeting the androgen receptor. Pharmacol. Therapeut. 2019, 200, 135–147. [Google Scholar] [CrossRef] [PubMed]
  34. Kaae, S. Testosterone propionate in the treatment of breast cancer. Acta Radiol. 1949, 31, 97–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Gordan, G.S.; Halden, A.; Horn, Y.; Fuery, J.J.; Parsons, R.J.; Walter, R.M. Calusterone (7-Beta, 17-Alpha-Dimethyltestosterone) as Primary and Secondary Therapy of Advanced Breast-Cancer. Oncology 1973, 28, 138–146. [Google Scholar] [CrossRef]
  36. Goldenberg, I.S. Testosterone Propionate Therapy in Breast Cancer. JAMA 1964, 188, 1069–1072. [Google Scholar] [CrossRef]
  37. Kennedy, B.J. Fluoxymesterone therapy in advanced breast cancer. N. Engl. J. Med. 1958, 259, 673–675. [Google Scholar] [CrossRef]
  38. Fels, E. Treatment of Breast Cancer with Testosterone Propionate. J. Clin. Endocrinol. Metab. 1944, 4, 121–125. [Google Scholar] [CrossRef]
  39. Dimitrakakis, C.; Jones, R.A.; Liu, A.; Bondy, C.A. Breast cancer incidence in postmenopausal women using testosterone in addition to usual hormone therapy. Menopause 2004, 11, 531–535. [Google Scholar] [CrossRef]
  40. Glaser, R.L.; Dimitrakakis, C. Reduced breast cancer incidence in women treated with subcutaneous testosterone, or testosterone with anastrozole: A prospective, observational study. Maturitas 2013, 76, 342–349. [Google Scholar] [CrossRef]
  41. Glaser, R.L.; York, A.E.; Dimitrakakis, C. Incidence of invasive breast cancer in women treated with testosterone implants: A prospective 10-year cohort study. BMC Cancer 2019, 19, 1271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Glaser, R.L.; York, A.E.; Dimitrakakis, C. Efficacy of subcutaneous testosterone on menopausal symptoms In breast cancer survivors. J. Clin. Oncol. 2014, 32, 109. [Google Scholar] [CrossRef]
  43. Glaser, R.L.; Dimitrakakis, C. Rapid response of breast cancer to neoadjuvant intramammary testosterone-anastrozole therapy: Neoadjuvant hormone therapy in breast cancer. Menopause 2014, 21, 673–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Glaser, R.L.; York, A.E.; Dimitrakakis, C. Subcutaneous testosterone-letrozole therapy before and concurrent with neoadjuvant breast chemotherapy: Clinical response and therapeutic implications. Menopause-J. N. Am. Menopause Soc. 2017, 24, 859–864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Boni, C.; Pagano, M.; Panebianco, M.; Bologna, A.; Sierra, N.M.; Gnoni, R.; Formisano, D.; Bisagni, G. Therapeutic activity of testosterone in metastatic breast cancer. Anticancer Res. 2014, 34, 1287–1290. [Google Scholar]
  46. Kono, M.; Fujii, T.; Lyons, G.R.; Huo, L.; Bassett, R.; Gong, Y.; Karuturi, M.S.; Tripathy, D.; Ueno, N.T. Impact of androgen receptor expression in fluoxymesterone-treated estrogen receptor-positive metastatic breast cancer refractory to contemporary hormonal therapy. Breast Cancer Res. Treat. 2016, 160, 101–109. [Google Scholar] [CrossRef]
  47. Overmoyer, B.; Sanz-Altamira, P.; Taylor, R.P.; Hancock, M.L.; Dalton, J.T.; Johnston, M.A.; Steiner, M.S. Enobosarm: A targeted therapy for metastatic, androgen receptor positive, breast cancer. J. Clin. Oncol. 2014, 32, 568. [Google Scholar] [CrossRef]
  48. Palmieri, C.; Linden, H.M.; Birrell, S.; Lim, E.; Schwartzberg, L.S.; Rugo, H.S.; Cobb, P.W.; Jain, K.; Vogel, C.L.; O’Shaughnessy, J.; et al. Efficacy of enobosarm, a selective androgen receptor (AR) targeting agent, correlates with the degree of AR positivity in advanced AR plus /estrogen receptor (ER) plus breast cancer in an international phase 2 clinical study. J. Clin. Oncol. 2021, 39, 1020. [Google Scholar] [CrossRef]
  49. Wardley, A.; Cortes, J.; Provencher, L.; Miller, K.; Chien, A.J.; Rugo, H.S.; Steinberg, J.; Sugg, J.; Tudor, I.C.; Huizing, M.; et al. The efficacy and safety of enzalutamide with trastuzumab in patients with HER2+ and androgen receptor-positive metastatic or locally advanced breast cancer. Breast Cancer Res. Treat. 2021, 187, 155–165. [Google Scholar] [CrossRef]
  50. Arce-Salinas, C.; Riesco-Martinez, M.C.; Hanna, W.; Bedard, P.; Warner, E. Complete Response of Metastatic Androgen Receptor-Positive Breast Cancer to Bicalutamide: Case Report and Review of the Literature. J. Clin. Oncol. 2016, 34, e21–e24. [Google Scholar] [CrossRef]
  51. Gucalp, A.; Tolaney, S.; Isakoff, S.J.; Ingle, J.N.; Liu, M.C.; Carey, L.A.; Blackwell, K.; Rugo, H.; Nabell, L.; Forero, A.; et al. Phase II trial of bicalutamide in patients with androgen receptor-positive, estrogen receptor-negative metastatic Breast Cancer. Clin. Cancer Res. 2013, 19, 5505–5512. [Google Scholar] [CrossRef] [PubMed]
  52. Traina, T.A.; Miller, K.; Yardley, D.A.; Eakle, J.; Schwartzberg, L.S.; O’Shaughnessy, J.; Gradishar, W.; Schmid, P.; Winer, E.; Kelly, C.; et al. Enzalutamide for the Treatment of Androgen Receptor-Expressing Triple-Negative Breast Cancer. J. Clin. Oncol. 2018, 36, 884–890. [Google Scholar] [CrossRef]
  53. Bonnefoi, H.; Lerebours, F.; Tredan, O.; Dalenc, F.; Levy, C.; Saghatchian, M.; Reynier, M.A.M.; Mollon, D.; Guiu, S.; Bouvet, L.V.; et al. First efficacy results of a 2-stage Simon’s design randomised phase 2 of darolutamide or capecitabine in patients with triplenegative, androgen receptor positive advanced breast cancer (UCBG06-3). Cancer Res. 2021, 81, PS12-05. [Google Scholar] [CrossRef]
  54. Lim, B.; Seth, S.; Huo, L.; Layman, R.M.; Valero, V.; Thompson, A.M.; White, J.B.; Litton, J.K.; Damodaran, S.; Candelaria, R.P.; et al. Comprehensive profiling of androgen receptor-positive (AR plus) triple-negative breast cancer (TNBC) patients (pts) treated with standard neoadjuvant therapy (NAT) +/− enzalutamide. J. Clin. Oncol. 2020, 38, 517. [Google Scholar] [CrossRef]
  55. Bryce, A.; Ryan, C.J. Development and Clinical Utility of Abiraterone Acetate as an Androgen Synthesis Inhibitor. Clin. Pharmacol. Ther. 2012, 91, 101–108. [Google Scholar] [CrossRef]
  56. Bonnefoi, H.; Grellety, T.; Tredan, O.; Saghatchian, M.; Dalenc, F.; Mailliez, A.; L’Haridon, T.; Cottu, P.; Abadie-Lacourtoisie, S.; You, B.; et al. A phase II trial of abiraterone acetate plus prednisone in patients with triple-negative androgen receptor positive locally advanced or metastatic breast cancer (UCBG 12-1). Ann. Oncol. 2016, 27, 812–818. [Google Scholar] [CrossRef]
  57. Toren, P.J.; Kim, S.; Pham, S.; Mangalji, A.; Adomat, H.; Guns, E.S.; Zoubeidi, A.; Moore, W.; Gleave, M.E. Anticancer activity of a novel selective CYP17A1 inhibitor in preclinical models of castrate-resistant prostate cancer. Mol. Cancer Ther. 2015, 14, 59–69. [Google Scholar] [CrossRef] [Green Version]
  58. Bardia, A.; Gucalp, A.; DaCosta, N.; Gabrail, N.; Danso, M.; Ali, H.; Blackwell, K.L.; Carey, L.A.; Eisner, J.R.; Baskin-Bey, E.S.; et al. Phase 1 study of seviteronel, a selective CYP17 lyase and androgen receptor inhibitor, in women with estrogen receptor-positive or triple-negative breast cancer. Breast Cancer Res. Treat. 2018, 171, 111–120. [Google Scholar] [CrossRef]
  59. Gucalp, A.; Danso, M.A.; Elias, A.D.; Bardia, A.; Ali, H.Y.; Potter, D.; Gabrail, N.Y.; Haley, B.B.; Khong, H.T.; Riley, E.C.; et al. Phase (Ph) 2 stage 1 clinical activity of seviteronel, a selective CYP17-lyase and androgen receptor (AR) inhibitor, in women with advanced AR plus triple-negative breast cancer (TNBC) or estrogen receptor (ER) plus BC: CLARITY-01. J. Clin. Oncol. 2017, 35, 1102. [Google Scholar] [CrossRef]
  60. Bluemn, E.G.; Coleman, I.M.; Lucas, J.M.; Coleman, R.T.; Hernandez-Lopez, S.; Tharakan, R.; Bianchi-Frias, D.; Dumpit, R.F.; Kaipainen, A.; Corella, A.N.; et al. Androgen Receptor Pathway-Independent Prostate Cancer Is Sustained through FGF Signaling. Cancer Cell 2017, 32, 474–489.e6. [Google Scholar] [CrossRef] [Green Version]
  61. Gordon, M.A.; D’Amato, N.C.; Gu, H.H.; Babbs, B.; Wulfkuhle, J.; Petricoin, E.F.; Gallagher, I.; Dong, T.; Torkko, K.; Liu, B.L.; et al. Synergy between Androgen Receptor Antagonism and Inhibition of mTOR and HER2 in Breast Cancer. Mol. Cancer Ther. 2017, 16, 1389–1400. [Google Scholar] [CrossRef] [PubMed]
  62. Lehmann, B.D.; Bauer, J.A.; Schafer, J.M.; Pendleton, C.S.; Tang, L.; Johnson, K.C.; Chen, X.; Balko, J.M.; Gomez, H.; Arteaga, C.L.; et al. PIK3CA mutations in androgen receptor-positive triple negative breast cancer confer sensitivity to the combination of PI3K and androgen receptor inhibitors. Breast Cancer Res. 2014, 16, 406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cuenca-Lopez, M.D.; Montero, J.C.; Morales, J.C.; Prat, A.; Pandiella, A.; Ocana, A. Phospho-kinase profile of triple negative breast cancer and androgen receptor signaling. BMC Cancer 2014, 14, 302. [Google Scholar] [CrossRef] [Green Version]
  64. Huang, R.; Han, J.; Liang, X.; Sun, S.; Jiang, Y.; Xia, B.; Niu, M.; Li, D.; Zhang, J.; Wang, S.; et al. Androgen Receptor Expression and Bicalutamide Antagonize Androgen Receptor Inhibit beta-Catenin Transcription Complex in Estrogen Receptor-Negative Breast Cancer. Cell Physiol. Biochem. 2017, 43, 2212–2225. [Google Scholar] [CrossRef] [PubMed]
  65. Barton, V.N.; D’Amato, N.C.; Gordon, M.A.; Lind, H.T.; Spoelstra, N.S.; Babbs, B.L.; Heinz, R.E.; Elias, A.; Jedlicka, P.; Jacobsen, B.M.; et al. Multiple Molecular Subtypes of Triple-Negative Breast Cancer Critically Rely on Androgen Receptor and Respond to Enzalutamide In Vivo. Mol. Cancer Ther. 2015, 14, 769–778. [Google Scholar] [CrossRef] [Green Version]
  66. Hoeferlin, L.A.; Chalfant, C.E.; Park, M.A. Challenges in the Treatment of Triple Negative and HER2-Overexpressing Breast Cancer. J. Surg. Sci. 2013, 1, 3–7. [Google Scholar] [PubMed]
  67. Castellano, I.; Allia, E.; Accortanzo, V.; Vandone, A.M.; Chiusa, L.; Arisio, R.; Durando, A.; Donadio, M.; Bussolati, G.; Coates, A.S.; et al. Androgen receptor expression is a significant prognostic factor in estrogen receptor positive breast cancers. Breast Cancer Res. Treat. 2010, 124, 607–617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. You, C.P.; Leung, M.H.; Tsang, W.C.; Khoo, U.S.; Tsoi, H. Androgen Receptor as an Emerging Feasible Biomarker for Breast Cancer. Biomolecules 2022, 12, 72. [Google Scholar] [CrossRef]
  69. Tsang, J.Y.S.; Ni, Y.-B.; Chan, S.-K.; Shao, M.-M.; Law, B.K.B.; Tan, P.H.; Tse, G.M. Androgen Receptor Expression Shows Distinctive Significance in ER Positive and Negative Breast Cancers. Ann. Surg. Oncol. 2014, 21, 2218–2228. [Google Scholar] [CrossRef]
  70. Niemeier, L.A.; Dabbs, D.J.; Beriwal, S.; Striebel, J.M.; Bhargava, R. Androgen receptor in breast cancer: Expression in estrogen receptor-positive tumors and in estrogen receptor-negative tumors with apocrine differentiation. Mod. Pathol. 2010, 23, 205–212. [Google Scholar] [CrossRef] [Green Version]
  71. Tokunaga, E.; Hisamatsu, Y.; Taketani, K.; Yamashita, N.; Akiyoshi, S.; Okada, S.; Tanaka, K.; Saeki, H.; Oki, E.; Aishima, S.; et al. Differential impact of the expression of the androgen receptor by age in estrogen receptor–positive breast cancer. Cancer Med. 2013, 2, 763–773. [Google Scholar] [CrossRef] [PubMed]
  72. Kraby, M.R.; Valla, M.; Opdahl, S.; Haugen, O.A.; Sawicka, J.E.; Engstrøm, M.J.; Bofin, A.M. The prognostic value of androgen receptors in breast cancer subtypes. Breast Cancer Res. Treat. 2018, 172, 283–296. [Google Scholar] [CrossRef] [PubMed]
  73. Loibl, S.; Müller, B.M.; Von Minckwitz, G.; Schwabe, M.; Roller, M.; Darb-Esfahani, S.; Ataseven, B.; Du Bois, A.; Fissler-Eckhoff, A.; Gerber, B.; et al. Androgen receptor expression in primary breast cancer and its predictive and prognostic value in patients treated with neoadjuvant chemotherapy. Breast Cancer Res. Treat. 2011, 130, 477–487. [Google Scholar] [CrossRef] [PubMed]
  74. Micello, D.; Marando, A.; Sahnane, N.; Riva, C.; Capella, C.; Sessa, F. Androgen receptor is frequently expressed in HER2-positive, ER/PR-negative breast cancers. Virchows Arch. 2010, 457, 467–476. [Google Scholar] [CrossRef] [PubMed]
  75. Bronte, G.; Bravaccini, S.; Ravaioli, S.; Puccetti, M.; Scarpi, E.; Andreis, D.; Tumedei, M.M..; Sarti, S.; Cecconetto, L.; Pietri, E.; et al. Androgen Receptor Expression in Breast Cancer: What Differences Between Primary Tumor and Metastases? Transl. Oncol. 2018, 11, 950–956. [Google Scholar] [CrossRef]
  76. Xiang, G.; Liu, F.; Liu, J.; Meng, Q.; Li, N.; Niu, Y. Prognostic role of Amphiregulin and the correlation with androgen receptor in invasive breast cancer. Pathol. Res. Pract. 2019, 215, 152414. [Google Scholar] [CrossRef]
  77. Zhao, S.; Ma, D.; Xiao, Y.; Li, X.-M.; Ma, J.-L.; Zhang, H.; Xu, X.-L.; Lv, H.; Jiang, W.-H.; Yang, W.-T.; et al. Molecular Subtyping of Triple-Negative Breast Cancers by Immunohistochemistry: Molecular Basis and Clinical Relevance. Oncologist 2020, 25, e1481–e1491. [Google Scholar] [CrossRef]
  78. Kanayama, M.; Lu, C.; Luo, J.; Antonarakis, E.S. AR Splicing Variants and Resistance to AR Targeting Agents. Cancers 2021, 13, 2563. [Google Scholar] [CrossRef]
  79. Lu, C.; Luo, J. Decoding the androgen receptor splice variants. Transl. Androl. Urol. 2013, 2, 178–186. [Google Scholar] [CrossRef]
  80. Sobhani, N.; Neeli, P.K.; D’Angelo, A.; Pittacolo, M.; Sirico, M.; Galli, I.C.; Roviello, G.; Nesi, G. AR-V7 in Metastatic Prostate Cancer: A Strategy beyond Redemption. Int. J. Mol. Sci. 2021, 22, 5515. [Google Scholar] [CrossRef]
  81. Wang, Z.; Shen, H.; Liang, Z.; Mao, Y.; Wang, C.; Xie, L. The characteristics of androgen receptor splice variant 7 in the treatment of hormonal sensitive prostate cancer: A systematic review and meta-analysis. Cancer Cell Int. 2020, 20, 149. [Google Scholar] [CrossRef] [PubMed]
  82. Khurana, N.; Chandra, P.K.; Kim, H.; Abdel-Mageed, A.B.; Mondal, D.; Sikka, S.C. Bardoxolone-Methyl (CDDO-Me) Suppresses Androgen Receptor and Its Splice-Variant AR-V7 and Enhances Efficacy of Enzalutamide in Prostate Cancer Cells. Antioxidants 2020, 9, 68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Liu, C.; Lou, W.; Zhu, Y.; Nadiminty, N.; Schwartz, C.T.; Evans, C.P.; Gao, A.C. Niclosamide inhibits androgen receptor variants expression and overcomes enzalutamide resistance in castration-resistant prostate cancer. Clin. Cancer Res. 2014, 20, 3198–3210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Luna Velez, M.V.; Verhaegh, G.W.; Smit, F.; Sedelaar, J.P.M.; Schalken, J.A. Suppression of prostate tumor cell survival by antisense oligonucleotide-mediated inhibition of AR-V7 mRNA synthesis. Oncogene 2019, 38, 3696–3709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Melnyk, J.E.; Steri, V.; Nguyen, H.G.; Hwang, Y.C.; Gordan, J.D.; Hann, B.; Feng, F.Y.; Shokat, K.M. Targeting a splicing-mediated drug resistance mechanism in prostate cancer by inhibiting transcriptional regulation by PKCbeta1. Oncogene 2022, 41, 1536–1549. [Google Scholar] [CrossRef]
  86. Yang, Z.; Wang, D.; Johnson, J.K.; Pascal, L.E.; Takubo, K.; Avula, R.; Chakka, A.B.; Zhou, J.; Chen, W.; Zhong, M.; et al. A Novel Small Molecule Targets Androgen Receptor and Its Splice Variants in Castration-Resistant Prostate Cancer. Mol. Cancer Ther. 2020, 19, 75–88. [Google Scholar] [CrossRef] [Green Version]
  87. Armstrong, C.M.; Gao, A.C. Current strategies for targeting the activity of androgen receptor variants. Asian J. Urol. 2019, 6, 42–49. [Google Scholar] [CrossRef]
  88. Kajiwara, D.; Minamiguchi, K.; Seki, M.; Mizutani, H.; Yamamura, K.; Okajima, S. The impact of TAS3681, a new type of androgen receptor antagonist, on aberrant AR signaling that drives tumor resistance to AR-targeted therapies by down-regulating full length and splice variant AR. J. Clin. Oncol. 2017, 35, 197. [Google Scholar] [CrossRef]
  89. Guo, Z.; Yang, X.; Sun, F.; Jiang, R.; Linn, D.E.; Chen, H.; Chen, H.; Kong, X.; Melamed, J.; Tepper, C.G.; et al. A novel androgen receptor splice variant is up-regulated during prostate cancer progression and promotes androgen depletion-resistant growth. Cancer Res. 2009, 69, 2305–2313. [Google Scholar] [CrossRef] [Green Version]
  90. Hu, R.; Dunn, T.A.; Wei, S.; Isharwal, S.; Veltri, R.W.; Humphreys, E.; Han, M.; Partin, A.W.; Vessella, R.L.; Isaacs, W.B.; et al. Ligand-independent androgen receptor variants derived from splicing of cryptic exons signify hormone-refractory prostate cancer. Cancer Res. 2009, 69, 16–22. [Google Scholar] [CrossRef] [Green Version]
  91. Watson, P.A.; Chen, Y.F.; Balbas, M.D.; Wongvipat, J.; Socci, N.D.; Viale, A.; Kim, K.; Sawyers, C.L. Constitutively active androgen receptor splice variants expressed in castration-resistant prostate cancer require full-length androgen receptor. Proc. Natl. Acad. Sci. USA 2010, 107, 16759–16765. [Google Scholar] [CrossRef] [PubMed]
  92. Hu, R.; Isaacs, W.B.; Luo, J. A snapshot of the expression signature of androgen receptor splicing variants and their distinctive transcriptional activities. Prostate 2011, 71, 1656–1667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Yang, X.; Guo, Z.; Sun, F.; Li, W.; Alfano, A.; Shimelis, H.; Chen, M.; Brodie, A.M.H.; Chen, H.; Xiao, Z.; et al. Novel membrane-associated androgen receptor splice variant potentiates proliferative and survival responses in prostate cancer cells. J. Biol. Chem. 2011, 286, 36152–36160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Jagla, M.; Feve, M.; Kessler, P.; Lapouge, G.; Erdmann, E.; Serra, S.; Bergerat, J.P.; Ceraline, J. A splicing variant of the androgen receptor detected in a metastatic prostate cancer exhibits exclusively cytoplasmic actions. Endocrinology 2007, 148, 4334–4343. [Google Scholar] [CrossRef] [Green Version]
  95. Ahrens-Fath, I.; Politz, O.; Geserick, C.; Haendler, B. Androgen receptor function is modulated by the tissue-specific AR45 variant. FEBS J. 2005, 272, 74–84. [Google Scholar] [CrossRef] [PubMed]
  96. Zhu, X.; Daffada, A.A.; Chan, C.M.; Dowsett, M. Identification of an exon 3 deletion splice variant androgen receptor mRNA in human breast cancer. Int. J. Cancer 1997, 72, 574–580. [Google Scholar] [CrossRef]
  97. Rangel, N.; Rondon-Lagos, M.; Annaratone, L.; Aristizabal-Pachon, A.F.; Cassoni, P.; Sapino, A.; Castellano, I. AR/ER Ratio Correlates with Expression of Proliferation Markers and with Distinct Subset of Breast Tumors. Cells 2020, 9, 1064. [Google Scholar] [CrossRef]
  98. Rangel, N.; Rondon-Lagos, M.; Annaratone, L.; Osella-Abate, S.; Metovic, J.; Mano, M.P.; Bertero, L.; Cassoni, P.; Sapino, A.; Castellano, I. The role of the AR/ER ratio in ER-positive breast cancer patients. Endocr. Relat. Cancer 2018, 25, 163–172. [Google Scholar] [CrossRef] [Green Version]
  99. Cao, L.; Xiang, G.; Liu, F.; Xu, C.; Liu, J.; Meng, Q.; Lyu, S.; Wang, S.; Niu, Y. A high AR:ERalpha or PDEF:ERalpha ratio predicts a sub-optimal response to tamoxifen therapy in ERalpha-positive breast cancer. Cancer Chemother. Pharmacol. 2019, 84, 609–620. [Google Scholar] [CrossRef]
  100. Tangen, I.L.; Onyango, T.B.; Kopperud, R.; Berg, A.; Halle, M.K.; Oyan, A.M.; Werner, H.M.; Trovik, J.; Kalland, K.H.; Salvesen, H.B.; et al. Androgen receptor as potential therapeutic target in metastatic endometrial cancer. Oncotarget 2016, 7, 49289–49298. [Google Scholar] [CrossRef] [Green Version]
  101. Rajarajan, S.; Korlimarla, A.; Alexander, A.; Anupama, C.E.; Ramesh, R.; Srinath, B.S.; Sridhar, T.S.; Prabhu, J.S. Pre-Menopausal Women with Breast Cancers Having High AR/ER Ratios in the Context of Higher Circulating Testosterone Tend to Have Poorer Outcomes. Front. Endocrinol. 2021, 12, 679756. [Google Scholar] [CrossRef]
  102. Lehmann, B.D.; Bauer, J.A.; Chen, X.; Sanders, M.E.; Chakravarthy, A.B.; Shyr, Y.; Pietenpol, J.A. Identification of human triple-negative breast cancer subtypes and preclinical models for selection of targeted therapies. J. Clin. Investig. 2011, 121, 2750–2767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Masuda, H.; Baggerly, K.A.; Wang, Y.; Zhang, Y.; Gonzalez-Angulo, A.M.; Meric-Bernstam, F.; Valero, V.; Lehmann, B.D.; Pietenpol, J.A.; Hortobagyi, G.N.; et al. Differential Response to Neoadjuvant Chemotherapy Among 7 Triple-Negative Breast Cancer Molecular Subtypes. Clin. Cancer Res. 2013, 19, 5533–5540. [Google Scholar] [CrossRef] [Green Version]
  104. Shufelt, C.L.; Braunstein, G.D. Safety of testosterone use in women. Maturitas 2009, 63, 63–66. [Google Scholar] [CrossRef] [PubMed]
  105. Katsambas, A.D.; Dessinioti, C. Hormonal therapy for acne: Why not as first line therapy? facts and controversies. Clin. Dermatol. 2010, 28, 17–23. [Google Scholar] [CrossRef] [PubMed]
  106. Sakamoto, K.M.; Kim, K.B.; Kumagai, A.; Mercurio, F.; Crews, C.M.; Deshaies, R.J. Protacs: Chimeric molecules that target proteins to the Skp1-Cullin-F box complex for ubiquitination and degradation. Proc. Natl. Acad. Sci. USA 2001, 98, 8554–8559. [Google Scholar] [CrossRef] [Green Version]
  107. Qi, S.M.; Dong, J.; Xu, Z.Y.; Cheng, X.D.; Zhang, W.D.; Qin, J.J. PROTAC: An Effective Targeted Protein Degradation Strategy for Cancer Therapy. Front. Pharmacol. 2021, 12, 692574. [Google Scholar] [CrossRef]
  108. Salami, J.; Alabi, S.; Willard, R.R.; Vitale, N.J.; Wang, J.; Dong, H.Q.; Jin, M.Z.; McDonnell, D.P.; Crew, A.P.; Neklesa, T.K.; et al. Androgen receptor degradation by the proteolysis-targeting chimera ARCC-4 outperforms enzalutamide in cellular models of prostate cancer drug resistance. Commun. Biol. 2018, 1, 100. [Google Scholar] [CrossRef]
  109. Zhao, L.; Han, X.; Lu, J.; McEachern, D.; Wang, S. A highly potent PROTAC androgen receptor (AR) degrader ARD-61 effectively inhibits AR-positive breast cancer cell growth in vitro and tumor growth in vivo. Neoplasia 2020, 22, 522–532. [Google Scholar] [CrossRef]
  110. Xiang, W.G.; Zhao, L.J.; Han, X.; Qin, C.; Miao, B.K.Y.; McEachern, D.; Wang, Y.; Metwally, H.; Kirchhoff, P.D.; Wang, L.; et al. Discovery of ARD-2585 as an Exceptionally Potent and Orally Active PROTAC Degrader of Androgen Receptor for the Treatment of Advanced Prostate Cancer. J. Med. Chem. 2021, 64, 13487–13509. [Google Scholar] [CrossRef]
  111. Chen, L.R.; Han, L.Q.; Mao, S.J.; Xu, P.; Xu, X.X.; Zhao, R.B.; Wu, Z.H.; Zhong, K.; Yu, G.L.; Wang, X.L. Discovery of A031 as effective proteolysis targeting chimera (PROTAC) androgen receptor (AR) degrader for the treatment of prostate cancer. Eur J. Med. Chem. 2021, 216, 113307. [Google Scholar] [CrossRef] [PubMed]
  112. Masoodi, K.Z.; Xu, Y.; Dar, J.A.; Eisermann, K.; Pascal, L.E.; Parrinello, E.; Ai, J.; Johnston, P.A.; Nelson, J.B.; Wipf, P.; et al. Inhibition of Androgen Receptor Nuclear Localization and Castration-Resistant Prostate Tumor Growth by Pyrroloimidazole-based Small Molecules. Mol. Cancer Ther. 2017, 16, 2120–2129. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The proposed mechanism for the suppression of AR on ERα as reported by Hickey et al. [19]. Under normal circumstances, active ERα interacts with its coactivators, such as p300/SRC-3, and then bind to estrogen response element (ERE) to induce the transcription. Active AR may act as an ERα suppressor by competitive binding to ERE or by interacting with ERα coactivators to hijack ERα from ERE to ARE, thereby preventing ERα from transcribing its target genes.
Figure 1. The proposed mechanism for the suppression of AR on ERα as reported by Hickey et al. [19]. Under normal circumstances, active ERα interacts with its coactivators, such as p300/SRC-3, and then bind to estrogen response element (ERE) to induce the transcription. Active AR may act as an ERα suppressor by competitive binding to ERE or by interacting with ERα coactivators to hijack ERα from ERE to ARE, thereby preventing ERα from transcribing its target genes.
Ijms 23 15342 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

You, C.-P.; Tsoi, H.; Man, E.P.S.; Leung, M.-H.; Khoo, U.-S. Modulating the Activity of Androgen Receptor for Treating Breast Cancer. Int. J. Mol. Sci. 2022, 23, 15342. https://doi.org/10.3390/ijms232315342

AMA Style

You C-P, Tsoi H, Man EPS, Leung M-H, Khoo U-S. Modulating the Activity of Androgen Receptor for Treating Breast Cancer. International Journal of Molecular Sciences. 2022; 23(23):15342. https://doi.org/10.3390/ijms232315342

Chicago/Turabian Style

You, Chan-Ping, Ho Tsoi, Ellen P. S. Man, Man-Hong Leung, and Ui-Soon Khoo. 2022. "Modulating the Activity of Androgen Receptor for Treating Breast Cancer" International Journal of Molecular Sciences 23, no. 23: 15342. https://doi.org/10.3390/ijms232315342

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop