Next Article in Journal
Synergy Assessment of Four Antimicrobial Bioactive Compounds for the Combinational Treatment of Bacterial Pathogens
Previous Article in Journal
Effect of Leflunomide–Metal Complexes on ROS, TNF, and Brain Indolamines in Comparison with Anti-Depressants as Adjunct Therapy in Rheumatoid Arthritic Model
Previous Article in Special Issue
Androstenedione and Follicle-Stimulating Hormone Concentration Predict the Progression of Frailty Syndrome at One Year Follow-Up in Patients with Localized Breast Cancer Treated with Aromatase Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Beyond Prostate Cancer: An Androgen Receptor Splice Variant Expression in Multiple Malignancies, Non-Cancer Pathologies, and Development

by
Kimberley D. Katleba
1,2,
Paramita M. Ghosh
1,3,4 and
Maria Mudryj
1,2,*
1
Veterans Affairs-Northern California Health Care System, 10535 Hospital Way, Mather, CA 95655, USA
2
Department of Medical Microbiology and Immunology, 1 Shields Avenue, UC Davis, Davis, CA 95616, USA
3
Department of Urologic Surgery, 4860 Y Street, UC Davis, Sacramento, CA 95718, USA
4
Department of Biochemistry and Molecular Medicine, 1 Shields Avenue, UC Davis, Davis, CA 95616, USA
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(8), 2215; https://doi.org/10.3390/biomedicines11082215
Submission received: 9 June 2023 / Revised: 19 July 2023 / Accepted: 21 July 2023 / Published: 7 August 2023
(This article belongs to the Special Issue Molecular Mechanisms of Steroid Hormone Action)

Abstract

:
Multiple studies have demonstrated the importance of androgen receptor (AR) splice variants (SVs) in the progression of prostate cancer to the castration-resistant phenotype and their utility as a diagnostic. However, studies on AR expression in non-prostatic malignancies uncovered that AR-SVs are expressed in glioblastoma, breast, salivary, bladder, kidney, and liver cancers, where they have diverse roles in tumorigenesis. AR-SVs also have roles in non-cancer pathologies. In granulosa cells from women with polycystic ovarian syndrome, unique AR-SVs lead to an increase in androgen production. In patients with nonobstructive azoospermia, testicular Sertoli cells exhibit differential expression of AR-SVs, which is associated with impaired spermatogenesis. Moreover, AR-SVs have been identified in normal cells, including blood mononuclear cells, neuronal lipid rafts, and the placenta. The detection and characterization of AR-SVs in mammalian and non-mammalian species argue that AR-SV expression is evolutionarily conserved and that AR-SV-dependent signaling is a fundamental regulatory feature in multiple cellular contexts. These discoveries argue that alternative splicing of the AR transcript is a commonly used mechanism that leads to an expansion in the repertoire of signaling molecules needed in certain tissues. Various malignancies appropriate this mechanism of alternative AR splicing to acquire a proliferative and survival advantage.

1. Introduction

The androgen receptor (AR) has a dominant role in mammalian physiology from early development, where it is critical in establishing the male phenotype, to post-natal development and adult maintenance of multiple tissues [1,2,3,4]. While expressed in most cells, this ligand-regulated transcription factor is dependent on androgen (e.g., testosterone or dihydrotestosterone (DHT)) binding for activation [5,6]. Testosterone is secreted by testicular interstitial (Leydig) cells and, at much lower levels, by the adrenal cortex and ovaries [2,7,8]. In addition, testosterone can be converted by 5α reductase to DHT, which binds to the AR with higher affinity [2,7]. There is a ~15-fold higher plasma concentration of testosterone in males than females, resulting in greater AR signaling [9].
Since the AR is a member of the 48-member steroid receptor superfamily, it has the same basic structure as other family members, consisting of an N-terminal activation (NTD), DNA binding (DBD), and ligand binding domains (LBD), where a hinge region lies between the DBD and LBD [3,4,10]. The AR gene maps to the X chromosome and consists of eight exons [11,12]. The NTD is encoded by exon 1, exons 2 and 3 code for the DBD, and the C-terminal LBD is encoded by the terminal four exons. Exon 4 encodes the hinge region, which bridges the DBD and LBD and contains a nuclear localization signal [5]. Prior to ligand binding, the AR is inactive and complexed in the cytoplasm with heat shock proteins (HSPs) [13]. Ligand binding results in a conformation change, release from cytoplasmic HSPs, exposure of the nuclear localization signal, phosphorylation, and ultimately translocation into the nucleus. The dimerized AR DBD tethers the receptor to AR regulatory elements (AREs) on the target DNA to serve as a platform for the assembly of regulatory molecules to drive or repress transcription [6,14,15]. While the LBD and DBD X-ray structures have been solved separately, more recently Yu and collaborators defined the three-dimensional structure of an androgen-occupied AR bound to DNA using cryo-electron microscopy (cryo-EM) [14]. This study confirms that AR dimerizes in a head-to-head and tail-to-tail fashion, shows that the LBD, DBD, and NTDs all have a dimerization interface, and shows that the LBD and the NTD interact. Cryo-EM shows that the NTDs wrap around the LBD and have a major role in recruiting and interacting with SRC-3 and p300 co-regulators (Figure 1).
While the AR has a critical role in male development and physiology, it is a pivotal regulatory molecule in prostate cancer (PCa). Its central function in the initiation and progression of this malignancy was proposed by Huggins and Hodge over 80 years ago [16]. Their visionary studies established that PCa is a hormonally dependent cancer and that limiting AR activity by reducing testosterone levels is efficacious in limiting disease. Hence, for the last eight decades, androgen deprivation therapy (ADT) has been the chief strategy for treating advanced PCa [17,18,19,20,21]. Unfortunately, while initially effective, this therapeutic intervention ultimately fails due to the development of castration-resistant cells [22,23,24]. Multiple mechanisms lead to castration-resistant prostate cancer (CRPC), where the cells remain reliant on AR, but AR signaling is no longer dependent on testosterone [25].
The emergence of AR splice variants (AR-SVs), which retain the NTD and DBD but no longer express the LBD and are therefore refractory to therapies that target this region, is greatly associated with the acquisition of castration resistance. Some AR-SVs retain the ability to translocate into the nucleus, bind to DNA, and regulate gene transcription constitutively [26,27,28,29,30,31].
Below, we summarize our comprehensive review of studies on AR-SV expression and function in multiple non-prostatic malignancies, non-cancer pathologies, and normal tissues in humans and non-human species. This comparative analysis underscores the wide prevalence of AR-SVs in differing cells, where these regulatory molecules make unique contributions to defining cellular physiology in various pathologies and normal processes.

2. AR Splice Variants Expressed in Prostate Cancer

2.1. Classification and Signaling

Historically, lower molecular weight (LMW) AR variants have been reported in studies on AR in androgen insensitivity syndrome and in malignancies, but their role in PCa was established only about a decade ago [32,33,34,35,36]. While several mechanisms can give rise to these AR forms, including mutations leading to premature termination signals and proteolytic processing that removes the LBD, splice variants (SVs) are the most prominent and best studied [26]. Multiple AR-SVs have been uncovered, and most encode proteins that share common features: they retain the NTD and DBD, are missing the LBD, but acquire a variant-specific short C-terminal sequence encoded by aberrant splicing to a cryptic exon within an intronic region [37]. AR-SVs can be broadly classified into three groups: (1) constitutively active, (2) conditionally active, and (3) inactive [38]. The most extensively characterized variant 7 (AR-V7, aka ARV3) retains the NTD and DBD, while the hinge and LBD are replaced by a 15-amino acid sequence that is derived from intron 3 sequences. Many splice variants are not transcriptionally active, but AR-V7 is representative of constitutively active variants, which can translocate into the nucleus and drive gene expression [5].
In the context of PCa, full-length-AR (FL-AR) and AR-V7 DNA binding site specificities, either homo or heterodimers, have been explored in several studies, but the results are inconsistent; some find that FL-AR and AR-V7 heterodimer and FL-AR homodimer binding sites exhibit significant overlap, while others report a unique AR-V7 binding profile [37,39,40,41] (Figure 2). AR-V7 binding to unique sites is enabled by interactions with co-regulators, including HOXV13 and ZFX, and AR-V7 binding favors open chromatin regions proximal to the transcriptional start site [40,41]. Distinct interactions with co-regulators would be expected since removal of the LBD would significantly alter the three-dimensional structure of a dimerized molecule, either exposing or masking regions capable of interaction with coregulators. Therefore, the FL-AR and AR-V7 transcriptomes have some common targets, but most are unique [33,35,40,42,43]. Multiple studies have established that, in PCa, AR-V7 expression is elevated in CRPC. This led to the development of diagnostics for the detection of AR-V7 in circulating tumor cells [30,44,45,46].
In PCa, expression of AR-SVs has two well-described etiologies: genomic rearrangements of the AR locus and aberrant RNA splicing. Genomic rearrangements associated with AR-SVs have been described in PCa in vitro models and in human CRPC lesions [32,47,48]. The alterations in patient samples are patient-specific, not recurrent, and most specimens harboring genomic alterations express AR-SVs, but there are exceptions [49,50]. AR-SVs are also expressed in cells that have no AR genomic alterations; therefore, a clear correlation between genomic alterations and AR-SVs is not firmly established.
Aberrant RNA splicing is sensitive to FL-AR signaling and therefore can be rapidly initiated in the absence of androgens but reversed when androgen dependent signaling is reinstated, arguing that this process is an adaptation rather than a clonal selection of AR-SV-expressing cells and that androgen deprivation drives AR-SV-expression [33,51,52]. Notably, AR-SV expression is associated with AR gene amplification or transcript elevation; hence, high levels of AR expression may enable AR-SV detection [30,38]. However, other mechanisms may underlie the expression of AR-SVs [53]. The splicing factor hnRNPA1 is upregulated in PCa cell lines harboring AR-SVs, and its recruitment of AR-SV mRNA splice junctions but not FL-AR mRNA splice sites is increased [52]. Splicing factors U2AF65 and ASF/SF2 recognize binding sites near AR exon 3 and facilitate the recruitment of the RNA splicesome to the AR-V7 3′ splice site [54]. Studies of HSP90 have implicated this molecular chaperone in AR-SV splicing, and HSP90 inhibition disrupts this process, resulting in decreased AR-V7 production [55]. The noncoding RNA PCGEM1 and splicing factors also contribute to mRNA splicing, and androgen deprivation induces PCGEM1 redistribution into nuclear speckles and facilitates the interaction of PCGEM1 with U2AF65 [56]. Thus, mechanisms governing splicing are likely cell-context-specific and subject to multiple layers of regulation.
Multiple animal models of cancer have been developed, but surprisingly, there is little information on AR-SVs in these models. The best studies are PCa models developed by genetically engineering mice (reviewed [57]). Many models of aggressive cancer are representative of neuroendocrine malignancies and do not express AR. However, the importance of AR-SVs has been directly addressed by modeling AR-V7 and AR-V567es in transgenic mice where prostate epithelial cell-specific AR-SV forced expression was driven by a modified probasin promoter [58,59]. AR-V7 overexpression promoted the development of prostatic interstitial neoplasia-like lesions. Notably, AR-V7 altered the expression of genes associated with prostate stem cells/progenitor cells and microRNAs deregulated in human cancers. Additionally, the upregulation of TGFβ signaling-associated genes suggested that AR-V7 overexpression may promote epithelial-mesenchymal transition. Prostate-specific expression of AR-V567es drove early hyperplasia and ultimately the development of invasive adenocarcinoma. Transcriptome studies identified enriched expression of inflammatory-related cytokines, transcription factors, and tumorigenesis-associated transcripts. Together, these studies confirm that AR-SV expression reprograms cellular transcription, elevating the expression of genes that promote tumor initiation and progression (Figure 3).

2.2. Developmental Therapy Targeting Androgen Receptor Splice Variants

Strides have been made in the development of agents that target the AR-NTD with the goal of inhibiting NTDs containing AR isoforms. Unlike the LBD, the intrinsically disordered AR-NTD makes structure-based drug design difficult, but inhibitor development is underway. In a recent review, Ji and co-authors surveyed the status of NTD-targeting agent development, which can broadly be assigned to four groups [60]:
EPI-001 and the analogues are bisphenol A diglyceride derivatives initially isolated from murine sponge. The original isolated compound was reported to target the AR-NTD [61,62]. While its mechanistic basis has been explored, EPI-001 has a secondary mechanism of action where it was found to selectively modulate peroxisome proliferator-activated receptor-gamma [63]. The most recent derivative, EPI-7386, has a favorable safety, absorption, distribution, and metabolism profile and is able to overcome resistance to current anti-androgen therapies. It therefore exhibits clinical potential and has entered a phase 1 clinical trial (NCT04421222).
Bicalutamide-derived selective AR degraders include several compounds that can directly bind to the AR-NTD of FL-AR and AR-SVs to block interactions with co-activators [64]. The most recently developed degrader demonstrates in vivo efficacy in reducing the growth of castrate-resistant prostate tumors [65].
Agents with dual targeting of STAT3 and AR inhibitory activity have been identified. These compounds interact with the AR-NTD to inhibit recruitment of the AR co-activator SRC-1, resulting in a loss of activity, blocking nuclear translocation of FL-AR and AR-SVs, and causing AR degradation by way of the ubiquitin-mediated pathway [66].
A biphasic antibody molecule that consists of a single-chain variable antibody fragment that recognizes DNA and a single-chain variable fragment that recognizes an epitope in the AR-NTD can accumulate in the nucleus and block signaling by the FL-AR and AR-SVs. This molecule can inhibit the proliferation of AR-dependent cells in a ligand-independent manner [67].

3. Sexual Dimorphism of Multiple Cancers

AR signaling has been suggested as a contributing factor in other cancers [68,69]. Varying degrees of male preference are observed in many malignancies, but for several, including the pharynx, larynx, esophagus, and bladder, the incidence in males is over three times greater than that in females (Figure 4). This is accompanied by higher cancer-related deaths in males than females [70]. Since the male and female hormonal milieus differ, where males have higher levels of testosterone while females have greater estrogen levels, the contribution of hormonal signaling, where androgens are tumor promoting and estrogens are protective, has been suggested [71].
Cell culture and animal models reiterate the contributions of androgens and/or AR to the development of cancer. However, clinical studies on the efficacy of AR-targeting therapeutics have been disappointing, showing little measurable benefit. Multiple explanations can reconcile the divergent preclinical and clinical results, including the expression of constitutively active AR-SVs, which, as in CRPC, could render cells refractory to current AR targeting therapeutics.

4. Androgen Receptor Splice Variants in Human Malignancies

Below, we review reports of AR-SV expression in non-prostatic malignancies. All the malignancies, with the exception of breast cancer (BrCa), exhibit a sex disparity where males are more likely to develop disease than females. However, counterintuitively, AR-SVs were detected in tumors from males and females, and males did not have an increase in incidence or levels of AR-SVs. In BrCa, a cancer that overwhelmingly occurs in females, AR-SVs are detected in most high-grade malignancies. It is notable that in the context of BrCa, the tumors that most frequently express AR-SVs are Triple Negative BrCa (TNBC), particularly tumors with apocrine features. Moreover, salivary duct carcinomas, the only salivary gland tumor that displays a sex disparity, also exhibit apocrine features and express AR-SVs. While the transcriptomes regulated by distinct AR-SVs have not been investigated in all contexts, the prevailing pattern suggests that the AR-SV-regulated transcriptome is different from the FL-AR-driven transcriptional program. Moreover, there is little, if any, similarity between the investigated AR-V7-regulated transcriptions in different cellular contexts. Lastly, another emerging common theme is that while the level of an SV is low in comparison to the amount of total AR, depletion of the constitutively active AR-SV has a disproportionate effect on the cells’ physiological properties, suggesting that its role is independent of other AR-SVs and FL-AR and needed for optimal cell survival and growth (Table 1).

4.1. Breast Cancer

Breast cancers (BrCa) are highly reliant on estrogens and progesterone, whose activity is mediated by their respective receptors [72,73,74]. Hence, in BrCa that expresses these receptors, blocking estrogen receptor (ER) signaling is the first-line treatment strategy. Both the ER and progesterone receptor (PR) are nuclear hormone receptors with structures similar to the AR. The ER is encoded by two distinct genes, ERα and ERβ. These two subunits form hetero or homodimers, which, when bound to estrogens, mediate estrogen signaling. The PR is encoded by a single gene, but two separate promoters are utilized to generate isoforms A and B. AR signaling has been implicated in BrCa as well, but its role in tumorigenesis is less well defined, and epidemiological studies yielded mixed results, even though immunohistochemical studies found that AR is expressed in 31% to 85% of human breast tumors [75,76,77]. Some have identified a significant increase in BrCa risk for women with elevated testosterone and low progesterone levels [78,79]. In other studies, low levels of the androgens dehydroepiandrosterone (DHEA) or dehydroepiandrosterone sulfate (DHEAS) are associated with increased BrCa risk in premenopausal women [80,81,82], while another study did not show a clear association between androgens and BrCa risk [83].
AR as a prognostic marker of disease progression is complicated and nuanced and appears to depend on whether tumors express ER and PR. In ERα-positive malignancies, AR may be a tumor suppressor, while in ERα-negative tumors, AR may be a tumor promoter [84]. AR-regulated genes positively correlate with node invasiveness and independently predict the likelihood of axillary metastases [85,86,87], but AR-positive tumors were associated with better disease-free survival rates [77]. The mechanistic basis of AR-dependent tumor suppression in ERα-positive malignancies may be due to crosstalk between ERα and AR, where the interaction of the two can inhibit the activity of both [88]. AR can compete with ER by binding to ERα response elements or by interacting with co-activators [89]. The AR can also upregulate tumor suppressor genes to directly repress cell growth [90]. Clinical studies that included testosterone for treatment of ERα-positive malignancies either alone or in combination with aromatase inhibitors show that testosterone treatment has clinical benefit [91,92,93].
In contrast, in ERα-negative malignancies, including triple negative (ER-, PR-, and HER2-negative) BrCa (TNBC), AR has oncogenic properties. Blocking AR signaling in ERα-negative cell models suppressed cell growth [94]. The combination of the AR antagonist enzalutamide and HER2-targeting trastuzamb synergizes to inhibit proliferation [95]. However, AR-dependent inhibition of proliferation and induction of cell death may be due to inhibition of WNT/β-catenin and EGFR signaling pathways [96,97]. Significantly, AR has been shown to increase HER2 expression and subsequent activation of MAPK, the WNT/β pathway, and SRC, activating this oncogenic signaling cascade [98,99,100]. Therefore, in ERα-negative malignancies, AR is associated with a poor prognosis [101,102,103], suggesting that AR inactivation therapy may be beneficial [104]. Given these results, AR suppressive agents have been tested either alone or in combination with other drugs in clinical trials in patients with ERα-negative BrCa and have shown efficacy in some patients [105,106,107], but these malignancies are aggressive, and patients who respond quickly relapse [108].
Several studies reported the presence of AR-SVs in BrCa [109,110,111,112,113]. Initial studies reported that five AR-SVs originally identified in PCa (AR-V1, AR-V3, AR-V7, AR-V9, and AR45) are expressed at varying levels in six studied BrCa cell lines. Novel AR-SVs encoding a ninth exon were identified as well [109]. A subsequent study examined AR-SV expression in primary breast tumors. Transcript analysis detected FL-AR in 98.1% and AR-V7 in 53.7% of tumors. AR-V7 protein expression is present in BrCa tumors and in BrCa-derived cell lines. In contrast with what is observed in PCa, FL-AR and AR-SVs are coordinately regulated in BrCa [110]. ERα-negative malignancies have higher levels of FL-AR and AR-V7 than ERα-positive tumors. In a cell culture model, AR-V7 is constitutively active, and activity is not altered by AR antagonists. AR-V7 overexpression in the presence and absence of androgens alters transcription of 64 and 47 transcripts, but only 19% and 11% are upregulated, respectively [110]. 28 transcripts are common to both, and of these, 27 are repressed. Alteration of AR-V7-dependent transcription is distinct from AR-FL, arguing that AR-V7 in this cellular context primarily represses gene expression. Over-expression of AR-V7 modulates transcripts involved in immune function and signaling as well as cell movement. This transcriptome is distinct from what has been observed with AR-V7 overexpression in PCa LNCaP cells, where overexpression primarily alters cell cycle-associated transcripts. These differences are most likely due to distinct cohorts of co-regulators that collaborate with AR-V7 in regulating transcription. Functionally, AR-V7 depletion reduced cell viability, and the reduction was comparable to that of FL-AR, even though AR-V7 is far less abundant. These results further indicate that AR-V7 has distinct pro-proliferative properties in ARα-negative BrCa [110]. Further studies examined the frequency of AR-V7 in BrCa and its correlation with clinical and histopathological features of primary and metastatic malignancies. The majority of AR-V7-positive cases are TNBC, and 71% of these malignancies have apocrine features. In contrast to PCa, AR-V7 is present in primary and metastatic lesions. Given these results, apocrine morphology may serve as a useful screen for the stratification of tumors for the presence of AR-V7 and consideration of AR antagonist therapy [113].

4.2. Salivary Duct Carcinoma

Salivary duct carcinoma is a rare cancer of the salivary glands, accounting for 2% of salivary gland tumors. This aggressive malignancy is frequently fatal, and for advanced, unresectable lesions, therapeutic options are limited. Most salivary malignancies do not display a sex disparity, but salivary duct carcinomas are an exception and are more common in males than females [114,115]. Immunohistochemical studies revealed that histologically, these tumors have an apocrine phenotype and share morphological features with invasive ductal carcinoma of the breast [116]. Similar to breast tumors, salivary duct carcinomas stain positive for AR, which is detected in up to 98% of tumors [117,118,119,120]. Notably, several case studies reported clinical benefit from ADT, suggesting that this is a potential therapeutic strategy for patients with recurring lesions [121,122,123,124]. A phase II study of combined androgen blockade (CAB), using leuprorelin acetate and bicalutamide, reported an overall response rate of ~42% of AR-positive malignancies, suggesting that this treatment has promise for recurrent, metastatic, or unresectable locally advanced carcinomas [125]. A larger study of HER2-/AR-positive malignancies found that patients treated with HER2 targeting and ADT therapies showed longer survival [126]. A study on the efficacy of enzalutamide for the treatment of AR-positive malignancies found that while some patients partly responded to treatment, overall, this treatment had limited success [127]. These studies found that while some patients respond to androgen blockade, others that express nuclear AR are refractory to this treatment, highlighting the need for biomarkers that could predict response to androgen deprivation therapies.
AR-SV expression was assessed in a cohort of male and female patients of varying ages [128]. The expression of three different AR-SVs, including AR-V7, was assessed by RT-PCR. AR-V7 is the most frequently detected variant and is present in 37% of tumors. AR-SV expression was confirmed by Western immunoblot analysis. This study also identified AR gene alterations in male and female patients. In vitro studies using salivary duct carcinoma-derived cells showed that depletion of AR using siRNA inhibits cell growth, but cell proliferation is unaltered in steroid-depleted media. The lack of androgen effects was reiterated in an in vivo assay of tumor growth [114]. A study that focused on AR-V7 found that 87% of tumors are positive for AR mRNA, and protein analysis corroborated AR expression. 70% of the AR-positive tumors express AR-V7 mRNA, and AR-V7 protein is detectable in three of five high expressing AR-V7 mRNA samples but is not detected in tumors with lower AR-V7 mRNA levels. In a similar study that evaluated several pathogenic genetic alterations, AR was expressed in all cases, and 50% expressed detectable levels of AR-V7, arguing that AR-V7 is a common feature of these malignancies [129]. Since PCa AR-V7 expression is associated with tumors that are refractory to AR-targeting therapies, AR-V7 expression in salivary duct carcinomas may serve a similar function; hence, detection of AR-SV in salivary duct malignancies may allow for better stratification of patients that would benefit from AR-targeting treatments.

4.3. Glioblastoma Multiforme

Glioblastoma multiforme (GBM) exhibits a sex disparity where males are 1.59 times more likely than females to develop this malignancy, and females exhibit a survival advantage [130,131]. Females also exhibit a better response to radiation and chemotherapy [132]. AR protein levels are elevated in 93% of GBMs derived from males and females and from GBM-derived cell lines when compared to normal brain tissue [133,134]. AR expression increases with tumor grade, regardless of sex. Interestingly, 5α reductase 2, an enzyme that converts testosterone to DHT, is elevated in GBMs when compared to normal brain tissue [135]. Glioma cells implanted into nude male and female mice exhibit differential tumor growth; in male animals, tumors are significantly larger and have shorter latency periods, and AR is preferentially detected in tumors propagated in male animals [136]. GBM-derived cell lines expressing AR are sensitive to AR modulation. siRNA-mediated depletion of AR significantly reduces cell proliferation and promotes apoptosis [137]. Likewise, AR antagonist treatments result in decreased cell proliferation, migration, and invasion [138,139,140]. The efficacy of an antagonist was demonstrated in an animal model where treatment with enzalutamide significantly reduced tumor cell growth [140]. Jointly, these studies strongly suggest that AR signaling is pro-tumorigenic in GBM and that AR antagonists curtail tumor cell growth in vitro and in vivo.
AR gene amplification was detected in 27% of GBMs from men and 38% of tumors from women, and further analysis shows that AR mRNA is overexpressed in 93% of tumor samples [140]. This detailed analysis identified a positive correlation between AR expression and age but no correlation between AR levels and survival, gender, race, or ethnicity. AR protein analysis of primary tumor samples and tumor-derived cell lines verified AR expression but also found that 30% of AR-expressing tumors and cells also expressed AR-V7, where the variant is always co-expressed with FL-AR. Treatment of cells expressing both FL-AR and AR-V7 found that they are sensitive to enzalutamide; therefore, in this context, the presence of AR-V7 does not render the cells refractory to this intervention. Thus far, it is unknown if modulating AR-V7 expression in these cells alters the transcriptional output or sensitivity to AR-targeting therapeutics.

4.4. Renal Cancer

Epidemiological studies found that males have a two-fold higher risk of developing kidney cancer than females, and this disparity is unaltered by age, year surveyed, or geographic location [141]. While multiple environmental and genetic factors influence the development of renal cell carcinoma (RCC), accumulating evidence supports a role for the androgen signaling axis [142]. Multiple studies report AR expression in RCC [143,144,145,146], but the correlations between AR protein expression and clinical outcomes are conflicting [145,147]. Cell culture and animal models demonstrated the role of AR signaling in RCC. Normal kidney cells transfected with AR-encoding sequences develop larger and larger colonies when treated with carcinogens and promote cell migration and invasion [141]. AR induces vascular endothelial cell proliferation, an effect that is mediated through the PI3K/AKT signaling pathway and results in increased CXCL5 expression [145]. Additional studies reported that the non-coding RNAs HOTAIR and AR are components of a feedback loop that drives transcription of the GLI2 transcription factor, leading to increased expression of the downstream genes VEGFA and PDGFA [147]. Studies of AR-expressing RCC cell lines show that DHT treatment promotes proliferation through STAT5 signaling, and AR inhibition demonstrated antitumor effects [148]. Animal studies utilizing AR-positive RCC tumor cells show that treatment with enzalutamide or limiting DHT production with abiraterone acetate significantly reduces tumor growth [149]. Thus far, there have been no completed clinical trials on the efficacy of hormone deprivation therapy in limiting RCC.
The two most common RCC subtypes are clear-cell RCC (ccRCC) and papillary RCC (pRCC). ccRCC accounts for 75% of all RCCs, while pRCC accounts for approximately 10%. The tumors differ in vascularization; ccRCCs are highly vascularized, while pRCCs are hypovascularized when compared to the surrounding parenchyma [150,151]. ccRCC patients have a worse prognosis than pRCC patients, and approximately 15% of them develop lung, liver, bone, or lymph node metastases [151]. AR splice variant expression (AR-V1, AR-V3, AR-V4, AR-V7, and AR-V12) was investigated using primary ccRCC and pRCC tumors, focusing on AR splice variants that are constitutively active [152]. ccRCC and normal kidney tissues express FL-AR and AR splice variants, but expression is more frequent in pRCC. This result agrees with a previous analysis showing higher AR expression in pRCC than in ccRCC [153]. AR-V7 is the dominant variant expressed, while AR-V12 is not expressed in the analyzed tumors. FL-AR and AR-SV expression are higher in high-grade malignancies (pT4), but there are no significant correlations between patient sex and AR-SV expression. Further inquiries investigated the relationship between relaxin 2 (RLN2), a protein that influences the AR signaling pathway and, in the kidney, is a renoprotective factor whose activity is sex-specific, and AR-SVs [154,155,156]. This study found a significant correlation between FL-AR, AR-V1, and AR-V4 transcript levels and RLN2 expression, a potential modulator of AR activity. More in vitro, in vivo, and genomic studies are needed to define the functional and mechanistic roles of FL-AR and AR-SVs in these malignancies.

4.5. Bladder Cancers

Bladder cancer (BlCa) displays a striking sex disparity where males are 3–4 times more likely to develop the disease than females [157]. Numerous studies have reported that the AR has a significant role in BlCa development. Immunohistochemical (IHC) analysis found that AR expression is higher in tumor tissue from males than from females [158]. Moreover, AR staining is more intense and localizes to the nucleus in tumors from males than females [159]. Animal models have confirmed the hypothesis that androgens contribute to bladder carcinogenesis. Supplementation of female rats with testosterone promotes N-butyl-N-(4-hydroxybutyl) nitrosamine (BBN) and induces bladder tumors [160], and urothelium-targeted AR knockout mice have a reduced incidence of BBN-induced tumors [161]. Men with BlCa treated with 5α-reductase inhibitors exhibit a lower risk of BlCa-related death than men not treated with these agents [162]. These studies indicate a significant role for the AR in BlCa. Despite this, two BlCa clinical trials that used the AR inhibitor enzalutamide to target AR were terminated early due to a lack of efficacy.
Most bladder cancer cells express AR-SV, and in some contexts, this is the major isoform expressed [163]. However, the subcellular distribution of AR varies, where some cells have abundant nuclear AR expression while in others the AR is confined to the cytoplasm, and subcellular localization is not dependent on whether the cells were obtained from malignancies in males or females. This study detected AR-V7 expression in ~50% of cells surveyed but also identified a new constitutively active variant, AR-V19, which, like AR-V7, is missing the LBD, contains C-terminal sequences derived from intron 3, localizes to the nucleus, and is constitutively active. Depletion of all AR forms in cells with nuclear AR expression reduces cell viability and increases apoptosis, but AR-V19-specific depletion also effectively reduces cell viability in cells expressing this isoform. Overexpression of FL-AR does not alter AR-SV expression, and conversely, overexpression of V19 does not reduce FL-AR; therefore, the reciprocal negative regulation observed in PCa is not apparent in BlCa. Transcript analysis following depletion of FL-AR, AR-V19, or total AR identified distinct but partly overlapping transcriptomes [164]. mTOR is decreased by deletion of Fl-AR, AR-V19, or all AR isoforms, and this is coupled with altered mTOR downstream signaling. A decrease in mitochondrial components and Increased expression of HIF1α- and HIF1α-regulated transcripts coupled with a significant decrease in oxygen consumption rate on AR depletion argue that an important role of AR signaling is enhanced mitochondrial activity to drive energy production. Taken together, AR signaling in BlCa is appropriated to increase metabolic processes to enhance BlCa proliferation.

4.6. Liver Cancer

The AR is implicated in the development and pathogenesis of hepatocellular carcinomas (HCC) [165,166,167]. Mouse HCC models reiterate the contribution of AR signaling, where conditional liver-specific AR knockout decreases the incidence of carcinogen-induced HCC [168]. AR also has a role in Hepatitis B virus (HBV)-induced HCC, an effect that can be attributed to functional AREs in the HBV genome, whereby AR is able to stimulate viral gene transcription, promoting hepatocarcinogenesis [169]. This result is consistent with the observed large differences in male and female incidence of HCC in HBV endemic regions [170,171]. AR protein levels are higher in HCC than in adjacent tissues, and higher AR levels correlate with increased tumor recurrence and decreased overall survival [168,172]. More recent studies find that cytoplasmic AR levels do not differ between tumors and adjacent normal tissue, but nuclear AR expression is significantly higher in tumor cells [173]. Moreover, higher AR mRNA and protein expression correlate with greater overall survival, but higher AR activity, measured as higher expression of androgen-responsive genes, correlates with a higher histological grade and a poorer prognosis [174]. The latter is substantiated by further studies supporting the emerging view that activated and nuclear-localized AR and the corresponding increased AR activity correlate with worse HCC outcomes than overall AR mRNA and protein expression. This suggests that limiting AR signaling would be a viable therapeutic strategy. However, clinical trials of either directly targeting AR with an antagonist or suppressing gonadal androgen synthesis have not shown clinical efficacy [175,176].
One explanation for these disappointing clinical results is that the presence of constitutively active AR-SVs enhances AR signaling and renders the cells refractory to therapeutic interventions. Dauki and collaborators re-examined HCC TCGA patient data and found that 78% of patient tumors expressed AR-SVs, which are detected in male and female patients but are higher in males [177]. This analysis detected the expression of known constitutively active AR SVs, V1, V3, and V7, variants that are active in this cellular context. Depletion of all AR forms reduces cellular proliferation in the presence of androgens, confirming the importance of AR in growth. Overexpression of AR-V7 enhances mTOR signaling, while AR-V7 depletion hinders it, arguing that AR-SVs modulate this critical signaling pathway. Moreover, depletion of AR-V7 reduces cellular invasion and migration, indicative of a role in epithelial mesenchymal transition (EMT). Hence, AR-SVs have varied functions in HCC [177]. It is currently unclear what mechanism(s) drive AR-SV expression, but one explanation is that the observed differential expression of mRNA splicing factors promotes AR-SV expression [178].

5. Androgen Receptor Splice Variants in Non-Malignant Cells

In this section, we summarize the studies on AR-SVs in non-cancer conditions as well as in normal tissues (Table 1).

5.1. Polycystic Ovarian Syndrome

Polycystic Ovarian Syndrome (PCOS) is the most frequent endocrine disorder of reproductive-aged women, affecting up to 15% of women worldwide [179]. PCOS can cause infertility, hyperandrogenism, and metabolic disorders including obesity, insulin resistance, dyslipidemia, cardiovascular disease, and type 2 diabetes [179]. While the etiology of PCOS has not been completely deciphered, a key feature of this disease is an excess of androgen driven AR signaling [180,181,182]. Excess androgen production is most likely due to increased ovarian output [183]. In the ovary, AR is predominantly expressed by granulosa cells throughout follicular development. Notably, mice with global AR knockout are phenotypically female but have reduced fertility with abnormal ovarian function [184]. Granulosa cell-specific AR knockdown results in premature ovarian failure, decreased fertility, and slower follicle growth. These studies strongly argue that nearly all reproductive phenotypes associated with AR knockdown are due to a lack of AR activity in granulosa cells [185].
Human genetic studies implicate AR CAG repeat polymorphisms and AR SNP rs615G/A with PCOS, suggesting that alterations of the AR gene have a role in PCOS pathogenesis [186,187]. A study of granulosa cells obtained from women undergoing in vitro fertilization and embryo transfer identified two alternatively spliced AR variants expressed exclusively in women with PCOS [188]. One variant is a 69-bp insertion (ins) between exons 2 and 3, while the other has a deletion (del) that skips exon 3; both retain the LBD. Wildtype AR is co-expressed with the splice variants. However, on androgen treatment, the variant AR nuclear translocation is compromised. Notably, women expressing the AR splice variants have higher serum testosterone and DHEA levels than women with PCOS who do not express the variant AR. Chromatin precipitation sequence (ChIP seq) studies following expression of FL-AR or AR-SVs in primary granulosa cells found that the AR-SVs exhibit a different DNA binding pattern than FL-AR. FL-AR binds preferentially to classical AREs, while AR-SVs bind to distinct motifs. Parallel RNA-seq studies found that the FL-AR drives expression of genes related to follliculogenesis, while the AR-SV gene expression profiles differ. Additionally, the expression profiles of the 69 bp ins AR-SV and exon 3 del AR-SV differ from each other as well, indicating that gene cohorts regulated by specific AR-SVs are unique to that AR-SV. In granulosa cells, FL-AR transactivates expression of CYP19A1, the gene encoding aromatase, the rate-limiting enzyme for the conversion of androgens to estrogens, but the AR-SVs are unable to up-regulate this critical enzyme. Therefore, the AR-SV-encoded protein fails to upregulate the ratios of estradiol to total testosterone and estrone to androstenedione. However, the 69-bp insertion in AR-SV activates expression of CYP17A1, which encodes 17α-hydroxylase, which catalyzes androstenedione synthesis. These results argue that granulosa cells harboring AR-SVs have a leading role in the observed excess androgen accumulation in PCOS.

5.2. Placenta

The placenta develops at the maternal-fetal interface to support fetal development for nutrient and gas exchange and endocrine signaling. The placenta secretes androgens throughout pregnancy and expresses the AR nuclear receptor and the non-genomic membrane receptor, suggesting a role for androgen signaling in placental function. Androgens and the FL-AR are expressed in the placenta of humans [189,190], rodents [191], sheep [192], and cows [193] for the transcriptional regulation of genes involved in fetal growth and development.
On average, male fetal growth leads to increased birthweights and birthweight centiles compared with female fetal growth in normal pregnancies [194]. Circulating testosterone levels in women with preeclampsia are elevated 2–3 fold, and preeclampsia is associated with elevated placental AR expression. Animal models recapitulated these findings [195]. The importance of understanding placental androgens and AR has been shown during intrauterine development and in a sex-specific manner associated specifically with pregnancy complications. Maternal complications during pregnancy can impact fetal growth in utero, in part due to altered androgen signaling, and alterations in placental AR expression and localization may contribute to disrupted fetal development [196]. Women with preeclampsia have elevated androgen levels and AR gene expression, and it has been hypothesized that androgens affect vascular function and that vascular smooth muscle is involved in the development of preeclampsia. Moreover, it is postulated that an abnormal level of androgens negatively impacts placental angiogenesis and/or alters trophoblast cell proliferation and differentiation [195,197], although the exact mechanisms are unclear. In human pregnancies resulting in preeclampsia, there is an increase in placental FL-AR mRNA and protein expression, regardless of sex [190]. In AR-knockout mice, abnormalities in intrauterine growth and development are more pronounced in male fetuses [198].
AR-SVs have recently been reported in the human placenta which include the FL-AR, AR-V1, AR-V7, and AR-45 along with unknown proteins that are detected by anti-AR antibody including 120, 90, and 55 kDa isoforms [196]. These variants in the human placenta vary by fetal sex, cellular localization, and in the presence of pregnancy complications [196]. AR-FL, AR-V1, AR-V7, and AR-45 have also been recently reported in sheep placenta and their distribution and cellular localization are altered in a pregnancy complication including asthma [199]. In this sheep model for maternal asthma there is a decrease in nuclear expression of AR-45 and cytoplasmic expression of V1, proposing that AR-45, not AR-FL, may be responsible for growth [199]. Understanding placental androgen AR-SV signaling, could give more insight to certain developmental diseases and fetal mortality.

5.3. Testis

The AR is essential for the development of testicular tissue [200]. Sertoli cells, the somatic cells present in seminiferous tubules, are subject to AR signaling, which regulates Sertoli cell proliferation and maturation and the maintenance of the blood-testis barrier between cells [201]. Moreover, AR-dependent activities are also important for the surrounding germ cells [201]. Alternative splicing of nuclear hormone receptors is most frequently detected in testicular tissue, and alternatively spliced isoforms of steroid hormone receptors are commonly generated [202]. Several alternatively spliced ER isoforms have been found in the testis, where they modulate ER activity and may have a role in the regulation of spermatogenesis [203,204,205,206,207,208,209].
Early studies on AR-SV expression in testis identified AR45, an AR-SV that encodes a truncated NTD but a complete DBD, hinge, and DBD. Functional studies of this variant in non-testicular tissue found that AR45 could activate or repress AR transcriptional activity, depending on co-regulator expression, but studies were not conducted in the context of testicular cells [210]. Subsequent studies identified novel AR-SVs in human testis that encoded the NTD and deletions/truncations of exons 2, 3, or 4 [211]. ARΔ23stop (a deletion of exon 2, partial retention of intron 2, encoding a protein with only the NTD) and ARΔ4 (in-frame deletion of 120 nucleotides of exon 4 resulting in a protein with an incomplete LBD) were only detected in testis, while ARΔ2stop (premature stop encoding only the NTD) and ARΔ3 (in-frame skipping of exon 3 encoding a protein missing the 2nd Zn finger) were only detected in testis. This AR-SV is the same as the AR-SV described in PCOS, in breast cancer samples, and in patients with AIS [211,212,213], which was detected not only in the testis but in other tissues as well.
Studies on impaired spermatogenesis found AR-V7 (called AR3 in the study) is differentially expressed in testicular tissue from controls and patients with nonobstructive azoospermia (NOA) [214]. The NOA patient tissue was further stratified into patients with spermatogenic arrest at the levels of round spermatids (MA), patients with hypo-spermatogenesis (hypo), and patients with Sertoli cell-only syndrome (SCOS). Control tissue expresses FL-AR and AR-SV, while tissues from NOA patients all express similar levels of FL-AR but decreased levels of AR-SV (Control > hypo > MA > SCOS). There is a negative correlation between AR-V7 and serum follicular-stimulating hormone (FSH) levels. Moreover, FSH levels negatively correlate with the Johnsen score, a scoring system that quantifies spermatogenesis according to the profile of cells in seminiferous tubules. Decreased AR-V7 but not Fl-AR levels correlate with NOA, arguing that AR-V7 has a distinct function in spermatogenesis. Future studies will be needed to define the AR-V7 transcriptome and uncover the mechanistic basis of these observations.

5.4. Neuronal Lipid Raft

Membrane AR has not been as well studied as classical cytoplasmic/nuclear AR. The AR is expressed in many areas of the brain, including the substantia niagra, hippocampus, and entorhinal cortex [215,216,217]. Some non-DNA-binding-dependent actions of androgens through membrane-bound protein receptors for intracellular signaling have been suggested [218], and cell-impermeable androgens can have fast cellular effects [219,220,221,222,223,224]. Testosterone conjugated to bovine serum albumin leads to rapid membrane effects in N27 and SH-SY5Y neuronal cell lines, resulting in intracellular calcium release [222,223]. In the prostate LNCaP cell line, non-DNA-binding-dependent actions of androgens through membrane AR led to apoptosis and the reduction of tumor size [225]. It has been proposed that this binding antagonizes DNA-binding-dependent actions, arresting androgen action in target tissues.
Recently, Garza-Contreras et al. [226] reported expression of AR45 in a dopaminergic N27 cell line and in brain regions including the substantia nigra pars compacta, entorhinal cortex, and hippocampus from young and middle-aged gonadally intact and gonadectomized rats. The AR45 splice variant that has been characterized in humans lacks the NTD, decreasing the protein molecular weight from 110 kDa to 45 kDa [210]. AR45 is expressed in muscle, lung, heart, breast, uterus, and prostate [210,227,228] and in human brain tissue from an aged population [109]. In dopaminergic neurons, AR45 localizes to plasma membrane lipid rafts and interacts with membrane-associated G proteins Gαq and Gαo, proteins involved in modulating intracellular calcium release via the PLC/IP3 pathway. Hence, the AR-SV may function as a membrane AR to mediate fast, nongenomic androgen actions (Table 1) [226].

6. Androgen Receptor Splice Variants in Immune Cells

6.1. Peripheral Blood Mononuclear Cells

The immune responses of males and females differ [231]. Females have greater antibody responses to antigens and reject skin allografts more rapidly, hallmarks of greater cellular immune responses that are advantageous for surviving infectious diseases [232,233]. Females have a reduced incidence of tumors but, conversely, an increased incidence of autoimmune diseases consistent with greater cellular immune responses [232]. Sex differences were reproduced in models of autoimmune diseases and led to studies of sex differences in immune response to distinguish between the effects of hormones and genetic factors [234]. Lymphocytes have a central role in autoimmune disease development and progression [232] and express nuclear sex hormone receptors, which control epigenetic and transcriptional processes that can regulate expression of key immune-related genes either directly or indirectly [235,236]. The protective effects of testosterone may underlie the decrease in male susceptibility to autoimmune diseases. In the mouse model of nonobese diabetic disease, castration of males increased disease prevalence, while supplementation of females with testosterone lowered disease incidence [237,238].
These studies led to an evaluation of AR expression in distinct populations of peripheral blood mononuclear cells (PBMC), including B and T lymphocytes, monocytes/macrophages, neutrophil lineage cells, and natural killer cells [231]. Most neutrophils express AR, and androgens can promote neutrophil differentiation and recruitment in humans and mice. In the mouse model, deletion of the AR gene results in severe neutropenia, and neutrophil numbers decrease in mice following castration. AR-deficient animals exhibit normal phagocytic properties but respond less to granulocyte-colony stimulating factor induced proliferation and to migratory signals in vitro. They also exhibited greater susceptibility to apoptotic stimuli and produced fewer proinflammatory chemokines and cytokines. In line with these findings, prostate cancer patients undergoing androgen blockade display neutropenia.
Sub-populations enriched for CD4+ T lymphocytes (helper cells), CD8+ T lymphocytes (effector cells), and macrophages all express AR, while AR expression in B lymphocytes is much lower [239]. In vitro treatment of CD4+ lymphocytes with physiologically relevant levels of DHT results in increased expression of IL-10, a cytokine that suppresses the function of NK cells and T cells while enhancing B-cell survival. This suggests that androgens modify lymphocyte cytokine expression, altering the male immune system [239].
PBMC obtained from non-cancer patients express not only FL-AR but also AR-V7 [229]. Of the PBMC subpopulations studied, AR-V7 is detected in monocytes in all patients and in T- and B-cells and natural killer cells in some patients. FL-AR is detected in all subpopulations. FL-AR and AR-V7 assessment in PBMC and circulating tumor cells (CTC) from PCa patients treated with abiraterone (AA), enzalutamide (E), or taxanes found that high PBMC AR-V7 levels correlate with worse AA/E and better taxane responses. In taxane-treated patients, higher FL-AR expression in PBMC correlates with longer progression-free survival. However, high FL-AR and AR-V7 expression in CTC correlates with shorter progression-free survival and overall survival; therefore, expression in PBMC and CTC has differing predictive roles for therapeutic response. It is also notable that CTC-enriched fractions have detectable levels of the white blood cell marker CD45, indicating that they are contaminated with white blood cells. Testing whole blood cells for AR-V7 expression is premised on the assumption that the primary source of AR-V7 expression is CTC [240]. Identification of AR-V7 expression in lymphocytes suggests that CTC contamination with lymphocytes would contribute to the presence of AR-V7.

6.2. Fish Immune Cells during Development

AR protein structure is evolutionarily conserved from fish to humans [241]. Androgens in teleosts, ray-finned fish, are mainly mediated by nuclear AR and found in immune-competent organs of the sea bass, including the head-kidney, liver, and spleen [242], as well as in zebrafish [243]. This suggests that androgens may play a role in immune responses in some fish. The gilthead seabream is a seasonally breeding vertebrate marine fish in the Western Mediterranean. As hermaphrodite teleosts, they are male during at least two reproductive cycles, and each reproductive cycle is divided into four stages, including spermatogenesis, spawning, post-spawning, and resting. It is after the second resting stage that testicular involution occurs, which allows for the sex change [244,245]. Post-spawning, in the testis, there is a massive infiltration of immune cells (acidophilic granulocytes), which are regulated by testicular factors and hormones [244,245,246,247]. Both acidophilic granulocytes and macrophages express AR, with its expression being modulated by androgens [248].
Sanches-Hernandez et al. [230] report a constitutively active AR splice variant that lacks the ligand binding domain, AR∆LBD, in the gilthead seabream testis, head-kidney, and acidophilic granulocytes (equivalent to mammalian neutrophils). This variant has the NTD and DBD but lacks the LBD. Wild-type AR and AR∆LBD show different expression profiles in the testis during the reproductive cycles, suggesting that they play different roles. In addition, the expression ratio in the testis is negatively correlated with testosterone and 11-ketotestosterone throughout the reproductive cycles, while there is a positive correlation between the AR and AR∆LBD expression ratios in the head and kidney. In acidophilic granulocytes, AR and AR∆LBD varied during reproductive cycles and presented a positive correlation with testosterone similar to that in the head and kidney. Bacterial immune stimulation of acidophilic granulocytes regulates alternative splicing of AR, depending on reproductive stage. This suggests a crosstalk between endocrine and immune stimuli in the regulation of AR∆LBD and acidophilic granulocyte function in fish. It is worth noting that the ARΔ2stop and ARΔ3 SVs are also present in phylogenetically distinct vertebrate species—rat, African clawed frog, and gilthead seabream, a teleost fish [211].

6.3. Role of Androgen Receptor Splice Variants in Immunotherapy

Androgen signaling has implications for the response to immunotherapy. In preclinical studies, melanoma growth following anti-PD-L1 treatment was significantly greater in female animals than in male animals. This effect could be in part explained by inhibition of Treg upon PD-L1 blockade and thus an enhanced immune response, suggesting that androgens present in the male may have a negative impact on responses to anti-PD-L1 therapy [249]. A meta-analysis on checkpoint inhibitor efficacy in treatment for melanoma and non-small cell carcinoma found that overall survival was improved by these therapies for all patients, but the greatest therapeutic benefit was observed in men [250]. In a different meta-analysis, which surveyed checkpoint inhibitor efficacy in patients with various solid metastatic malignancies, significant differences in responses between men and women were not observed [251]. Moreover, where patients received a combination of chemotherapies and checkpoint inhibitors, women had better responses [252], hence the ambiguous data on sex-based responses to checkpoint inhibitors. Currently, the contribution of AR-SVs to checkpoint inhibitor therapies has not been documented.
Most of the studies on androgen signaling blockade and treatment efficacy focus on prostate cancer, where androgen deprivation therapy is a standard treatment for advanced malignancies [231]. Most immunotherapeutic approaches gave disappointing results due to a low tumor mutational burden and a “cold” tumor microenvironment [253]. Single-agent checkpoint inhibitor treatments in PD-L1-positive metastatic prostate cancers achieved positive outcomes in only a small percentage of patients, and in PD-L1-negative malignancies, the outcomes were even worse [254]. It has been noted that treatment with enzalutamide upregulated PD-L1 expression, and androgen deprivation is associated with T-cell infiltration of tumors and increased T-cell response in preclinical models [255,256,257]. Clinical trials of the combination treatment of enzalutamide and checkpoint inhibitors were remarkably effective in over 50% of patients [258,259]. Even though AR-SVs are commonly known to contribute to resistance to anti-androgen therapies, their contribution to or correlation with dual hormonal and immune checkpoint blockage has yet to be uncovered.
Expression of AR-SVs in human neutrophils has not been explored, but given their role in fish development and response to bacterial stimuli, such studies would further our understanding of the endocrine-immune axis in humans.

7. Implications, Unresolved Questions, and Conclusions

This comprehensive review of numerous studies that identified AR-SVs in various tissues reveals that AR-SV expression is pervasive across tumor and non-tumor tissues and in diverse species. When analyzed in total, the studies lead to several conclusions: (1) AR-SV expression is conserved through evolution, as evidenced by its presence in fish, frogs, and mammalian species. The AR-SV is similar across species and tissues, where most retain the NTD while the LBD is removed. The exception to this is AR45, which is missing the NTD but retains the DBD and LBD. In most variants, the DBD remains intact, but in some cases, the second ZN finger (encoded by exon 3) may be missing. (2) AR-SVs drive variant-specific transcriptomes that are distinct from those of the FL-AR, but there may be overlapping gene targets. Likewise, the AR-SV binding motifs differ from the FL-AR AREs. Studies in PCOS demonstrated that two different AR-SVs, generated by the insertion and deletion of sequences, drive distinct transcriptional programs; therefore, AR-SV-driven transcriptional programs are specific to a splice variant. Hence, AR-SVs are not simply substituting for FL-AR signaling but complementing, extending, or, in some cases, replacing FL-AR-specific gene expression. (3) AR-SV-driven transcription does not require FL-AR since some cells express only AR-SVs. Moreover, in many cells, a particular AR-SV may constitute a small percentage of total AR present in cells, yet its depletion exerts profound effects on cell proliferation, apoptosis, invasion, and migration, further arguing that AR-SVs have distinct and vital roles that are independent of FL-AR. Furthermore, AR-SVs and FL-ARs can heterodimerize. Is the transcriptional program driven by heterodimers distinct from either homodimer? (4) More recent studies have identified novel constitutively active AR-SVs, but most of the above studies focused on AR-SVs that were originally identified as constitutively active in PCa. It is possible that novel AR-SVs exist but remain undetected, and if true, the prevalence of AR-SVs may be even greater. Malignancies that are not as prevalent or as well studied as those reviewed above may also express these molecules where they may contribute to disease or therapeutic responses, so the prevalence of AR-SVs may be even greater. This requires further study. Moreover, defining the importance of AR-SVs in malignancies may expose a therapeutic opportunity. AR-SVs were detected in PCa cells, which are heavily reliant on AR activity; hence, significant efforts have focused on identifying agents that would target these isoforms along with the FL-AR. Since AR-SVs have a role in promoting the viability and tumorgenicity of multiple malignancies, such agents, when identified, could prove beneficial for the treatment of various malignancies where FL-AR and AR-SVs have a demonstrated role in pathogenesis.

Author Contributions

Conceptualization, K.D.K. and M.M.; original draft preparation, K.D.K. and M.M.; review and editing, P.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

Studies were supported in part by Veterans Affairs Biomedical laboratory Research and Development MERIT awards BX003458 (M.M.) and BX0004000 (P.M.G.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The contents reported/presented within do not represent the views of the Department of Veterans Affairs or the United States Government.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ipulan-Colet, L.A. Sexual dimorphism through androgen signaling; from external genitalia to muscles. Front. Endocrinol. 2022, 13, 940229. [Google Scholar] [CrossRef]
  2. Gerald, T.; Raj, G. Testosterone and the Androgen Receptor. Urol. Clin. North Am. 2022, 49, 603–614. [Google Scholar] [CrossRef]
  3. Mangelsdorf, D.J.; Thummel, C.; Beato, M.; Herrlich, P.; Schutz, G.; Umesono, K.; Blumberg, B.; Kastner, P.; Mark, M.; Chambon, P.; et al. The nuclear receptor superfamily: The second decade. Cell 1995, 83, 835–839. [Google Scholar] [CrossRef] [Green Version]
  4. Evans, R.M. The nuclear receptor superfamily: A rosetta stone for physiology. Mol. Endocrinol. 2005, 19, 1429–1438. [Google Scholar] [CrossRef]
  5. Messner, E.A.; Steele, T.M.; Tsamouri, M.M.; Hejazi, N.; Gao, A.C.; Mudryj, M.; Ghosh, P.M. The Androgen Receptor in Prostate Cancer: Effect of Structure, Ligands and Spliced Variants on Therapy. Biomedicines 2020, 8, 422. [Google Scholar] [CrossRef]
  6. Dahiya, U.R.; Heemers, H.V. Analyzing the Androgen Receptor Interactome in Prostate Cancer: Implications for Therapeutic Intervention. Cells 2022, 11, 936. [Google Scholar] [CrossRef] [PubMed]
  7. Rey, R.A. The Role of Androgen Signaling in Male Sexual Development at Puberty. Endocrinology 2021, 162, bqaa215. [Google Scholar] [CrossRef] [PubMed]
  8. Starka, L.; Duskova, M.; Vitku, J. 11-Keto-testosterone and other androgens of adrenal origin. Physiol. Res. 2020, 69, S187–S192. [Google Scholar] [CrossRef] [PubMed]
  9. Handelsman, D.J.; Hirschberg, A.L.; Bermon, S. Circulating Testosterone as the Hormonal Basis of Sex Differences in Athletic Performance. Endocr. Rev. 2018, 39, 803–829. [Google Scholar] [CrossRef] [Green Version]
  10. Weikum, E.R.; Liu, X.; Ortlund, E.A. The nuclear receptor superfamily: A structural perspective. Protein Sci. 2018, 27, 1876–1892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Migeon, B.R.; Brown, T.R.; Axelman, J.; Migeon, C.J. Studies of the locus for androgen receptor: Localization on the human X chromosome and evidence for homology with the Tfm locus in the mouse. Proc. Natl. Acad. Sci. USA. 1981, 78, 6339–6343. [Google Scholar] [CrossRef] [PubMed]
  12. Brinkmann, A.O.; Faber, P.W.; van Rooij, H.C.; Kuiper, G.G.; Ris, C.; Klaassen, P.; van der Korput, J.A.; Voorhorst, M.M.; van Laar, J.H.; Mulder, E.; et al. The human androgen receptor: Domain structure, genomic organization and regulation of expression. J. Steroid Biochem. 1989, 34, 307–310. [Google Scholar] [CrossRef] [PubMed]
  13. Veldscholte, J.; Berrevoets, C.A.; Zegers, N.D.; van der Kwast, T.H.; Grootegoed, J.A.; Mulder, E. Hormone-induced dissociation of the androgen receptor-heat-shock protein complex: Use of a new monoclonal antibody to distinguish transformed from nontransformed receptors. Biochemistry 1992, 31, 7422–7430. [Google Scholar] [CrossRef] [PubMed]
  14. Yu, X.; Yi, P.; Hamilton, R.A.; Shen, H.; Chen, M.; Foulds, C.E.; Mancini, M.A.; Ludtke, S.J.; Wang, Z.; O’Malley, B.W. Structural Insights of Transcriptionally Active, Full-Length Androgen Receptor Coactivator Complexes. Mol. Cell 2020, 79, 812–823.e814. [Google Scholar] [CrossRef]
  15. Smith, D.F.; Toft, D.O. Minireview: The intersection of steroid receptors with molecular chaperones: Observations and questions. Mol. Endocrinol. 2008, 22, 2229–2240. [Google Scholar] [CrossRef] [Green Version]
  16. Huggins, C.; Hodges, C.V. Studies on prostatic cancer. I. The effect of castration, of estrogen and of androgen injection on serum phosphatases in metastatic carcinoma of the prostate. J. Urol. 2002, 167, 948–951; discussion 952. [Google Scholar] [CrossRef]
  17. Magee, D.E.; Singal, R.K. Androgen deprivation therapy: Indications, methods of utilization, side effects and their management. Can. J. Urol. 2020, 27, 11–16. [Google Scholar]
  18. Shore, N.D.; Antonarakis, E.S.; Cookson, M.S.; Crawford, E.D.; Morgans, A.K.; Albala, D.M.; Hafron, J.; Harris, R.G.; Saltzstein, D.; Brown, G.A.; et al. Optimizing the role of androgen deprivation therapy in advanced prostate cancer: Challenges beyond the guidelines. Prostate 2020, 80, 527–544. [Google Scholar] [CrossRef] [Green Version]
  19. Yu, E.M.; Aragon-Ching, J.B. Advances with androgen deprivation therapy for prostate cancer. Exp. Opin. Pharmacother. 2022, 23, 1015–1033. [Google Scholar] [CrossRef]
  20. Roy, S.; Sayyid, R.; Saad, F.; Sun, Y.; Lajkosz, K.; Ong, M.; Klaassen, Z.; Malone, S.; Spratt, D.E.; Wallis, C.J.D.; et al. Addition of Docetaxel to Androgen Receptor Axis-targeted Therapy and Androgen Deprivation Therapy in Metastatic Hormone-sensitive Prostate Cancer: A Network Meta-analysis. Eur. Urol. Oncol. 2022, 5, 494–502. [Google Scholar]
  21. Yanagisawa, T.; Rajwa, P.; Thibault, C.; Gandaglia, G.; Mori, K.; Kawada, T.; Fukuokaya, W.; Shim, S.R.; Mostafaei, H.; Motlagh, R.S.; et al. Androgen Receptor Signaling Inhibitors in Addition to Docetaxel with Androgen Deprivation Therapy for Metastatic Hormone-sensitive Prostate Cancer: A Systematic Review and Meta-Analysis. Eur. Urol. 2022, 82, 584–598. [Google Scholar]
  22. Harris, W.P.; Mostaghel, E.A.; Nelson, P.S.; Montgomery, B. Androgen deprivation therapy: Progress in understanding mechanisms of resistance and optimizing androgen depletion. Nat. Clin. Pract. Urol. 2009, 6, 76–85. [Google Scholar] [CrossRef] [PubMed]
  23. Karantanos, T.; Corn, P.G.; Thompson, T.C. Prostate cancer progression after androgen deprivation therapy: Mechanisms of castrate resistance and novel therapeutic approaches. Oncogene 2013, 32, 5501–5511. [Google Scholar] [PubMed]
  24. Wadosky, K.M.; Koochekpour, S. Molecular mechanisms underlying resistance to androgen deprivation therapy in prostate cancer. Oncotarget 2016, 7, 64447–64470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Dai, C.; Dehm, S.M.; Sharifi, N. Targeting the Androgen Signaling Axis in Prostate Cancer. J. Clin. Oncol. 2023, JCO2300433. [Google Scholar] [CrossRef]
  26. Mudryj, M.; Tepper, C.G. On the origins of the androgen receptor low molecular weight species. Horm. Cancer 2013, 4, 259–269. [Google Scholar] [CrossRef] [Green Version]
  27. Ware, K.E.; Garcia-Blanco, M.A.; Armstrong, A.J.; Dehm, S.M. Biologic and clinical significance of androgen receptor variants in castration resistant prostate cancer. Endocr. Relat. Cancer 2014, 21, T87–T103. [Google Scholar] [CrossRef]
  28. Bryce, A.H.; Antonarakis, E.S. Androgen receptor splice variant 7 in castration-resistant prostate cancer: Clinical considerations. Int. J. Urol. 2016, 23, 646–653. [Google Scholar] [CrossRef]
  29. Imamura, Y.; Sadar, M.D. Androgen receptor targeted therapies in castration-resistant prostate cancer: Bench to clinic. Int. J. Urol. 2016, 23, 654–665. [Google Scholar] [CrossRef]
  30. Antonarakis, E.S.; Armstrong, A.J.; Dehm, S.M.; Luo, J. Androgen receptor variant-driven prostate cancer: Clinical implications and therapeutic targeting. Prostate Cancer Prostatic Dis. 2016, 19, 231–241. [Google Scholar] [CrossRef] [Green Version]
  31. 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] [PubMed]
  32. Dehm, S.M.; Schmidt, L.J.; Heemers, H.V.; Vessella, R.L.; Tindall, D.J. Splicing of a novel androgen receptor exon generates a constitutively active androgen receptor that mediates prostate cancer therapy resistance. Cancer Res. 2008, 68, 5469–5477. [Google Scholar] [CrossRef] [Green Version]
  33. 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]
  34. Libertini, S.J.; Tepper, C.G.; Rodriguez, V.; Asmuth, D.M.; Kung, H.J.; Mudryj, M. Evidence for calpain-mediated androgen receptor cleavage as a mechanism for androgen independence. Cancer Res. 2007, 67, 9001–9005. [Google Scholar] [CrossRef] [Green Version]
  35. 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]
  36. Sun, S.; Sprenger, C.C.; Vessella, R.L.; Haugk, K.; Soriano, K.; Mostaghel, E.A.; Page, S.T.; Coleman, I.M.; Nguyen, H.M.; Sun, H.; et al. Castration resistance in human prostate cancer is conferred by a frequently occurring androgen receptor splice variant. J. Clin. Investig. 2010, 120, 2715–2730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Luo, J.; Attard, G.; Balk, S.P.; Bevan, C.; Burnstein, K.; Cato, L.; Cherkasov, A.; De Bono, J.S.; Dong, Y.; Gao, A.C.; et al. Role of Androgen Receptor Variants in Prostate Cancer: Report from the 2017 Mission Androgen Receptor Variants Meeting. Eur. Urol. 2018, 73, 715–723. [Google Scholar]
  38. Lu, C.; Luo, J. Decoding the androgen receptor splice variants. Transl. Androl. Urol. 2013, 2, 178–186. [Google Scholar] [PubMed]
  39. Cato, L.; de Tribolet-Hardy, J.; Lee, I.; Rottenberg, J.T.; Coleman, I.; Melchers, D.; Houtman, R.; Xiao, T.; Li, W.; Uo, T.; et al. ARv7 Represses Tumor-Suppressor Genes in Castration-Resistant Prostate Cancer. Cancer Cell 2019, 35, 401–413.e406. [Google Scholar] [CrossRef]
  40. Basil, P.; Robertson, M.J.; Bingman, W.E., 3rd.; Dash, A.K.; Krause, W.C.; Shafi, A.A.; Piyarathna, B.; Coarfa, C.; Weigel, N.L. Cistrome and transcriptome analysis identifies unique androgen receptor (AR) and AR-V7 splice variant chromatin binding and transcriptional activities. Sci. Rep. 2022, 12, 5351. [Google Scholar] [CrossRef]
  41. Cai, L.; Tsai, Y.H.; Wang, P.; Wang, J.; Li, D.; Fan, H.; Zhao, Y.; Bareja, R.; Lu, R.; Wilson, E.M.; et al. ZFX Mediates Non-canonical Oncogenic Functions of the Androgen Receptor Splice Variant 7 in Castrate-Resistant Prostate Cancer. Mol. Cell 2018, 72, 341–354.e346. [Google Scholar] [CrossRef] [Green Version]
  42. Hu, R.; Lu, C.; Mostaghel, E.A.; Yegnasubramanian, S.; Gurel, M.; Tannahill, C.; Edwards, J.; Isaacs, W.B.; Nelson, P.S.; Bluemn, E.; et al. Distinct transcriptional programs mediated by the ligand-dependent full-length androgen receptor and its splice variants in castration-resistant prostate cancer. Cancer Res. 2012, 72, 3457–3462. [Google Scholar] [CrossRef] [Green Version]
  43. Krause, W.C.; Shafi, A.A.; Nakka, M.; Weigel, N.L. Androgen receptor and its splice variant, AR-V7, differentially regulate FOXA1 sensitive genes in LNCaP prostate cancer cells. Int. J. Biochem. Cell Biol. 2014, 54, 49–59. [Google Scholar] [CrossRef] [Green Version]
  44. Zhang, T.; Armstrong, A.J. Clinical Utility of Circulating Tumor Cells in Advanced Prostate Cancer. Curr. Oncol. Rep. 2016, 18, 3. [Google Scholar] [CrossRef]
  45. Zhang, T.; Karsh, L.I.; Nissenblatt, M.J.; Canfield, S.E. Androgen Receptor Splice Variant, AR-V7, as a Biomarker of Resistance to Androgen Axis-Targeted Therapies in Advanced Prostate Cancer. Clin. Genitourin. Cancer 2020, 18, 1–10. [Google Scholar] [CrossRef] [Green Version]
  46. Enikeev, D.; Morozov, A.; Babaevskaya, D.; Bazarkin, A.; Malavaud, B. A Systematic Review of Circulating Tumor Cells Clinical Application in Prostate Cancer Diagnosis. Cancers 2022, 14, 3802. [Google Scholar] [CrossRef]
  47. Li, Y.; Alsagabi, M.; Fan, D.; Bova, G.S.; Tewfik, A.H.; Dehm, S.M. Intragenic rearrangement and altered RNA splicing of the androgen receptor in a cell-based model of prostate cancer progression. Cancer Res. 2011, 71, 2108–2117. [Google Scholar] [CrossRef] [Green Version]
  48. Tepper, C.G.; Boucher, D.L.; Ryan, P.E.; Ma, A.H.; Xia, L.; Lee, L.F.; Pretlow, T.G.; Kung, H.J. Characterization of a novel androgen receptor mutation in a relapsed CWR22 prostate cancer xenograft and cell line. Cancer Res. 2002, 62, 6606–6614. [Google Scholar] [PubMed]
  49. Henzler, C.; Li, Y.; Yang, R.; McBride, T.; Ho, Y.; Sprenger, C.; Liu, G.; Coleman, I.; Lakely, B.; Li, R.; et al. Truncation and constitutive activation of the androgen receptor by diverse genomic rearrangements in prostate cancer. Nat. Commun. 2016, 7, 13668. [Google Scholar] [CrossRef] [PubMed]
  50. De Laere, B.; van Dam, P.J.; Whitington, T.; Mayrhofer, M.; Diaz, E.H.; Van den Eynden, G.; Vandebroek, J.; Del-Favero, J.; Van Laere, S.; Dirix, L.; et al. Comprehensive Profiling of the Androgen Receptor in Liquid Biopsies from Castration-resistant Prostate Cancer Reveals Novel Intra-AR Structural Variation and Splice Variant Expression Patterns. Eur. Urol. 2017, 72, 192–200. [Google Scholar] [CrossRef] [PubMed]
  51. Paschalis, A.; Sharp, A.; Welti, J.C.; Neeb, A.; Raj, G.V.; Luo, J.; Plymate, S.R.; de Bono, J.S. Alternative splicing in prostate cancer. Nat. Rev. Clin. Oncol. 2018, 15, 663–675. [Google Scholar] [CrossRef]
  52. Nadiminty, N.; Tummala, R.; Liu, C.; Lou, W.; Evans, C.P.; Gao, A.C. NF-kappaB2/p52:c-Myc:hnRNPA1 Pathway Regulates Expression of Androgen Receptor Splice Variants and Enzalutamide Sensitivity in Prostate Cancer. Mol. Cancer Ther. 2015, 14, 1884–1895. [Google Scholar] [CrossRef] [Green Version]
  53. Cao, S.; Zhan, Y.; Dong, Y. Emerging data on androgen receptor splice variants in prostate cancer. Endocr. Relat. Cancer 2016, 23, T199–T210. [Google Scholar] [CrossRef]
  54. Liu, L.L.; Xie, N.; Sun, S.; Plymate, S.; Mostaghel, E.; Dong, X. Mechanisms of the androgen receptor splicing in prostate cancer cells. Oncogene 2014, 33, 3140–3150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Ferraldeschi, R.; Welti, J.; Powers, M.V.; Yuan, W.; Smyth, T.; Seed, G.; Riisnaes, R.; Hedayat, S.; Wang, H.; Crespo, M.; et al. Second-Generation HSP90 Inhibitor Onalespib Blocks mRNA Splicing of Androgen Receptor Variant 7 in Prostate Cancer Cells. Cancer Res. 2016, 76, 2731–2742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zhang, Z.; Zhou, N.; Huang, J.; Ho, T.T.; Zhu, Z.; Qiu, Z.; Zhou, X.; Bai, C.; Wu, F.; Xu, M.; et al. Regulation of androgen receptor splice variant AR3 by PCGEM1. Oncotarget 2016, 7, 15481–15491. [Google Scholar] [CrossRef]
  57. Adamiecki, R.; Hryniewicz-Jankowska, A.; Ortiz, M.A.; Li, X.; Porter-Hansen, B.A.; Nsouli, I.; Bratslavsky, G.; Kotula, L. In Vivo Models for Prostate Cancer Research. Cancers 2022, 14, 5321. [Google Scholar] [CrossRef] [PubMed]
  58. Sun, F.; Chen, H.G.; Li, W.; Yang, X.; Wang, X.; Jiang, R.; Guo, Z.; Chen, H.; Huang, J.; Borowsky, A.D.; et al. Androgen receptor splice variant AR3 promotes prostate cancer via modulating expression of autocrine/paracrine factors. J. Biol. Chem. 2014, 289, 1529–1539. [Google Scholar] [CrossRef] [Green Version]
  59. Liu, G.; Sprenger, C.; Sun, S.; Epilepsia, K.S.; Haugk, K.; Zhang, X.; Coleman, I.; Nelson, P.S.; Plymate, S. AR variant ARv567es induces carcinogenesis in a novel transgenic mouse model of prostate cancer. Neoplasia 2013, 15, 1009–1017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Ji, Y.; Zhang, R.; Han, X.; Zhou, J. Targeting the N-terminal domain of the androgen receptor: The effective approach in therapy of CRPC. Eur. J. Med. Chem. 2023, 247, 115077. [Google Scholar] [CrossRef]
  61. Andersen, R.J.; Mawji, N.R.; Wang, J.; Wang, G.; Haile, S.; Myung, J.K.; Watt, K.; Tam, T.; Yang, Y.C.; Banuelos, C.A.; et al. Regression of castrate-recurrent prostate cancer by a small-molecule inhibitor of the amino-terminus domain of the androgen receptor. Cancer Cell 2010, 17, 535–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Myung, J.K.; Banuelos, C.A.; Fernandez, J.G.; Mawji, N.R.; Wang, J.; Tien, A.H.; Yang, Y.C.; Tavakoli, I.; Haile, S.; Watt, K.; et al. An androgen receptor N-terminal domain antagonist for treating prostate cancer. J. Clin. Investig. 2013, 123, 2948–2960. [Google Scholar] [CrossRef] [Green Version]
  63. Brand, L.J.; Olson, M.E.; Ravindranathan, P.; Guo, H.; Kempema, A.M.; Andrews, T.E.; Chen, X.; Raj, G.V.; Harki, D.A.; Dehm, S.M. EPI-001 is a selective peroxisome proliferator-activated receptor-gamma modulator with inhibitory effects on androgen receptor expression and activity in prostate cancer. Oncotarget 2015, 6, 3811–3824. [Google Scholar] [CrossRef] [Green Version]
  64. Ponnusamy, S.; Coss, C.C.; Thiyagarajan, T.; Watts, K.; Hwang, D.J.; He, Y.; Selth, L.A.; McEwan, I.J.; Duke, C.B.; Pagadala, J.; et al. Novel Selective Agents for the Degradation of Androgen Receptor Variants to Treat Castration-Resistant Prostate Cancer. Cancer Res. 2017, 77, 6282–6298. [Google Scholar] [CrossRef] [Green Version]
  65. He, Y.; Hwang, D.J.; Ponnusamy, S.; Thiyagarajan, T.; Mohler, M.L.; Narayanan, R.; Miller, D.D. Exploration and Biological Evaluation of Basic Heteromonocyclic Propanamide Derivatives as SARDs for the Treatment of Enzalutamide-Resistant Prostate Cancer. J. Med. Chem. 2021, 64, 11045–11062. [Google Scholar] [CrossRef] [PubMed]
  66. Tran, T.T.; Song, C.H.; Kim, K.J.; Lee, K. A new compound targets the AF-1 of androgen receptor and decreases its activity and protein levels in prostate cancer cells. Am. J. Cancer Res. 2020, 10, 4607–4623. [Google Scholar]
  67. Goicochea, N.L.; Garnovskaya, M.; Blanton, M.G.; Chan, G.; Weisbart, R.; Lilly, M.B. Development of cell-penetrating bispecific antibodies targeting the N-terminal domain of androgen receptor for prostate cancer therapy. Protein Eng. Des. Sel. 2017, 30, 785–793. [Google Scholar] [CrossRef] [PubMed]
  68. Rubin, J.B.; Lagas, J.S.; Broestl, L.; Sponagel, J.; Rockwell, N.; Rhee, G.; Rosen, S.F.; Chen, S.; Klein, R.S.; Imoukhuede, P.; et al. Sex differences in cancer mechanisms. Biol. Sex Differ. 2017, 11, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Lopes-Ramos, C.M.; Quackenbush, J.; DeMeo, D.L. Genome-Wide Sex and Gender Differences in Cancer. Front. Oncol. 2020, 10, 597788. [Google Scholar] [CrossRef]
  70. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer statistics, 2023. CA Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef]
  71. Clocchiatti, A.; Cora, E.; Zhang, Y.; Dotto, G.P. Sexual dimorphism in cancer. Nat. Rev. Cancer 2016, 16, 330–339. [Google Scholar] [CrossRef] [Green Version]
  72. Santen, R.J. Assessing individual risk for breast cancer: Role of oestrogens and androgens. Breast Cancer Res. 2008, 10, S10. [Google Scholar] [CrossRef]
  73. Santen, R.J. Menopausal hormone therapy and breast cancer. J. Steroid Biochem. Mol. Biol. 2014, 142, 52–61. [Google Scholar] [CrossRef] [PubMed]
  74. Dai, X.; Xiang, L.; Li, T.; Bai, Z.; Hallmarks, C. Biomarkers and Breast Cancer Molecular Subtypes. J. Cancer 2016, 7, 1281–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hall, R.E.; Clements, J.A.; Birrell, S.N.; Tilley, W.D. Prostate-specific antigen and gross cystic disease fluid protein-15 are co-expressed in androgen receptor-positive breast tumours. Br. J. Cancer 1998, 78, 360–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kuenen-Boumeester, V.; Van der Kwast, T.H.; van Putten, W.L.; Claassen, C.; van Ooijen, B.; Henzen-Logmans, S.C. Immunohistochemical determination of androgen receptors in relation to oestrogen and progesterone receptors in female breast cancer. Int. J. Cancer 1992, 52, 581–584. [Google Scholar] [CrossRef]
  77. Agoff, S.N.; Swanson, P.E.; Linden, H.; Hawes, S.E.; Lawton, T.J. Androgen receptor expression in estrogen receptor-negative breast cancer. Immunohistochemical, clinical, and prognostic associations. Am. J. Clin. Pathol. 2003, 120, 725–731. [Google Scholar] [CrossRef] [PubMed]
  78. Kaaks, R.; Rinaldi, S.; Key, T.J.; Berrino, F.; Peeters, P.H.; Biessy, C.; Dossus, L.; Lukanova, A.; Bingham, S.; Khaw, K.T.; et al. Postmenopausal serum androgens, oestrogens and breast cancer risk: The European prospective investigation into cancer and nutrition. Endocr. Relat. Cancer 2005, 12, 1071–1082. [Google Scholar] [CrossRef] [Green Version]
  79. Secreto, G.; Toniolo, P.; Pisani, P.; Recchione, C.; Cavalleri, A.; Fariselli, G.; Totis, A.; Di Pietro, S.; Berrino, F. Androgens and breast cancer in premenopausal women. Cancer Res. 1989, 49, 471–476. [Google Scholar]
  80. Grattarola, R. The Premenstrual Endometrial Pattern of Women with Breast Cancer. A Study of Progestational Activity. Cancer 1964, 17, 1119–1122. [Google Scholar] [CrossRef]
  81. Somboonporn, W.; Davis, S.R. Postmenopausal testosterone therapy and breast cancer risk. Maturitas 2004, 49, 267–275. [Google Scholar] [CrossRef] [PubMed]
  82. Key, T.; Appleby, P.; Barnes, I.; Reeves, G.; Endogenous, H.; Collaborative, G.B.C. Endogenous sex hormones and breast cancer in postmenopausal women: Reanalysis of nine prospective studies. J. Natl. Cancer Inst. 2002, 94, 606–616. [Google Scholar]
  83. Secreto, G.; Zumoff, B. Abnormal production of androgens in women with breast cancer. Anticancer Res. 1994, 14, 2113–2117. [Google Scholar] [PubMed]
  84. 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. [Google Scholar]
  85. Brys, M.; Wojcik, M.; Romanowicz-Makowska, H.; Krajewska, W.M. Androgen receptor status in female breast cancer: RT-PCR and Western blot studies. J. Cancer Res. Clin. Oncol. 2002, 128, 85–90. [Google Scholar]
  86. Bieche, I.; Parfait, B.; Tozlu, S.; Lidereau, R.; Vidaud, M. Quantitation of androgen receptor gene expression in sporadic breast tumors by real-time RT-PCR: Evidence that MYC is an AR-regulated gene. Carcinogenesis 2001, 22, 1521–1526. [Google Scholar] [CrossRef] [Green Version]
  87. Soreide, J.A.; Lea, O.A.; Varhaug, J.E.; Skarstein, A.; Kvinnsland, S. Androgen receptors in operable breast cancer: Relation to other steroid hormone receptors, correlations to prognostic factors and predictive value for effect of adjuvant tamoxifen treatment. Eur. J. Surg. Oncol. 1992, 18, 112–118. [Google Scholar]
  88. 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]
  89. 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] [PubMed]
  90. 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]
  91. 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] [Green Version]
  92. 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 2017, 24, 859–864. [Google Scholar] [CrossRef] [Green Version]
  93. 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]
  94. 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.e476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Gordon, M.A.; D’Amato, N.C.; Gu, H.; Babbs, B.; Wulfkuhle, J.; Petricoin, E.F.; Gallagher, I.; Dong, T.; Torkko, K.; Liu, B.; 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] [Green Version]
  96. 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]
  97. 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] [PubMed] [Green Version]
  98. 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] [PubMed] [Green Version]
  99. Ni, M.; Chen, Y.; 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]
  100. 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]
  101. 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] [PubMed]
  102. 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] [PubMed] [Green Version]
  103. 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] [PubMed]
  104. Isola, J.J. Immunohistochemical demonstration of androgen receptor in breast cancer and its relationship to other prognostic factors. J. Pathol. 1993, 170, 31–35. [Google Scholar] [CrossRef] [PubMed]
  105. 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]
  106. 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] [Green Version]
  107. 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]
  108. 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]
  109. Hu, D.G.; Hickey, T.E.; Irvine, C.; Wijayakumara, D.D.; Lu, L.; Tilley, W.D.; Selth, L.A.; Mackenzie, P.I. Identification of androgen receptor splice variant transcripts in breast cancer cell lines and human tissues. Horm. Cancer 2014, 5, 61–71. [Google Scholar]
  110. Hickey, T.E.; Irvine, C.M.; Dvinge, H.; Tarulli, G.A.; Hanson, A.R.; Ryan, N.K.; Pickering, M.A.; Birrell, S.N.; Hu, D.G.; Mackenzie, P.I.; et al. Expression of androgen receptor splice variants in clinical breast cancers. Oncotarget 2015, 6, 44728–44744. [Google Scholar] [CrossRef] [Green Version]
  111. Aceto, N.; Bardia, A.; Wittner, B.S.; Donaldson, M.C.; O’Keefe, R.; Engstrom, A.; Bersani, F.; Zheng, Y.; Comaills, V.; Niederhoffer, K.; et al. AR Expression in Breast Cancer CTCs Associates with Bone Metastases. Mol. Cancer Res. 2018, 16, 720–727. [Google Scholar] [CrossRef] [Green Version]
  112. Kasimir-Bauer, S.; Keup, C.; Hoffmann, O.; Hauch, S.; Kimmig, R.; Bittner, A.K. Circulating Tumor Cells Expressing the Prostate Specific Membrane Antigen (PSMA) Indicate Worse Outcome in Primary, Non-Metastatic Triple-Negative Breast Cancer. Front. Oncol. 2020, 10, 1658. [Google Scholar] [PubMed]
  113. Ferguson, D.C.; Mata, D.A.; Tay, T.K.; Traina, T.A.; Gucalp, A.; Chandarlapaty, S.; D’Alfonso, T.M.; Brogi, E.; Mullaney, K.; Ladanyi, M.; et al. Androgen receptor splice variant-7 in breast cancer: Clinical and pathologic correlations. Mod. Pathol. 2022, 35, 396–402. [Google Scholar] [CrossRef] [PubMed]
  114. Mitani, Y.; Rao, P.H.; Maity, S.N.; Lee, Y.C.; Ferrarotto, R.; Post, J.C.; Licitra, L.; Lippman, S.M.; Kies, M.S.; Weber, R.S.; et al. Alterations associated with androgen receptor gene activation in salivary duct carcinoma of both sexes: Potential therapeutic ramifications. Clin. Cancer Res. 2014, 20, 6570–6581.e814. [Google Scholar]
  115. Dalin, M.G.; Desrichard, A.; Katabi, N.; Makarov, V.; Walsh, L.A.; Lee, K.W.; Wang, Q.; Armenia, J.; West, L.; Dogan, S.; et al. Comprehensive Molecular Characterization of Salivary Duct Carcinoma Reveals Actionable Targets and Similarity to Apocrine Breast Cancer. Clin. Cancer Res. 2016, 22, 4623–4633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Simpson, R.H. Salivary duct carcinoma: New developments—Morphological variants including pure in situ high grade lesions; proposed molecular classification. Head Neck Pathol. 2013, 7, S48–S58. [Google Scholar] [PubMed] [Green Version]
  117. Di Palma, S.; Simpson, R.H.; Marchio, C.; Skalova, A.; Ungari, M.; Sandison, A.; Whitaker, S.; Parry, S.; Reis-Filho, J.S. Salivary duct carcinomas can be classified into luminal androgen receptor-positive, HER2 and basal-like phenotypes. Histopathology 2012, 61, 629–643. [Google Scholar] [CrossRef]
  118. Chiosea, S.I.; Williams, L.; Griffith, C.C.; Thompson, L.D.; Weinreb, I.; Bauman, J.E.; Luvison, A.; Roy, S.; Seethala, R.R.; Nikiforova, M.N. Molecular characterization of apocrine salivary duct carcinoma. Am. J. Surg. Pathol. 2015, 39, 744–752. [Google Scholar] [CrossRef]
  119. Luk, P.P.; Weston, J.D.; Yu, B.; Selinger, C.I.; Ekmejian, R.; Eviston, T.J.; Lum, T.; Gao, K.; Boyer, M.; O’Toole, S.A.; et al. Salivary duct carcinoma: Clinicopathologic features, morphologic spectrum, and somatic mutations. Head Neck 2016, 38, E1838–E1847. [Google Scholar]
  120. Williams, L.; Thompson, L.D.; Seethala, R.R.; Weinreb, I.; Assaad, A.M.; Tuluc, M.; Din, N.U.; Purgina, B.; Lai, C.; Griffith, C.C.; et al. Salivary duct carcinoma: The predominance of apocrine morphology, prevalence of histologic variants and androgen receptor expression. Am. J. Surg. Pathol. 2015, 39, 705–713. [Google Scholar]
  121. Locati, L.D.; Quattrone, P.; Bossi, P.; Marchiano, A.V.; Cantu, G.; Licitra, L. A complete remission with androgen-deprivation therapy in a recurrent androgen receptor-expressing adenocarcinoma of the parotid gland. Ann. Oncol. 2003, 14, 1327–1328. [Google Scholar] [CrossRef]
  122. Jaspers, H.C.; Verbist, B.M.; Schoffelen, R.; Mattijssen, V.; Slootweg, P.J.; van der Graaf, W.T.; van Herpen, C.M. Androgen receptor-positive salivary duct carcinoma: A disease entity with promising new treatment options. J. Clin. Oncol. 2011, 29, e473–e476. [Google Scholar] [PubMed]
  123. Boon, E.; van Boxtel, W.; Buter, J.; de Jong, R.J.B.; van Es, R.J.J.; Bel, M.; Fiets, E.; Oosting, S.F.; Slingerland, M.; Hoeben, A.; et al. Androgen deprivation therapy for androgen receptor-positive advanced salivary duct carcinoma: A nationwide case series of 35 patients in The Netherlands. Head Neck 2018, 40, 605–613. [Google Scholar] [PubMed]
  124. van Boxtel, W.; Locati, L.D.; van Engen-van Grunsven, A.C.H.; Bergamini, C.; Jonker, M.A.; Fiets, E.; Cavalieri, S.; Tooten, S.; Bos, E.; Quattrone, P.; et al. Adjuvant androgen deprivation therapy for poor-risk, androgen receptor-positive salivary duct carcinoma. Eur. J. Cancer 2019, 110, 62–70. [Google Scholar] [CrossRef]
  125. Fushimi, C.; Tada, Y.; Takahashi, H.; Nagao, T.; Ojiri, H.; Masubuchi, T.; Matsuki, T.; Miura, K.; Kawakita, D.; Hirai, H.; et al. A prospective phase II study of combined androgen blockade in patients with androgen receptor-positive metastatic or locally advanced unresectable salivary gland carcinoma. Ann. Oncol. 2018, 29, 979–984. [Google Scholar] [PubMed] [Green Version]
  126. Kawakita, D.; Nagao, T.; Takahashi, H.; Kano, S.; Honma, Y.; Hirai, H.; Saigusa, N.; Akazawa, K.; Tani, K.; Ojiri, H.; et al. Survival benefit of HER2-targeted or androgen deprivation therapy in salivary duct carcinoma. Ther. Adv. Med. Oncol. 2022, 14, 17588359221119538. [Google Scholar] [CrossRef]
  127. Ho, A.L.; Foster, N.R.; Zoroufy, A.J.; Campbell, J.D.; Worden, F.; Price, K.; Adkins, D.; Bowles, D.W.; Kang, H.; Burtness, B.; et al. Phase II Study of Enzalutamide for Patients with Androgen Receptor-Positive Salivary Gland Cancers (Alliance A091404). J. Clin. Oncol. 2022, 40, 4240–4249. [Google Scholar]
  128. Yang, R.K.; Zhao, P.; Lu, C.; Luo, J.; Hu, R. Expression pattern of androgen receptor and AR-V7 in androgen-deprivation therapy-naive salivary duct carcinomas. Hum. Pathol. 2019, 84, 173–182. [Google Scholar] [CrossRef] [PubMed]
  129. Gargano, S.M.; Senarathne, W.; Feldman, R.; Florento, E.; Stafford, P.; Swensen, J.; Vranic, S.; Gatalica, Z. Novel therapeutic targets in salivary duct carcinoma uncovered by comprehensive molecular profiling. Cancer Med. 2019, 8, 7322–7329. [Google Scholar] [CrossRef] [Green Version]
  130. Ostrom, Q.T.; Patil, N.; Cioffi, G.; Waite, K.; Kruchko, C.; Barnholtz-Sloan, J.S. CBTRUS Statistical Report: Primary Brain and Other Central Nervous System Tumors Diagnosed in the United States in 2013–2017. Neuro-Oncology 2020, 22, iv1–iv96. [Google Scholar]
  131. Gittleman, H.; Ostrom, Q.T.; Stetson, L.C.; Waite, K.; Hodges, T.R.; Wright, C.H.; Wright, J.; Rubin, J.B.; Berens, M.E.; Lathia, J.; et al. Sex is an important prognostic factor for glioblastoma but not for nonglioblastoma. Neurooncol. Pract. 2019, 6, 451–462. [Google Scholar] [CrossRef]
  132. Broestl, L.; Warrington, N.M.; Grandison, L.; Abou-Antoun, T.; Tung, O.; Shenoy, S.; Tallman, M.M.; Rhee, G.; Yang, W.; Sponagel, J.; et al. Gonadal sex patterns p21-induced cellular senescence in mouse and human glioblastoma. Commun. Biol. 2022, 5, 781. [Google Scholar] [CrossRef]
  133. Yu, X.; Jiang, Y.; Wei, W.; Cong, P.; Ding, Y.; Xiang, L.; Wu, K. Androgen receptor signaling regulates growth of glioblastoma multiforme in men. Tumour Biol. 2015, 36, 967–972. [Google Scholar] [CrossRef]
  134. Bao, D.; Cheng, C.; Lan, X.; Xing, R.; Chen, Z.; Zhao, H.; Sun, J.; Wang, Y.; Niu, C.; Zhang, B.; et al. Regulation of p53wt glioma cell proliferation by androgen receptor-mediated inhibition of small VCP/p97-interacting protein expression. Oncotarget 2017, 8, 23142–23154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Rodriguez-Lozano, D.C.; Velazquez-Vazquez, D.E.; Del Moral-Morales, A.; Camacho-Arroyo, I. Dihydrotestosterone Induces Proliferation, Migration and Invasion of Human Glioblastoma Cell Lines. Onco Targets Ther. 2020, 13, 8813–8823. [Google Scholar] [CrossRef] [PubMed]
  136. Verzat, C.; Delisle, M.B.; Courriere, P.; Hollande, E. Influence of host sex on the growth of a human glioblastoma line in athymic mice. Neuropathol. Appl. Neurobiol. 1990, 16, 141–151. [Google Scholar] [CrossRef]
  137. Chen, T.C.; Chuang, J.Y.; Ko, C.Y.; Kao, T.J.; Yang, P.Y.; Yu, C.H.; Liu, M.S.; Hu, S.L.; Tsai, Y.T.; Chan, H.; et al. AR ubiquitination induced by the curcumin analog suppresses growth of temozolomide-resistant glioblastoma through disrupting GPX4-Mediated redox homeostasis. Redox Biol. 2020, 30, 101413. [Google Scholar] [CrossRef]
  138. Rodriguez-Lozano, D.C.; Pina-Medina, A.G.; Hansberg-Pastor, V.; Bello-Alvarez, C.; Camacho-Arroyo, I. Testosterone Promotes Glioblastoma Cell Proliferation, Migration and Invasion Through Androgen Receptor Activation. Front. Endocrinol. 2019, 10, 16. [Google Scholar] [CrossRef] [Green Version]
  139. Orozco, M.; Valdez, R.A.; Ramos, L.; Cabeza, M.; Segovia, J.; Romano, M.C. Dutasteride combined with androgen receptor antagonists inhibit glioblastoma U87 cell metabolism, proliferation and invasion capacity: Androgen regulation. Steroids 2020, 164, 108733. [Google Scholar] [PubMed]
  140. Zalcman, N.; Canello, T.; Ovadia, H.; Charbit, H.; Zelikovitch, B.; Mordechai, A.; Fellig, Y.; Rabani, S.; Shahar, T.; Lossos, A.; et al. Androgen receptor: A potential therapeutic target for glioblastoma. Oncotarget 2018, 9, 19980–19993. [Google Scholar] [CrossRef] [Green Version]
  141. Scelo, G.; Li, P.; Chanudet, E.; Muller, D.C. Variability of Sex Disparities in Cancer Incidence over 30 Years: The Striking Case of Kidney Cancer. Eur. Urol. Focus 2018, 4, 586–590. [Google Scholar]
  142. Peired, A.J.; Campi, R.; Angelotti, M.L.; Antonelli, G.; Conte, C.; Lazzeri, E.; Becherucci, F.; Calistri, L.; Serni, S.; Romagnani, P. Sex and Gender Differences in Kidney Cancer: Clinical and Experimental Evidence. Cancers 2021, 13, 4588. [Google Scholar] [CrossRef]
  143. Langner, C.; Ratschek, M.; Rehak, P.; Schips, L.; Zigeuner, R. Steroid hormone receptor expression in renal cell carcinoma: An immunohistochemical analysis of 182 tumors. J. Urol. 2004, 171, 611–614. [Google Scholar]
  144. Zhu, G.; Liang, L.; Li, L.; Dang, Q.; Song, W.; Yeh, S.; He, D.; Chang, C. The expression and evaluation of androgen receptor in human renal cell carcinoma. Urology 2014, 83, 510.e19–510.e24. [Google Scholar] [CrossRef]
  145. Ha, Y.S.; Lee, G.T.; Modi, P.; Kwon, Y.S.; Ahn, H.; Kim, W.J.; Kim, I.Y. Increased Expression of Androgen Receptor mRNA in Human Renal Cell Carcinoma Cells is Associated with Poor Prognosis in Patients with Localized Renal Cell Carcinoma. J. Urol. 2015, 194, 1441–1448. [Google Scholar] [CrossRef]
  146. Yuan, P.; Ge, Y.; Liu, X.; Wang, S.; Ye, Z.; Xu, H.; Chen, Z. The Association of Androgen Receptor Expression with Renal Cell Carcinoma Risk: A Systematic Review and Meta-Analysis. Pathol. Oncol. Res. 2020, 26, 605–614. [Google Scholar] [CrossRef] [PubMed]
  147. Noh, S.J.; Kang, M.J.; Kim, K.M.; Bae, J.S.; Park, H.S.; Moon, W.S.; Chung, M.J.; Lee, H.; Lee, D.G.; Jang, K.Y. Acetylation status of P53 and the expression of DBC1, SIRT1 and androgen receptor are associated with survival in clear cell renal cell carcinoma patients. Pathology 2013, 45, 574–580. [Google Scholar] [CrossRef]
  148. Pak, S.; Kim, W.; Kim, Y.; Song, C.; Ahn, H. Dihydrotestosterone promotes kidney cancer cell proliferation by activating the STAT5 pathway via androgen and glucocorticoid receptors. J. Cancer Res. Clin. Oncol. 2019, 145, 2293–2301. [Google Scholar] [CrossRef] [PubMed]
  149. Lee, G.T.; Han, C.S.; Kwon, Y.S.; Patel, R.; Modi, P.K.; Kwon, S.J.; Faiena, I.; Patel, N.; Singer, E.A.; Ahn, H.J.; et al. Intracrine androgen biosynthesis in renal cell carcinoma. Br. J. Cancer 2017, 116, 937–943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Makhov, P.; Joshi, S.; Ghatalia, P.; Kutikov, A.; Uzzo, R.G.; Kolenko, V.M. Resistance to Systemic Therapies in Clear Cell Renal Cell Carcinoma: Mechanisms and Management Strategies. Mol. Cancer Ther. 2018, 17, 1355–1364. [Google Scholar]
  151. Muglia, V.F.; Prando, A. Renal cell carcinoma: Histological classification and correlation with imaging findings. Radiol. Bras. 2015, 48, 166–174. [Google Scholar]
  152. Bialek, J.; Piwonka, M.; Kawan, F.; Fornara, P.; Theil, G. Differential Expression of the Androgen Receptor, Splice Variants and Relaxin 2 in Renal Cancer. Life 2021, 11, 731. [Google Scholar] [CrossRef] [PubMed]
  153. Foersch, S.; Schindeldecker, M.; Keith, M.; Tagscherer, K.E.; Fernandez, A.; Stenzel, P.J.; Pahernik, S.; Hohenfellner, M.; Schirmacher, P.; Roth, W.; et al. Prognostic relevance of androgen receptor expression in renal cell carcinomas. Oncotarget 2017, 8, 78545–78555. [Google Scholar] [CrossRef] [Green Version]
  154. Neschadim, A.; Summerlee, A.J.; Silvertown, J.D. Targeting the relaxin hormonal pathway in prostate cancer. Int. J. Cancer 2015, 137, 2287–2295. [Google Scholar] [CrossRef]
  155. Samuel, C.S.; Hewitson, T.D. Relaxin and the progression of kidney disease. Curr. Opin. Nephrol. Hypertens. 2009, 18, 9–14. [Google Scholar] [CrossRef]
  156. Samuel, C.S.; Zhao, C.; Bond, C.P.; Hewitson, T.D.; Amento, E.P.; Summers, R.J. Relaxin-1-deficient mice develop an age-related progression of renal fibrosis. Kidney Int. 2004, 65, 2054–2064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Schafer, E.J.; Jemal, A.; Wiese, D.; Sung, H.; Kratzer, T.B.; Islami, F.; Dahut, W.L.; Knudsen, K.E. Disparities and Trends in Genitourinary Cancer Incidence and Mortality in the USA. Eur. Urol. 2022, 84, 117–126. [Google Scholar] [CrossRef]
  158. Laor, E.; Schiffman, Z.J.; Braunstein, J.D.; Reid, R.E.; Tolia, B.M.; Koss, L.G.; Freed, S.Z. Androgen receptors in bladder tumors. Urology 1985, 25, 161–163. [Google Scholar] [CrossRef] [PubMed]
  159. Lin, C.; Yin, Y.; Stemler, K.; Humphrey, P.; Kibel, A.S.; Mysorekar, I.U.; Ma, L. Constitutive beta-catenin activation induces male-specific tumorigenesis in the bladder urothelium. Cancer Res. 2013, 73, 5914–5925. [Google Scholar] [CrossRef] [Green Version]
  160. Terada, S.; Suzuki, N.; Uchide, K.; Akasofu, K.; Nishida, E. Effect of testosterone on the development of bladder tumors and calculi in female rats. Gynecol. Obstet. Investig. 1992, 34, 105–110. [Google Scholar] [CrossRef]
  161. Hsu, J.W.; Hsu, I.; Xu, D.; Miyamoto, H.; Liang, L.; Wu, X.R.; Shyr, C.R.; Chang, C. Decreased tumorigenesis and mortality from bladder cancer in mice lacking urothelial androgen receptor. Am. J. Pathol. 2013, 182, 1811–1820. [Google Scholar]
  162. Wang, C.S.; Li, C.C.; Juan, Y.S.; Wu, W.J.; Lee, H.Y. 5alpha-reductase inhibitors impact prognosis of urothelial carcinoma. BMC Cancer 2020, 20, 872. [Google Scholar] [CrossRef]
  163. Katleba, K.; Lombard, A.P.; Tsamouri, M.M.; Baek, H.B.; Nishida, K.S.; Libertini, S.J.; Platero, A.J.; Ma, A.H.; Pan, C.X.; Ghosh, P.M.; et al. Depletion of androgen receptor low molecular weight isoform reduces bladder tumor cell viability and induces apoptosis. Cancer Lett. 2021, 504, 49–57. [Google Scholar] [CrossRef] [PubMed]
  164. Katleba, K.D.; Tsamouri, M.M.; Jathal, M.; Baek, H.B.; Armenta, R.B.; Tepper, C.G.; Cortopassi, G.; Ghosh, P.M.; Mudryj, M. Androgen receptor-dependent regulation of metabolism in high grade bladder cancer cells. Sci. Rep. 2023, 13, 1762. [Google Scholar] [CrossRef] [PubMed]
  165. Ma, W.L.; Lai, H.C.; Yeh, S.; Cai, X.; Chang, C. Androgen receptor roles in hepatocellular carcinoma, fatty liver, cirrhosis and hepatitis. Endocr. Relat. Cancer 2014, 21, R165–R182. [Google Scholar] [CrossRef] [Green Version]
  166. Kanda, T.; Jiang, X.; Yokosuka, O. Androgen receptor signaling in hepatocellular carcinoma and pancreatic cancers. World J. Gastroenterol. 2014, 20, 9229–9236. [Google Scholar] [PubMed]
  167. Montgomery, E.J.; Xing, E.; Campbell, M.J.; Li, P.K.; Blachly, J.S.; Tsung, A.; Coss, C.C. Constitutively Active Androgen Receptor in Hepatocellular Carcinoma. Int. J. Mol. Sci. 2022, 23, 13768. [Google Scholar] [CrossRef]
  168. Ma, W.L.; Hsu, C.L.; Wu, M.H.; Wu, C.T.; Wu, C.C.; Lai, J.J.; Jou, Y.S.; Chen, C.W.; Yeh, S.; Chang, C. Androgen receptor is a new potential therapeutic target for the treatment of hepatocellular carcinoma. Gastroenterology 2008, 135, 947–955.e5. [Google Scholar]
  169. Wu, M.H.; Ma, W.L.; Hsu, C.L.; Chen, Y.L.; Ou, J.H.; Ryan, C.K.; Hung, Y.C.; Yeh, S.; Chang, C. Androgen receptor promotes hepatitis B virus-induced hepatocarcinogenesis through modulation of hepatitis B virus RNA transcription. Sci. Transl. Med. 2010, 2, 32ra35. [Google Scholar] [CrossRef] [Green Version]
  170. Liu, W.C.; Liu, Q.Y. Molecular mechanisms of gender disparity in hepatitis B virus-associated hepatocellular carcinoma. World J. Gastroenterol. 2014, 20, 6252–6261. [Google Scholar] [CrossRef] [PubMed]
  171. Wang, S.H.; Chen, P.J.; Yeh, S.H. Gender disparity in chronic hepatitis B: Mechanisms of sex hormones. J. Gastroenterol. Hepatol. 2015, 30, 1237–1245. [Google Scholar] [CrossRef] [Green Version]
  172. Kalra, M.; Mayes, J.; Assefa, S.; Kaul, A.K.; Kaul, R. Role of sex steroid receptors in pathobiology of hepatocellular carcinoma. World J. Gastroenterol. 2008, 14, 5945–5961. [Google Scholar]
  173. Zhang, H.; Li, X.X.; Yang, Y.; Zhang, Y.; Wang, H.Y.; Zheng, X.F.S. Significance and mechanism of androgen receptor overexpression and androgen receptor/mechanistic target of rapamycin cross-talk in hepatocellular carcinoma. Hepatology 2018, 67, 2271–2286. [Google Scholar] [CrossRef] [PubMed]
  174. Acosta-Lopez, S.; Diaz-Bethencourt, D.; Concepcion-Massip, T.; de Basoa, M.C.M.-F.; Plata-Bello, A.; Gonzalez-Rodriguez, A.; Perez-Hernandez, F.; Plata-Bello, J. The androgen receptor expression and its activity have different relationships with prognosis in hepatocellular carcinoma. Sci. Rep. 2020, 10, 22046. [Google Scholar] [CrossRef] [PubMed]
  175. Rajaram, P.; Rivera, A.; Muthima, K.; Olveda, N.; Muchalski, H.; Chen, Q.H. Second-Generation Androgen Receptor Antagonists as Hormonal Therapeutics for Three Forms of Prostate Cancer. Molecules 2020, 25, 2448. [Google Scholar] [CrossRef]
  176. Grimaldi, C.; Bleiberg, H.; Gay, F.; Messner, M.; Rougier, P.; Kok, T.C.; Cirera, L.; Cervantes, A.; De Greve, J.; Paillot, B.; et al. Evaluation of antiandrogen therapy in unresectable hepatocellular carcinoma: Results of a European Organization for Research and Treatment of Cancer multicentric double-blind trial. J. Clin. Oncol. 1998, 16, 411–417. [Google Scholar]
  177. Dauki, A.M.; Blachly, J.S.; Kautto, E.A.; Ezzat, S.; Abdel-Rahman, M.H.; Coss, C.C. Transcriptionally Active Androgen Receptor Splice Variants Promote Hepatocellular Carcinoma Progression. Cancer Res. 2020, 80, 561–575. [Google Scholar] [CrossRef]
  178. Lee, S.E.; Alcedo, K.P.; Kim, H.J.; Snider, N.T. Alternative Splicing in Hepatocellular Carcinoma. Cell Mol. Gastroenterol. Hepatol. 2020, 10, 699–712. [Google Scholar] [CrossRef]
  179. Wang, F.; Pan, J.; Liu, Y.; Meng, Q.; Lv, P.; Qu, F.; Ding, G.L.; Klausen, C.; Leung, P.C.; Chan, H.C.; et al. Alternative splicing of the androgen receptor in polycystic ovary syndrome. Proc. Natl. Acad. Sci. USA 2015, 112, 4743–4748. [Google Scholar] [CrossRef]
  180. Marin-Aguilera, M.; Jimenez, N.; Reig, O.; Montalbo, R.; Verma, A.K.; Castellano, G.; Mengual, L.; Victoria, I.; Pereira, M.V.; Mila-Guasch, M.; et al. Androgen Receptor and Its Splicing Variant 7 Expression in Peripheral Blood Mononuclear Cells and in Circulating Tumor Cells in Metastatic Castration-Resistant Prostate Cancer. Cells 2020, 9, 203. [Google Scholar] [CrossRef] [Green Version]
  181. Meakin, A.S.; Saif, Z.; Tuck, A.R.; Clifton, V.L. Human placental androgen receptor variants: Potential regulators of male fetal growth. Placenta 2019, 80, 18–26. [Google Scholar] [CrossRef]
  182. Meakin, A.S.; Morrison, J.L.; Bradshaw, E.L.; Holman, S.L.; Saif, Z.; Gatford, K.L.; Wallace, M.J.; Bischof, R.J.; Moss, T.J.M.; Clifton, V.L. Identification of placental androgen receptor isoforms in a sheep model of maternal allergic asthma. Placenta 2021, 104, 232–235. [Google Scholar] [CrossRef] [PubMed]
  183. Laurentino, S.S.; Pinto, P.I.; Tomas, J.; Cavaco, J.E.; Sousa, M.; Barros, A.; Power, D.M.; Canario, A.V.; Socorro, S. Identification of androgen receptor variants in testis from humans and other vertebrates. Andrologia 2013, 45, 187–194. [Google Scholar] [CrossRef] [PubMed]
  184. Tian, F.; Wu, Y.S.; Zhao, J.; Li, W. AR3 messenger ribonucleic acid expression and its functional implication in human primary testicular failure. Andrologia 2014, 46, 859–866. [Google Scholar] [CrossRef] [PubMed]
  185. Garza-Contreras, J.; Duong, P.; Snyder, B.D.; Schreihofer, D.A.; Cunningham, R.L. Presence of Androgen Receptor Variant in Neuronal Lipid Rafts. eNeuro 2017, 4, ENEURO.0109-17.2017. [Google Scholar] [CrossRef] [Green Version]
  186. Sanchez-Hernandez, M.; Arizcun, M.; Garcia-Alcazar, A.; Sarropoulou, E.; Mulero, V.; Garcia-Ayala, A. Fish granulocytes express a constitutively active androgen receptor variant. Dev. Comp. Immunol. 2014, 45, 115–122. [Google Scholar] [CrossRef]
  187. Ben-Batalla, I.; Vargas-Delgado, M.E.; von Amsberg, G.; Janning, M.; Loges, S. Influence of Androgens on Immunity to Self and Foreign: Effects on Immunity and Cancer. Front. Immunol. 2020, 11, 1184. [Google Scholar] [CrossRef]
  188. Dodd, K.C.; Menon, M. Sex bias in lymphocytes: Implications for autoimmune diseases. Front. Immunol. 2022, 13, 945762. [Google Scholar] [CrossRef]
  189. Kronzer, V.L.; Bridges, S.L., Jr.; Davis, J.M., 3rd. Why women have more autoimmune diseases than men: An evolutionary perspective. Evol. Appl. 2021, 14, 629–633. [Google Scholar] [CrossRef]
  190. Lee, T.P.; Chiang, B.L. Sex differences in spontaneous versus induced animal models of autoimmunity. Autoimmun. Rev. 2012, 11, A422–A429. [Google Scholar] [CrossRef]
  191. Brown, M.A.; Su, M.A. An Inconvenient Variable: Sex Hormones and Their Impact on T Cell Responses. J. Immunol. 2019, 202, 1927–1933. [Google Scholar] [CrossRef] [Green Version]
  192. Bereshchenko, O.; Bruscoli, S.; Riccardi, C. Glucocorticoids, Sex Hormones and Immunity. Front. Immunol. 2018, 9, 1332. [Google Scholar] [CrossRef] [Green Version]
  193. Fitzpatrick, F.; Lepault, F.; Homo-Delarche, F.; Bach, J.F.; Dardenne, M. Influence of castration, alone or combined with thymectomy, on the development of diabetes in the nonobese diabetic mouse. Endocrinology 1991, 129, 1382–1390. [Google Scholar] [CrossRef] [PubMed]
  194. Fox, H.S. Androgen treatment prevents diabetes in nonobese diabetic mice. J. Exp. Med. 1992, 175, 1409–1412. [Google Scholar] [CrossRef] [PubMed]
  195. Liva, S.M.; Voskuhl, R.R. Testosterone acts directly on CD4+ T lymphocytes to increase IL-10 production. J. Immunol. 2001, 167, 2060–2067. [Google Scholar] [CrossRef]
  196. Qu, F.; Xie, W.; Nakabayashi, M.; Zhang, H.; Jeong, S.H.; Wang, X.; Komura, K.; Sweeney, C.J.; Sartor, O.; Lee, G.M.; et al. Association of AR-V7 and Prostate-Specific Antigen RNA Levels in Blood with Efficacy of Abiraterone Acetate and Enzalutamide Treatment in Men with Prostate Cancer. Clin. Cancer Res. 2017, 23, 726–734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Dapas, M.; Dunaif, A. Deconstructing a Syndrome: Genomic Insights into PCOS Causal Mechanisms and Classification. Endocr. Rev. 2022, 43, 927–965. [Google Scholar] [CrossRef]
  198. Conway, G.; Dewailly, D.; Diamanti-Kandarakis, E.; Escobar-Morreale, H.F.; Franks, S.; Gambineri, A.; Kelestimur, F.; Macut, D.; Micic, D.; Pasquali, R.; et al. The polycystic ovary syndrome: A position statement from the European Society of Endocrinology. Eur. J. Endocrinol. 2014, 171, P1–P29. [Google Scholar]
  199. Norman, R.J.; Dewailly, D.; Legro, R.S.; Hickey, T.E. Polycystic ovary syndrome. Lancet 2007, 370, 685–697. [Google Scholar] [CrossRef] [Green Version]
  200. Walters, K.A. Role of androgens in normal and pathological ovarian function. Reproduction 2015, 149, R193–R218. [Google Scholar] [CrossRef] [Green Version]
  201. Walters, K.A.; Handelsman, D.J. Androgen receptor splice variants and polycystic ovary syndrome: Cause or effect? Asian J. Androl. 2016, 18, 442–443. [Google Scholar] [CrossRef] [PubMed]
  202. Hu, Y.C.; Wang, P.H.; Yeh, S.; Wang, R.S.; Xie, C.; Xu, Q.; Zhou, X.; Chao, H.T.; Tsai, M.Y.; Chang, C. Subfertility and defective folliculogenesis in female mice lacking androgen receptor. Proc. Natl. Acad. Sci. USA 2004, 101, 11209–11214. [Google Scholar] [CrossRef] [PubMed]
  203. Sen, A.; Hammes, S.R. Granulosa cell-specific androgen receptors are critical regulators of ovarian development and function. Mol. Endocrinol. 2010, 24, 1393–1403. [Google Scholar] [CrossRef] [PubMed]
  204. Shah, N.A.; Antoine, H.J.; Pall, M.; Taylor, K.D.; Azziz, R.; Goodarzi, M.O. Association of androgen receptor CAG repeat polymorphism and polycystic ovary syndrome. J. Clin. Endocrinol. Metab. 2008, 93, 1939–1945. [Google Scholar] [CrossRef] [Green Version]
  205. Peng, C.Y.; Long, X.Y.; Lu, G.X. Association of AR rs6152G/A gene polymorphism with susceptibility to polycystic ovary syndrome in Chinese women. Reprod. Fertil. Dev. 2010, 22, 881–885. [Google Scholar] [CrossRef]
  206. Escobar, J.C.; Patel, S.S.; Beshay, V.E.; Suzuki, T.; Carr, B.R. The human placenta expresses CYP17 and generates androgens de novo. J. Clin. Endocrinol. Metab. 2011, 96, 1385–1392. [Google Scholar] [CrossRef] [Green Version]
  207. Sathishkumar, K.; Balakrishnan, M.; Chinnathambi, V.; Chauhan, M.; Hankins, G.D.; Yallampalli, C. Fetal sex-related dysregulation in testosterone production and their receptor expression in the human placenta with preeclampsia. J. Perinatol. 2012, 32, 328–335. [Google Scholar] [CrossRef] [Green Version]
  208. Fornes, R.; Maliqueo, M.; Hu, M.; Hadi, L.; Jimenez-Andrade, J.M.; Ebefors, K.; Nystrom, J.; Labrie, F.; Jansson, T.; Benrick, A.; et al. The effect of androgen excess on maternal metabolism, placental function and fetal growth in obese dams. Sci. Rep. 2017, 7, 8066. [Google Scholar] [CrossRef] [Green Version]
  209. Cleys, E.R.; Halleran, J.L.; Enriquez, V.A.; da Silveira, J.C.; West, R.C.; Winger, Q.A.; Anthony, R.V.; Bruemmer, J.E.; Clay, C.M.; Bouma, G.J. Androgen receptor and histone lysine demethylases in ovine placenta. PLoS ONE 2015, 10, e0117472. [Google Scholar] [CrossRef] [Green Version]
  210. Khatri, P.; Hoffmann, B.; Schuler, G. Androgen receptor is widely expressed in bovine placentomes and up-regulated during differentiation of bovine trophoblast giant cells. Placenta 2013, 34, 416–423. [Google Scholar] [CrossRef]
  211. Dobbins, T.A.; Sullivan, E.A.; Roberts, C.L.; Simpson, J.M. Australian national birthweight percentiles by sex and gestational age, 1998–2007. Med. J. Aust. 2012, 197, 291–294. [Google Scholar] [CrossRef] [Green Version]
  212. Kumar, S.; Gordon, G.H.; Abbott, D.H.; Mishra, J.S. Androgens in maternal vascular and placental function: Implications for preeclampsia pathogenesis. Reproduction 2018, 156, R155–R167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. McWhorter, E.S.; West, R.C.; Russ, J.E.; Ali, A.; Winger, Q.A.; Bouma, G.J. LIN28B regulates androgen receptor in human trophoblast cells through Let-7c. Mol. Reprod. Dev. 2019, 86, 1086–1093. [Google Scholar] [CrossRef]
  214. Zhou, X. Roles of androgen receptor in male and female reproduction: Lessons from global and cell-specific androgen receptor knockout (ARKO) mice. J. Androl. 2010, 31, 235–243. [Google Scholar]
  215. Cunha, G.R.; Cao, M.; Aksel, S.; Derpinghaus, A.; Baskin, L.S. Mouse-human species differences in early testicular development and its implications. Differentiation 2023, 129, 79–95. [Google Scholar] [CrossRef] [PubMed]
  216. Wang, J.M.; Li, Z.F.; Yang, W.X. What Does Androgen Receptor Signaling Pathway in Sertoli Cells During Normal Spermatogenesis Tell Us? Front. Endocrinol. 2022, 13, 838858. [Google Scholar] [CrossRef] [PubMed]
  217. Elliott, D.J.; Grellscheid, S.N. Alternative RNA splicing regulation in the testis. Reproduction 2006, 132, 811–819. [Google Scholar] [CrossRef]
  218. Ogawa, S.; Inoue, S.; Watanabe, T.; Orimo, A.; Hosoi, T.; Ouchi, Y.; Muramatsu, M. Molecular cloning and characterization of human estrogen receptor betacx: A potential inhibitor ofestrogen action in human. Nucleic Acids Res. 1998, 26, 3505–3512. [Google Scholar] [CrossRef] [Green Version]
  219. Inoue, S.; Ogawa, S.; Horie, K.; Hoshino, S.; Goto, W.; Hosoi, T.; Tsutsumi, O.; Muramatsu, M.; Ouchi, Y. An estrogen receptor beta isoform that lacks exon 5 has dominant negative activity on both ERalpha and ERbeta. Biochem. Biophys. Res. Commun. 2000, 279, 814–819. [Google Scholar] [CrossRef]
  220. Hirata, S.; Shoda, T.; Kato, J.; Hoshi, K. Novel isoforms of the mRNA for human female sex steroid hormone receptors. J. Steroid Biochem. Mol. Biol. 2002, 83, 25–30. [Google Scholar] [CrossRef]
  221. Saunders, P.T.; Millar, M.R.; Macpherson, S.; Irvine, D.S.; Groome, N.P.; Evans, L.R.; Sharpe, R.M.; Scobie, G.A. ERbeta1 and the ERbeta2 splice variant (ERbetacx/beta2) are expressed in distinct cell populations in the adult human testis. J. Clin. Endocrinol. Metab. 2002, 87, 2706–2715. [Google Scholar]
  222. Scobie, G.A.; Macpherson, S.; Millar, M.R.; Groome, N.P.; Romana, P.G.; Saunders, P.T. Human oestrogen receptors: Differential expression of ER alpha and beta and the identification of ER beta variants. Steroids 2002, 67, 985–992. [Google Scholar] [CrossRef] [PubMed]
  223. Shoda, T.; Hirata, S.; Kato, J.; Hoshi, K. Cloning of the novel isoform of the estrogen receptor beta cDNA (ERbeta isoform M cDNA) from the human testicular cDNA library. J. Steroid Biochem. Mol. Biol. 2002, 82, 201–208. [Google Scholar] [CrossRef]
  224. Aschim, E.L.; Saether, T.; Wiger, R.; Grotmol, T.; Haugen, T.B. Differential distribution of splice variants of estrogen receptor beta in human testicular cells suggests specific functions in spermatogenesis. J. Steroid Biochem. Mol. Biol. 2004, 92, 97–106. [Google Scholar] [CrossRef]
  225. 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]
  226. Zhu, Y.; Dalrymple, S.L.; Coleman, I.; Zheng, S.L.; Xu, J.; Hooper, J.E.; Antonarakis, E.S.; De Marzo, A.M.; Meeker, A.K.; Nelson, P.S.; et al. Role of androgen receptor splice variant-7 (AR-V7) in prostate cancer resistance to 2nd-generation androgen receptor signaling inhibitors. Oncogene 2020, 39, 6935–6949. [Google Scholar] [CrossRef]
  227. Quigley, C.A.; Evans, B.A.; Simental, J.A.; Marschke, K.B.; Sar, M.; Lubahn, D.B.; Davies, P.; Hughes, I.A.; Wilson, E.M.; French, F.S. Complete androgen insensitivity due to deletion of exon C of the androgen receptor gene highlights the functional importance of the second zinc finger of the androgen receptor in vivo. Mol. Endocrinol. 1992, 6, 1103–1112. [Google Scholar]
  228. Kritzer, M.F. Selective colocalization of immunoreactivity for intracellular gonadal hormone receptors and tyrosine hydroxylase in the ventral tegmental area, substantia nigra, and retrorubral fields in the rat. J. Comp. Neurol. 1997, 379, 247–260. [Google Scholar] [CrossRef]
  229. Hatzoglou, A.; Kampa, M.; Kogia, C.; Charalampopoulos, I.; Theodoropoulos, P.A.; Anezinis, P.; Dambaki, C.; Papakonstanti, E.A.; Stathopoulos, E.N.; Stournaras, C.; et al. Membrane androgen receptor activation induces apoptotic regression of human prostate cancer cells in vitro and in vivo. J. Clin. Endocrinol. Metab. 2005, 90, 893–903. [Google Scholar] [CrossRef] [Green Version]
  230. Aguila, S.; Castillo-Briceno, P.; Sanchez, M.; Cabas, I.; Garcia-Alcazar, A.; Meseguer, J.; Mulero, V.; Garcia-Ayala, A. Specific and non-overlapping functions of testosterone and 11-ketotestosterone in the regulation of professional phagocyte responses in the teleost fish gilthead seabream. Mol. Immunol. 2013, 53, 218–226. [Google Scholar] [CrossRef] [PubMed]
  231. Kritzer, M.F.; Creutz, L.M. Region and sex differences in constituent dopamine neurons and immunoreactivity for intracellular estrogen and androgen receptors in mesocortical projections in rats. J. Neurosci. 2008, 28, 9525–9535. [Google Scholar] [CrossRef] [Green Version]
  232. Sarkey, S.; Azcoitia, I.; Garcia-Segura, L.M.; Garcia-Ovejero, D.; DonCarlos, L.L. Classical androgen receptors in non-classical sites in the brain. Horm. Behav. 2008, 53, 753–764. [Google Scholar] [CrossRef] [Green Version]
  233. Papakonstanti, E.A.; Kampa, M.; Castanas, E.; Stournaras, C. A rapid, nongenomic, signaling pathway regulates the actin reorganization induced by activation of membrane testosterone receptors. Mol. Endocrinol. 2003, 17, 870–881. [Google Scholar] [CrossRef] [Green Version]
  234. Gorczynska, E.; Handelsman, D.J. Androgens rapidly increase the cytosolic calcium concentration in Sertoli cells. Endocrinology 1995, 136, 2052–2059. [Google Scholar] [CrossRef]
  235. Benten, W.P.; Lieberherr, M.; Stamm, O.; Wrehlke, C.; Guo, Z.; Wunderlich, F. Testosterone signaling through internalizable surface receptors in androgen receptor-free macrophages. Mol. Biol. Cell 1999, 10, 3113–3123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Estrada, M.; Espinosa, A.; Muller, M.; Jaimovich, E. Testosterone stimulates intracellular calcium release and mitogen-activated protein kinases via a G protein-coupled receptor in skeletal muscle cells. Endocrinology 2003, 144, 3586–3597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Estrada, M.; Uhlen, P.; Ehrlich, B.E. Ca2+ oscillations induced by testosterone enhance neurite outgrowth. J. Cell Sci. 2006, 119, 733–743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Holmes, S.; Abbassi, B.; Su, C.; Singh, M.; Cunningham, R.L. Oxidative stress defines the neuroprotective or neurotoxic properties of androgens in immortalized female rat dopaminergic neuronal cells. Endocrinology 2013, 154, 4281–4292. [Google Scholar] [CrossRef] [Green Version]
  239. Steinsapir, J.; Socci, R.; Reinach, P. Effects of androgen on intracellular calcium of LNCaP cells. Biochem. Biophys. Res. Commun. 1991, 179, 90–96. [Google Scholar] [CrossRef]
  240. Weiss, B.; Faus, H.; Haendler, B. Phylogenetic conservation of the androgen receptor AR45 variant form in placental mammals. Gene 2007, 399, 105–111. [Google Scholar] [CrossRef]
  241. Wu, Z.Y.; Chen, K.; Haendler, B.; McDonald, T.V.; Bian, J.S. Stimulation of N-terminal truncated isoform of androgen receptor stabilizes human ether-a-go-go-related gene-encoded potassium channel protein via activation of extracellular signal regulated kinase 1/2. Endocrinology 2008, 149, 5061–5069. [Google Scholar] [CrossRef] [PubMed]
  242. Schuppe, E.R.; Miles, M.C.; Fuxjager, M.J. Evolution of the androgen receptor: Perspectives from human health to dancing birds. Mol. Cell Endocrinol. 2020, 499, 110577. [Google Scholar] [CrossRef] [PubMed]
  243. Blazquez, M.; Piferrer, F. Sea bass (Dicentrarchus labrax) androgen receptor: cDNA cloning, tissue-specific expression and mRNA levels during early development and sex differentiation. Mol. Cell Endocrinol. 2005, 237, 37–48. [Google Scholar] [CrossRef]
  244. de Waal, P.P.; Wang, D.S.; Nijenhuis, W.A.; Schulz, R.W.; Bogerd, J. Functional characterization and expression analysis of the androgen receptor in zebrafish (Danio rerio) testis. Reproduction 2008, 136, 225–234. [Google Scholar] [CrossRef] [Green Version]
  245. Chaves-Pozo, E.; Mulero, V.; Meseguer, J.; Ayala, A.G. An overview of cell renewal in the testis throughout the reproductive cycle of a seasonal breeding teleost, the gilthead seabream (Sparus aurata L.). Biol. Reprod. 2005, 72, 593–601. [Google Scholar] [CrossRef] [Green Version]
  246. Liarte, S.; Chaves-Pozo, E.; Garcia-Alcazar, A.; Mulero, V.; Meseguer, J.; Garcia-Ayala, A. Testicular involution prior to sex change in gilthead seabream is characterized by a decrease in DMRT1 gene expression and by massive leukocyte infiltration. Reprod. Biol. Endocrinol. 2007, 5, 20. [Google Scholar] [CrossRef] [Green Version]
  247. Chaves-Pozo, E.; Pelegrin, P.; Mulero, V.; Meseguer, J.; Ayala, A.G. A role for acidophilic granulocytes in the testis of the gilthead seabream (Sparus aurata L., Teleostei). J. Endocrinol. 2003, 179, 165–174. [Google Scholar] [CrossRef] [Green Version]
  248. Chaves-Pozo, E.; Mulero, V.; Meseguer, J.; Ayala, A.G. Professional phagocytic granulocytes of the bony fish gilthead seabream display functional adaptation to testicular microenvironment. J. Leukoc. Biol. 2005, 78, 345–351. [Google Scholar] [CrossRef] [Green Version]
  249. Lin, P.Y.; Sun, L.; Thibodeaux, S.R.; Ludwig, S.M.; Vadlamudi, R.K.; Hurez, V.J.; Bahar, R.; Kious, M.J.; Livi, C.B.; Wall, S.R.; et al. B7-H1-dependent sex-related differences in tumor immunity and immunotherapy responses. J. Immunol. 2010, 185, 2747–2753. [Google Scholar] [CrossRef] [Green Version]
  250. Conforti, F.; Pala, L.; Bagnardi, V.; De Pas, T.; Martinetti, M.; Viale, G.; Gelber, R.D.; Goldhirsch, A. Cancer immunotherapy efficacy and patients’ sex: A systematic review and meta-analysis. Lancet Oncol. 2018, 19, 737–746. [Google Scholar] [PubMed]
  251. Wang, S.; Cowley, L.A.; Liu, X.S. Sex Differences in Cancer Immunotherapy Efficacy, Biomarkers and Therapeutic Strategy. Molecules 2019, 24, 3214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Conforti, F.; Pala, L.; Bagnardi, V.; Viale, G.; De Pas, T.; Pagan, E.; Pennacchioli, E.; Cocorocchio, E.; Ferrucci, P.F.; De Marinis, F.; et al. Sex-Based Heterogeneity in Response to Lung Cancer Immunotherapy: A Systematic Review and Meta-Analysis. J. Natl. Cancer Inst. 2019, 111, 772–781. [Google Scholar] [CrossRef]
  253. Gamat, M.; McNeel, D.G. Androgen deprivation and immunotherapy for the treatment of prostate cancer. Endocr. Relat. Cancer 2017, 24, T297–T310. [Google Scholar] [CrossRef] [PubMed]
  254. Antonarakis, E.S.; Piulats, J.M.; Gross-Goupil, M.; Goh, J.; Ojamaa, K.; Hoimes, C.J.; Vaishampayan, U.; Berger, R.; Sezer, A.; Alanko, T.; et al. Pembrolizumab for Treatment-Refractory Metastatic Castration-Resistant Prostate Cancer: Multicohort, Open-Label Phase II KEYNOTE-199 Study. J. Clin. Oncol. 2020, 38, 395–405. [Google Scholar]
  255. Bishop, J.L.; Sio, A.; Angeles, A.; Roberts, M.E.; Azad, A.A.; Chi, K.N.; Zoubeidi, A. PD-L1 is highly expressed in Enzalutamide resistant prostate cancer. Oncotarget 2015, 6, 234–242. [Google Scholar] [CrossRef] [Green Version]
  256. Mercader, M.; Bodner, B.K.; Moser, M.T.; Kwon, P.S.; Park, E.S.; Manecke, R.G.; Ellis, T.M.; Wojcik, E.M.; Yang, D.; Flanigan, R.C.; et al. T cell infiltration of the prostate induced by androgen withdrawal in patients with prostate cancer. Proc. Natl. Acad. Sci. USA 2001, 98, 14565–14570. [Google Scholar] [CrossRef] [PubMed]
  257. Kissick, H.T.; Sanda, M.G.; Dunn, L.K.; Pellegrini, K.L.; On, S.T.; Noel, J.K.; Arredouani, M.S. Androgens alter T-cell immunity by inhibiting T-helper 1 differentiation. Proc. Natl. Acad. Sci. USA 2014, 111, 9887–9892. [Google Scholar] [CrossRef]
  258. Graff, J.N.; Alumkal, J.J.; Drake, C.G.; Thomas, G.V.; Redmond, W.L.; Farhad, M.; Cetnar, J.P.; Ey, F.S.; Bergan, R.C.; Slottke, R.; et al. Early evidence of anti-PD-1 activity in enzalutamide-resistant prostate cancer. Oncotarget 2016, 7, 52810–52817. [Google Scholar] [CrossRef] [Green Version]
  259. Graff, J.N.; Beer, T.M.; Alumkal, J.J.; Slottke, R.E.; Redmond, W.L.; Thomas, G.V.; Thompson, R.F.; Wood, M.A.; Koguchi, Y.; Chen, Y.; et al. A phase II single-arm study of pembrolizumab with enzalutamide in men with metastatic castration-resistant prostate cancer progressing on enzalutamide alone. J. Immunother. Cancer 2020, 8, e000642. [Google Scholar] [CrossRef]
Figure 1. Schematic of the cryo-EM-derived three-dimensional structure of ligand bound AR binding to AREs, interacting with co-regulators p300 and SRC-3. Note that in the dimerized molecule, the NTDs have slightly different confirmations and that the hinge region is hidden. The NTDs interact with each other and wrap around the LBD.
Figure 1. Schematic of the cryo-EM-derived three-dimensional structure of ligand bound AR binding to AREs, interacting with co-regulators p300 and SRC-3. Note that in the dimerized molecule, the NTDs have slightly different confirmations and that the hinge region is hidden. The NTDs interact with each other and wrap around the LBD.
Biomedicines 11 02215 g001
Figure 2. Schematic of homo and hetero dimers of FL-AR and AR-SV. The structures of the AR-SV homodimer and FL-AR and AR-SV heterodimers are unknown; therefore, they are solely the authors’ representation to denote differences. Since the DBD (green) is retained in FL-AR and AR-SV isoforms, the molecules are tethered to DNA sequences. The absence of the LBD (orange) would alter the intermolecular and intramolecular interactions of the NTD (blue); hence, the structure of the dimer would be different, as would the assembly of AR interactors (represented by the light-colored clouds). The unique complement of AR interactors would regulate AR DNA binding, transcriptional activity, and subsequent distinct transcriptional output.
Figure 2. Schematic of homo and hetero dimers of FL-AR and AR-SV. The structures of the AR-SV homodimer and FL-AR and AR-SV heterodimers are unknown; therefore, they are solely the authors’ representation to denote differences. Since the DBD (green) is retained in FL-AR and AR-SV isoforms, the molecules are tethered to DNA sequences. The absence of the LBD (orange) would alter the intermolecular and intramolecular interactions of the NTD (blue); hence, the structure of the dimer would be different, as would the assembly of AR interactors (represented by the light-colored clouds). The unique complement of AR interactors would regulate AR DNA binding, transcriptional activity, and subsequent distinct transcriptional output.
Biomedicines 11 02215 g002
Figure 3. Schematic of the gene encoding AR, noting canonical and cryptic exons that give rise to the FL-AR and a subset of AR-SVs. The exons coding for the different domains are colored NTD-blue, alternative NTD-red, DBD-orange, hinge-yellow, and LBD-green. The cryptic intronic sequences found in AR-SVs are in shades of purple or black.
Figure 3. Schematic of the gene encoding AR, noting canonical and cryptic exons that give rise to the FL-AR and a subset of AR-SVs. The exons coding for the different domains are colored NTD-blue, alternative NTD-red, DBD-orange, hinge-yellow, and LBD-green. The cryptic intronic sequences found in AR-SVs are in shades of purple or black.
Biomedicines 11 02215 g003
Figure 4. Male-to-female ratio of cancer incidence of the most common cancers (2023) [70].
Figure 4. Male-to-female ratio of cancer incidence of the most common cancers (2023) [70].
Biomedicines 11 02215 g004
Table 1. Summary of AR-SVs detected in various cell types.
Table 1. Summary of AR-SVs detected in various cell types.
SVs DetectedBiological ProcessesKnown/Putative TargetsReferences
Prostate CancerV1–V14 (and others)AR signaling, cell cycle, mTORC1, and OX/PHOSUBE2C, KLK3, EDN2, ETS2, SRD5A1, ORM1 BIRC3, FKBP5, and HES1[37,39,40,42]
Breast CancerV1, V3, V7, V45, V9, V15, V16, V17, and V18immune function, signaling, and cell movementmultiple[109,110,111,112,113]
Salivary Duct CarcinomaV7viability/
proliferation
unknown[128,129]
Glioblastoma MultiformeV7unknownunknown[140]
Renal CarcinomaV1, V3, V4, and V7expression correlates with RLN2unknown[152]
Bladder CancerV1, V7, and V19cell cycle, mTOR, OXPHOS, ribosome, mitochondria, and HIF1amTOR and FKBP5[163,164]
Liver CancerV1, V3, and V7migration and invasion; mTOR signalingSNA12[177]
Polycystic Ovarian Syndrome69 bp ins between exons 2 and 3, exon 3 deletionlimits conversion of androgens to estrogen; fails to upregulate ratios of estradiol to total testosteroneCYP17A1[188]
Peripheral Blood Mononuclear CellsV7response to AA/ENZ and taxanes [229]
PlacentaV1, V7 and AR45unknownunknown[196,199]
TestisAR∆2Stop, AR∆223Stop, AR∆3, and AR∆4spermatogenesisunknown[211,214]
Neuronal Lipid RaftsAR45interacts with GPCR proteins Gαq and Gαounknown[226]
Fish Granulocytes (functional equivalent of mammalian neutrophils)AR∆LBDimmunocompetance [230]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Katleba, K.D.; Ghosh, P.M.; Mudryj, M. Beyond Prostate Cancer: An Androgen Receptor Splice Variant Expression in Multiple Malignancies, Non-Cancer Pathologies, and Development. Biomedicines 2023, 11, 2215. https://doi.org/10.3390/biomedicines11082215

AMA Style

Katleba KD, Ghosh PM, Mudryj M. Beyond Prostate Cancer: An Androgen Receptor Splice Variant Expression in Multiple Malignancies, Non-Cancer Pathologies, and Development. Biomedicines. 2023; 11(8):2215. https://doi.org/10.3390/biomedicines11082215

Chicago/Turabian Style

Katleba, Kimberley D., Paramita M. Ghosh, and Maria Mudryj. 2023. "Beyond Prostate Cancer: An Androgen Receptor Splice Variant Expression in Multiple Malignancies, Non-Cancer Pathologies, and Development" Biomedicines 11, no. 8: 2215. https://doi.org/10.3390/biomedicines11082215

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