Next Article in Journal
Prolonged Dystocic Labor in Neuraxial Analgesia and the Role of Enkephalin Neurotransmitters: An Experimental Study
Next Article in Special Issue
KDM5D Histone Demethylase Identifies Platinum-Tolerant Head and Neck Cancer Cells Vulnerable to Mitotic Catastrophe
Previous Article in Journal
Pathophysiological Aspects of Muscle Atrophy and Osteopenia Induced by Chronic Constriction Injury (CCI) of the Sciatic Nerve in Rats
Previous Article in Special Issue
DNA Methylation as a Diagnostic, Prognostic, and Predictive Biomarker in Head and Neck Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

‘Toxic Masculinity’: What Is Known about the Role of Androgen Receptors in Head and Neck Squamous Cell Carcinoma

Laboratory for Hereditary Cancer, Division of Molecular Medicine, Ruđer Bošković Institute, HR-10000 Zagreb, Croatia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(4), 3766; https://doi.org/10.3390/ijms24043766
Submission received: 22 December 2022 / Revised: 3 February 2023 / Accepted: 9 February 2023 / Published: 13 February 2023
(This article belongs to the Special Issue Molecular Mechanisms and Pathophysiology of Head and Neck Diseases)

Abstract

:
Head and neck squamous cell carcinoma (HNSCC), the most prevalent cancer in the head and neck region, develops from the mucosal epithelium of the upper aerodigestive tract. Its development directly correlates with alcohol and/or tobacco consumption and infection with human papillomavirus. Interestingly, the relative risk for HNSCC is up to five times higher in males, so it is considered that the endocrine microenvironment is another risk factor. A gender-specific risk for HNSCC suggests either the existence of specific risk factors that affect only males or that females have defensive hormonal and metabolic features. In this review, we summarized the current knowledge about the role of both nuclear and membrane androgen receptors (nAR and mARs, respectively) in HNSCC. As expected, the significance of nAR is much better known; it was shown that increased nAR expression was observed in HNSCC, while treatment with dihydrotestosterone increased proliferation, migration, and invasion of HNSCC cells. For only three out of five currently known mARs—TRPM8, CaV1.2, and OXER1—it was shown either their increased expression in various types of HNSCC or that their increased activity enhanced the migration and invasion of HNSCC cells. The primary treatments for HNSCC are surgery and radiotherapy, but targeted immunotherapies are on the rise. On the other hand, given the evidence of elevated nAR expression in HNSCC, this receptor represents a potential target for antiandrogen therapy. Moreover, there is still plenty of room for further examination of mARs’ role in HNSCC diagnosis, prognosis, and treatment.

1. Introduction

1.1. Head and Neck Squamous Cell Carcinoma

Head and neck squamous cell carcinoma (HNSCC) is the sixth most common cancer worldwide and the most prevalent type of cancer in the head and neck region. In 2020, 879,000 new cases and 445,000 deaths were reported worldwide [1]. The mean age at the diagnosis is 63 years [2] with a five-year survival rate below 50% in the advanced stages of the disease [3]. HNSCC develops from the mucosal epithelial cells of the oral cavity, larynx, nasopharynx, oropharynx, hypopharynx, and sinonasal tract, and its development directly correlates with several risk factors that include alcohol [4] and/or tobacco consumption [5], long-term exposure to environmental pollutants [6], and infection with oncogenic strains of human papillomavirus (HPV) [7] or Epstein–Barr virus (EBV) [8]. Most of the risk factors listed above have a cultural and geographical prevalence due to the different lifestyles. For instance, India’s highest prevalence of oral cavity cancer is connected to tobacco and betel nut chewing and exposure to air pollutants [9].
Furthermore, HPV-positive (HPV+) patients have the highest risk of developing tonsillar (TSCC) and oropharyngeal squamous cell carcinomas (OPSCC), in which the HPV-16 subtype is found in almost 90% of OPSCCs followed by HPV-18 (3%) [10]. On the other hand, laryngeal and oral cavity carcinomas are called HPV-negative (HPV-) because they are associated with smoking and other non-HPV risk factors. As with many other tumors, the formation and progression of HNSCC from the premalignant lesion consists of several steps that include hyperplasia, dysplasia, carcinoma in situ, and invasive carcinoma [11].
After a diagnosis, potential treatments for HNSCC are surgical resection followed by radiotherapy and/or chemotherapy [10]. A lack of screening strategies and consequent disease detection at advanced stages as well as the ineffectiveness of available therapies and treatments leads to a poor prognosis. So far, several molecular biomarkers for clinical diagnosis or prognosis of HNSCC have been approved by the United States Food and Drug Administration (FDA), while only a few of them have proven significant in clinical trials (such as CD44, CD133, and aldehyde dehydrogenase 1 (ALDH1)) [12]. Accordingly, new targeted therapies and predictive, prognostic, and diagnostic biomarkers for the early clinical detection of HNSCC are necessary. Various new biomarkers were evaluated in clinical trials in the last decade; these included hormonal receptors [12,13,14,15].
Interestingly, the relative risk for HNSCC development followed by death is up to five times higher in males compared to females [16]. The endocrine microenvironment is considered to be another risk factor in HNSCC tumorigenesis, although the role of sex hormones in this tumor type is still controversial. However, the gender-specific risk for HNSCC development suggests either the existence of specific risk factors that affect only males or that females have defensive hormonal and metabolic features as a response to common risk factors [17] (Figure 1). In this short but comprehensive review, we present the current knowledge about the role of androgen receptors (ARs) in HNSCC; to the best of our knowledge, this was the first attempt to synthesize what is known about both the nuclear and membrane types of ARs.

1.2. Sex Steroid Hormones and Their Receptors

Many types of tumors depend on steroids for growth, proliferation, and survival. Although steroid hormones control development, reproduction, and metabolism as endocrine molecules, they can also mediate tumor initiation and progression. The well-known hormone-sensitive cancers are breast [18], ovarian [19], prostate [20], thyroid [21], testicular [22], and uterine or endometrial cancer [23], whose progression correlates directly with sex steroid hormones [24]. Steroid hormones are derived from cholesterol, and their synthesis is localized in the mitochondria and smooth endoplasmic reticulum of the adrenal cortex, gonads, and placenta [25]. The synthesis of sex hormones is under the control of the hypothalamic–pituitary–gonadal axis. Secretion of gonadotropin-releasing hormone (GnRH) from the hypothalamus activates the anterior pituitary production of follicle-stimulating hormone (FSH) and luteinizing hormone (LH), which control gametogenesis and steroidogenesis [26]. Cholesterol-derived steroid hormones are divided into two main subgroups: sex steroids (androgens, estrogens, and progestogens) and corticosteroids (glucocorticoids and mineralocorticoids) [27]. The main role of steroid hormones is the regulation of gene expression in the cell nucleus or mediation of the rapid modulation of intracellular molecular pathways. Steroids interact with specific nuclear or membrane receptors to accomplish their role [28]. Membrane steroid receptors are integrated transmembrane proteins activated by steroid binding and therefore can activate or inhibit other proteins and modulate different signaling pathways [29]. On the other hand, nuclear hormone receptors function as transcription factors in the nucleus and receptors on the cell membrane. After post-translational covalent attachment of palmitic acid or association to plasma membrane scaffold proteins, some nuclear hormone receptors can be anchored to the cell membrane and act independently of their nuclear function [30]. Therefore, any signaling dysregulation or unregulated production of sex steroid hormones and/or their receptors can lead to aberrant growth and development as well as the initiation and proliferation of hormone-sensitive cancers [31].

2. Androgen Receptors as a Potential Culprit of Sex-Related Disparities in HNSCC

Several studies have shown that the sex hormone receptors (SHRs) can be prognostic and potential therapeutic targets of the HNSCC because they promote DNA hypermutation and potentiate HPV integration [32]. As already mentioned, due to the high male–female incidence and mortality rate ratios, endocrine homeostasis disruption is one of the potential risk factors for HNSCC development together with well-established risk factors such as alcohol and tobacco consumption. Park et al. recently published a large follow-up study that compared sex differences in HNSCC prevalence in which they showed the most evident differences between the sexes in the incidence of the upper aerodigestive tract tumors, primary laryngeal, and hypopharyngeal [33]. A higher prevalence of laryngeal cancer in males might be connected to hormonal control because, during puberty, the larynx goes through different physiological changes related to sex hormone receptors [33]. Moreover, the destruction of liver function and metabolic processes (including steroidogenesis) related to heavy alcohol consumption is another connection between the steroid receptors and HNSCC prevalence [34]. Therefore, the sex hormone receptors represent a potential therapeutic target for applying hormone receptor modulators and significant predictors of disease and treatment outcomes in some HNSCC subgroups.

2.1. Structure, Function, and Role of Nuclear Androgen Receptor in HNSCC

The nuclear androgen receptor (nAR) together with the estrogen, progesterone, and corticosteroid receptors belong to the steroid hormone receptor subfamily of the nuclear receptor superfamily. Structural elements of domains enable affinity binding to their ligands and response elements of the target genes for direct regulation of gene transcription [35]. The AR gene is located on the X chromosome (Xp11-12) and contains eight exons and seven introns that encode four structural domains that are common to all SHRs [36]. The prominent variability among the SHR family members is due to the amino acid sequence differences and the size of the N-terminal domain (NTD). This domain contains the ligand-independent activation function region AF1 and acts as a transcription driver or repressor depending on post-translational modifications [37,38]. Further, located next to the NTD is a highly conserved DNA binding domain (DBD) that contains two zinc finger motifs of four cysteine residues in each of them. Zinc finger proteins recognize DNA consensus sequences and promote the binding of the nAR to the androgen response elements (AREs) in promoters and enhancers of AR-regulated genes [39]. The short region between the DBD and C-terminal domain or ligand-binding domain (LBD) is the hinge region. Like the NTD, the hinge region is poorly conserved among the nuclear steroid receptors. Except for the nuclear localization signal, which can be contained within the hinge region, this region is also a site of post-translational modifications [38]. The LBD is a conserved domain amongst the nuclear receptors located at the C-terminal end of the receptor, and its structure creates a variable hydrophobic region for ligand binding and exhibits a ligand-dependent activation function motif (AF2) [40]. Furthermore, this region interacts with transcriptional intermediary factors (TIFs), co-activators, and co-repressors upon interaction with the ligand as well as with the N-terminal domain due to the stabilization of bounded androgens [41]. Importing the receptor into the nucleus is mediated by the nuclear localization signal (NLS) located between the DBD and hinge region. In contrast, the nuclear export signal (NES), as a part of the LBD, is responsible for the export of the nAR to the cytoplasm [42].
The nuclear androgen receptor is a 110 kDa ligand-inducible protein composed of 919 amino acids that is located in the cytoplasm and has two main mechanisms of action. The binding of androgens to the nAR results in a conformational change and phosphorylation of the protein as well as translocation of the androgen/nAR complex to the nucleus in the form of a homodimer. Once dimerized, nAR binds to the AREs within target genes and modulates gene transcription with various co-regulators [43] (Figure 2). In addition to the ligand binding, an alternative or noncanonical mechanism of nAR activation mediated by the activation of secondary signaling pathways such as mitogen-activated protein kinase (MAPK), protein kinase B (Akt), and extracellular signal-regulated kinase (ERK) also exists [44]. Interestingly, Trnski et al. recently showed the noncanonical activation of nAR via the active form of Sonic hedgehog protein (SHH-N) binding directly to the nAR through its cholesterol modification in androgen-independent LNCaP prostate cancer cells [45]. This interaction is one of the possible mechanisms of nAR activation in tumors.
AREs contain a pair of conserved sequences of 5’-GGTACAnnnTGTTCT-3’ that often are arranged as inverted repeats separated by a three-nucleotide spacer (5’-CGG-3’) [46,47]. According to the JASPAR 2022 database (https://jaspar.uio.no/matrix/MA0007.2 accessed on 21 December 2022), there are 11,206 AREs in the human genome (build GRCh37/hg19) [48]. However, so far fewer than 2000 human genes have been characterized as androgen-responsive (those whose transcription could be regulated either positively or negatively by nAR) [49]. The development of next-generation sequencing techniques—especially whole transcriptome sequencing (RNA-Seq) and chromatin immunoprecipitation followed by sequencing (ChIP-seq)—has enabled the examination of nAR-mediated transcriptional regulation at the genomic scale. Despite some differences in sets of discovered genes, which was due to the use of different sequencing platforms and experimental conditions, a core set of androgen-responsive genes has been repeatedly identified [50]. According to the Molecular Signatures Database (MSigDB), there are 101 hallmark androgen-responsive genes (MsigDB systematic name M5908) (Figure 2) [51]. These genes belong to several general gene families such as those that encode proteins involved in the secretory pathway, polyamine synthesis and lipogenesis, transcription, splicing, ribosomal biogenesis, mitogenesis, bioenergetics, and redox processes [52]. Dysregulation of those pathways is related to cancer etiopathology, so in addition to the well-known role of nAR in prostate cancer [53], nAR’s connection with breast [54], bladder [55], hepatocellular [56], ovarian [57], endometrial [58], uterine [59], and many other tumor types [60] is also known.
Currently, there is limited evidence of the significance of nAR expression in different subtypes of head and neck cancers. For example, immunohistochemical staining determined the nAR expression in 64% of laryngeal carcinoma cases; the expression was higher in well-differentiated cases and lower in poorly differentiated cases and those with lymphatic invasion [61]. Similarly, Fei et al. determined the significantly upregulated expression of nAR mRNA in laryngeal squamous cell carcinoma compared to the healthy tissue; in addition, AR-positive (AR+) nuclear immunostaining was observed in 77.1% of the samples [62]. Therefore, AR+ cytoplasmic staining in the neoplastic oral squamous cell carcinoma (OSCC) epithelium greater than 20% was significantly correlated with the increased risk of OSCC progression [63] and frequency of VEGF-positive lymphocytes and Ki67 in metastases [64]. In addition, in nonsquamous salivary duct carcinoma (SDC), nAR is one of the potential diagnostic immunohistochemical markers [65]. Furthermore, several studies have reported significantly increased nAR expression as well as the expression of androgen-responsive genes and an enhanced growth rate of the SCC-4, SCC-25, OECM-1 and SAS OSCC cell lines after treatment with dihydrotestosterone (DHT) [66]. Likewise, the nAR knockdown resulted in a 75% lower growth rate, significant apoptotic cell death, and inhibition of tumorigenicity in the above-mentioned cells [66]. Similarly, the DHT stimulation of nAR promoted cell migration and aggressiveness by increasing phosphorylation of the epidermal growth factor receptor (EGFR) and Akt in AR+ OSCC tumors. At the same time, this effect was not observed in an AR-negative (AR-) SCC-25 cell line, which suggested that nAR promotes the EGFR signaling pathway in head and neck tumors, therefore making it a perfect target for targeted therapy in this tumor type [67]. On the other hand, Collela et al. showed contradictory results; in fact, they found a lower expression level of nAR in OSCC than in the control samples [68]. Furthermore, the alterations in the [CAG]n repeats of the AR gene are also associated with poor outcomes in male patients with oral cavity or oropharyngeal cancers; shorter repeats (≤20) were correlated with a more aggressive tumor subset [69]. Taken together, in HNSCC patients, nuclear AR+ staining generally prevails and is associated with well- and moderately differentiated tumors without lymphatic invasion, while cytoplasmic AR+ staining is correlated with metastases. Table 1 summarizes the current knowledge about the role of the nuclear and currently known membrane ARs in HNSCC.

2.2. Structure, Function, and Role of Membrane Androgen Receptors in HNSCC

In contrast to the nuclear receptors, membrane steroid receptors mediate the rapid nongenomic effects of steroids through the activation of secondary messengers, different signaling pathways, and calcium ion flux. Immediate androgen actions have been described in various cell and tissue types. Research on membrane androgen receptors (mARs) has accelerated after rapid intracellular signaling pathway activation in AR- cells was observed [76]. So far, five multifunctional proteins have been described as plasma mARs: Ca2+ ion channels TRPM8 and CaV1.2, G-protein-coupled receptors OXER1 and GPRC6A, and zinc transporter ZIP9 (Figure 2).

2.2.1. Transient Receptor Potential Cation Channel Subfamily M Member 8 (TRPM8)

TRPM8 is the first of two calcium ion channels and androgen-related receptors at once located at the plasma membrane and endoplasmic reticulum of the cells. Due to its expression in the peripheral nervous system, in pain- and temperature-sensing neurons, respectively, it was shown that TRPM8 responded to cold physical stimulations, menthol, and icilin [77]. In addition, testosterone was shown to increase intracellular Ca2+ influx and activate the TRPM8 channel in PC-3 prostate cancer cells, dorsal root ganglion and hippocampal neurons, and the human embryonic kidney cell line HEK293 with almost the same affinity and specificity as for the nAR [78]. On the other hand, in the HEK293 cell line, which stably expresses recombinant TRPM8, testosterone did not cause significant Ca2+ oscillations. However, even that effect was partially reversed using AR siRNA or the nAR antagonist hydroxyflutamide, thereby indicating the importance of nAR in the absence of TRPM8 [79]. Furthermore, sustained exposure of the TRMP8 to testosterone caused channel desensitization—especially in men—due to higher circulating levels of testosterone [79]. TRPM8 is a potential early-stage prostate cancer biomarker because it is significantly upregulated in early stages and decreased in advanced stages of prostate tumors [80]. Likewise, TRPM8 is also a predicted diagnostic indicator of breast cancer, which promotes metastatic potential by activating the Akt/GSK-3β signaling pathway and thus regulates the epithelial–mesenchymal transition [81] as well as cellular autophagy via activation of AMP-activated protein kinase (AMPK) and Unc-51 like autophagy activating kinase 1 (ULK1) [82].
Studies have shown both intracellular and plasma membrane expression of the TRPM8 channel in HNSCC cell lines, contrary to exclusively membranous localization of other mARs. Interestingly, since cancer cells secrete matrix metalloproteinases, which degrade the basement membrane and the extracellular matrix and thus facilitate the cell migration and invasion ability, this mechanism was shown to potentially be associated with an elevated intracellular Ca2+ level in the androgen-dependent LNCaP prostate cancer cell line [83]. Similarly, the menthol-induced activity of the TRPM8 enhanced the migration and invasion potential of the HSC-3 and HSC-4 OSCC cell lines via an increase in the activation and gelatinase activity of matrix metallopeptidase 9 (MMP-9). On the other hand, the TRPM8 antagonist RQ-00203078 suppressed this activity, which suggested that stimulation of TRPM8 positively regulates MMP-9 [70]. Recent bioinformatical analysis of the HNSCC dataset from The Cancer Genome Atlas (TCGA-HNSC) indicated TRPM8 as one of the risk factors in HNSCC evolvement [71], but its main role in this type of tumorigenesis is still unknown.

2.2.2. Voltage-Dependent L-Type Calcium Channel Subunit Alpha-1C (CaV1.2)

CaV1.2 is a protein encoded by the CACNA1C gene whose primary role is pore formation for Ca2+ ion transport into the cell. It is a member of the L-type voltage-gated Ca2+ channel family [84]. The nomenclature is based on the type of response to voltage, meaning this type of channel has ‘long-lasting’ activity, and its activation requires a strong membrane depolarization [85]. Furthermore, the protein consists of three subunits (CaVα1, CaVα2δ, and CaVβ), and four CaVα1 subunits set up the ion-conducting channel. Alternative splice variants of the CACNA1C gene are associated with various abnormalities such as Timothy and Brugada syndromes and cancer progression [85,86,87]. Increased influx of the Ca2+ ions through the plasma membrane is related to the activation of Fos and c-Jun transcription factors, cAMP response element-binding (CREB) protein, and the nuclear factor of activated T cells (NFAT) [88]. Likewise, Ca2+ is indispensable in coordinating transitions between G1/S phases of the cell cycle via regulation of D-cyclin expression. All the above changes are implicated in breast cancer cell proliferation, migration, and invasion [89].
When comparing the gene expression profiles in OSCC-derived cell lines with normal oral tissues using a microarray analysis, CACNA1C was one of the detected genes with a confirmed significantly decreased expression in the HSC-2, HSC-3, Ca9-22, H-1, Sa3, and OK-92 OSCC cell lines compared to human normal oral keratinocytes [72]. In addition, another study compared genes dysregulated in the TCGA-HNSC dataset with the early molecular alterations in the oral cavity and esophagus induced by (S)-N’-nitrosonornicotine, a potent tobacco carcinogen, in rats; except for the genes involved in immune regulation and inflammation, most of the dysregulated genes in rats were tumorigenesis-associated and overlapped with altered genes in esophageal and head and neck tumors. This included CACNA1C, which was amplificated and/or upregulated in 10% of cases [73].

2.2.3. Oxoeicosanoid Receptor 1 (OXER1)

OXER1 is a G-protein-coupled receptor (GPCR) that is also known as G-protein-coupled receptor 170 (GPR170), 5-oxo-ETE G-protein-coupled receptor, or TG1019. The receptor can be activated by 5-oxoeicosatretraenoic acid (5-oxo-ETE) or some other product of the arachidonic acid metabolites and by 5-lipoxygenase (5-LOX) and peroxidase [90]. High expression of OXER1 in lung, liver, spleen, kidney, prostate, and breast cancer cells as well as in inflammatory cells (eosinophils, lymphocytes, monocytes, and neutrophils) and its activation by endogenous ligand 5-oxo-ETE mediated the intracellular actions in cell proliferation and survival, inflammatory responses, and steroidogenesis stimulation [91]. On the one hand, receptor activation led to the inhibition of cAMP production mediated by the Gαi subunit. In contrast, the calcium mobilization, chemotactic response, and activation of other molecular pathways were mediated by a stable dimeric complex Gβγ [90]. Most of these roles are related to the immune system, but OXER1 is also essential in promoting survival, apoptosis inhibition, and infiltration of inflammatory cells in prostate and breast cancers. Recently, it was reported that testosterone is an OXER1 antagonist that mediated the effects of 5-oxo-ETE on PI3K/Akt, FAK, and p38α activation [92]. The inhibition of the above signaling pathways (such as blocking 5-LOX) reduces cancer cell proliferation, migration, invasion, and apoptosis induction.
So far, no direct studies have investigated the role of OXER1 in HNSCC, but upregulation of the OXER1 gene was reported in HPV+ oropharyngeal and oral cancers [74]. As stated above, HPV infection has an essential role in HNSCC progression. Thus, OXER1 is potentially connected with HNSCC tumorigenesis and could present a novel drug target. Interestingly, López-Ozuna et al. identified the upregulated expression of OXER1 in 2N and 11N human normal oral epithelial cell lines exposed to water-pipe smoking (WPS) solution, thereby implying the potential risk of WPS consumption for HNSCC development [75].

2.2.4. G-Protein-Coupled Receptor Class C Group 6 Member A (GPRC6A)

GPRC6A is the second GPCR mAR. It is activated by a diverse range of ligands such as essential L-α-amino acids (L-arginine, L-lysine, and L-ornithine), cations (calcium, zinc, and magnesium), osteocalcin [93], and steroid hormones [94]. The expression of GPRC6A in hepatocytes, skeletal muscle myocytes, and adipocytes is related to its involvement in different physiological and pathological functions such as glucose and fat metabolism, inflammatory responses, steroid production and insulin secretion from β-cells, and tumorigenesis [95]. Furthermore, the role of GPRC6A in the male reproductive system was recently discovered. Oury et al. showed expression of GPRC6A in Leydig cells, where it regulated spermatogenesis and testosterone biosynthesis in male mice by acting as an osteocalcin receptor [96]. In addition, they demonstrated a reduction in circulating testosterone levels, a decreased sperm number, and a reduced size of the testis in GPRC6ALeydig KO mice. Likewise, in an AR- and GPRC6A-negative HEK293 cell line, testosterone and synthetic androgen R1881 could stimulate ERK activity in a dose-dependent manner after transfection with human GPRC6A. On the other hand, ERK activation was blocked using a G-protein inhibitor, which confirmed the hypothesis of androgen-mediated GPRC6A activation [94]. Additionally, it recently was shown that due to the similar structure of the undercarboxylated form of osteocalcin and the sex hormone-binding protein (SHBG), the latter—in an unliganded state—binds to the extracellular domain of GPRC6A and activates the receptor [97]. Computational modeling also predicted several testosterone binding sites on the GPRC6A receptor, although the specificity of its binding and GPRC6A receptor activation in vivo is still unclear [98]. Several studies confirmed a connection between GPRC6A upregulation and prostate cancer cell line proliferation [99,100,101], while there is still no evidence for the potential role of the GPRC6A receptor in HNSCC tumorigenesis.

2.2.5. Zinc Transporter Member 9 (ZIP9)

ZIP9 is a member of the zinc transporter Zrt- and Irt-like protein (ZIP) family encoded by the SLC39A9 gene in humans. The central role of ZIP transporters is the regulation of zinc homeostasis by increasing the intracellular zinc concentration and therefore maintaining the normal physiological function of the organism. Compared to the other ZIP family members that contain eight transmembrane domains and an extracellular C-terminus, ZIP9 is the only member with seven transmembrane domains and an intracellular C-terminal domain, which enables signalization via G proteins [102]. In addition to regulating cellular functions associated with zinc transport, ZIP9 was first identified as a mAR in Atlantic croaker fish ovarian follicle cells [103] and then immediately afterwards in the triple-negative human breast cancer cell lines MDA-MB-468 and MDA-MB-231 as well as in the PC-3 prostate cancer cell line, all of which were stably transfected with ZIP9 [104]. Treatment with testosterone in those cell lines led to G-protein activation, thereby increasing the cytosolic zinc concentrations and apoptosis induction through the upregulation of the proapoptotic genes BAX, TP53, and MAPK8. On the other hand, testosterone-induced increased expression of the proapoptotic factors was blocked by using a MAPK inhibitor (PD98059) or a zinc chelator, which suggested that the apoptotic response was mediated by MAPK activation and zinc influx [104]. Furthermore, when comparing the binding affinity and activation of nAR and ZIP9 via testosterone, DHT, and androstenedione, ZIP9 showed a higher affinity for testosterone regarding DHT or androstenedione [102]. Except for its role in apoptosis induction, an activated form of the ZIP9 receptor affected cell migration in prostate cancer cells [105]. Due to its broad expression in most tissues and cell types related to zinc homeostasis, ZIP9 is involved in cell proliferation, growth, and apoptosis and represents a potential drug target in different types of carcinomas. However, the connection between ZIP9 and HNSCC tumorigenesis is still unexplored.

3. Current Therapeutic Approaches to the Treatment of HNSCC

Since the primary treatment for most patients with HNSCC is surgery and radiotherapy, research is focused on developing effective targeted therapies and the identification of new biomarkers to reduce mortality and improve the treatment outcome as well as the life quality of the patients. The biggest challenge in this, as is the case with many other tumors, is a weak response to therapy and the development of therapy resistance due to the heterogeneous nature and aggressiveness of HNSCC. Likewise, the roles of both the tumor microenvironment and the oral microbiome in HNSCC pathogenesis are increasingly being investigated.

3.1. Immunotherapy Strategies for Head and Neck Cancer

Interestingly, about 70% of HPV-negative HNSCCs cases have the loss-of-function (LoF) mutations in tumor-suppressor p53 while 30% of patients have them in the mammalian target of rapamycin (mTOR) and around 15% have them in the EGFR and NOTCH oncogenes. In comparison, the EGFR is overexpressed in more than 80% of HNSCCs [10]. Therefore, to date, the FDA has fully approved three targeted immunotherapies for HNSCC treatment: the anti-EGFR monoclonal antibody cetuximab as well as pembrolizumab and nivolumab, which are programmed cell death protein 1 (PD-1) inhibitors [13]. Cetuximab mediates in an oncogenic signal blocking and killing tumor cells by activating antibody-dependent cell-mediated cytotoxicity (ADCC) via natural killer (NK) cells and monocytes [64]. It was used for a decade as the first line of HNSCC treatment together with other platinum-based drugs such as cisplatin and 5-fluorouracil (5-FU). In contrast, the latter two antibodies block the interaction of PD-1 with its programmed death-ligand 1 (PD-L1) and are used to treat recurrent or metastatic (R/M) HNSCC [106]. Since the KEYNOTE-048 trial was published in 2019, pembrolizumab has been approved as a first-line treatment for PD-L1-positive R/M HNSCC and in combination with chemotherapy for patients with any PD-L1 status [107]. However, the efficacy of targeted therapies is reduced due to the development of drug resistance. For instance, EGFR shares common downstream pathways (the RAS-RAF-MAPK and PI3K-AKT-mTOR) with alternative receptor tyrosine kinases (RTK) such as hepatocyte growth factor receptor (c-MET), human epidermal growth factor receptor 3 (HER3), and AXL receptor tyrosine kinase, so any mutations or aberrant signaling activity that lead to epidermal growth factor (EGF)-independent signal transduction can cause resistance development [108].

3.2. Applicability of Antiandrogen Therapy in HNSCC

On the other hand, given the evidence of elevated SHR expression in HNSCC, nAR also represents a potential target for antiandrogen therapy in these tumors (Figure 2)—notably in SDC. SDC is a very rare and aggressive subtype of salivary gland carcinoma (1–3%) with a 5-year survival rate below 20% [109]. Interestingly, nAR is expressed in 70–98% of SDC cases [110,111] and significantly more often in men than in women [112]. Therefore, based on nAR expression, in recent years the first-line treatment for AR+ SDC is androgen deprivation therapy (ADT) using the first-generation of nonsteroidal antiandrogens such as bicalutamide [113] and flutamide as well as abiraterone acetate [114], an inhibitor of androgen synthesis [115]. Several studies have shown the benefits of ADT use in AR+ SDC in terms of a higher response rate and a better prognosis and clinical outcome [109,116,117,118]. Locati et al. reported a 64.7% overall response rate and three complete clinical remissions in a cohort of 17 AR+ SDC patients who received ADT [119]. Furthermore, ADT proved less toxic but equally effective in both R/M and unresectable locally advanced SDC compared with conventional chemotherapy [120]. First-line ADT therapy also demonstrated a higher response rate than the first-line chemotherapy treatment in SDC patients [109]. Nonetheless, combining chemotherapy or radiotherapy with ADT seems to be the most effective for AR+ HNSCC treatment because nAR expression positively correlates with overall and metastases-free survival [121]. Interestingly, an in vitro study also showed that flutamide treatment of the HSY salivary gland cancer cell line could block the effects of DHT, which has been shown to increase cell proliferation, migration, and invasion [114].
Since the vast majority of SDC tumors express nAR, treatment with nonsteroidal antiandrogens is often used in combination with GnRH agonists such as triptorelin and goserelin and luteinizing hormone-releasing hormone (LHRH) agonists (leuprorelin acetate) [122]. This type of combined androgen blockade (CAB) has an established role in prostate cancer treatment, and therefore translating these combined treatment strategies to SDC was studied in a phase II study that suggested the equivalent efficacy and lower toxicity of CAB for SDC patients compared with chemotherapy [120]. Currently, in the treatment of prostate cancer, new-generation nonsteroidal antiandrogens such as abiraterone acetate, enzalutamide, apalutamide, and darolutamide also are used [123]. Enzalutamide monotherapy treatment was used for AR+ SDC patients in a phase II trial in which only 15% of patients (7/46) had a partial response and 2/46 patients maintained the response until 8 weeks [124]. On the other hand, in the first published report of SDC responding to abiraterone, a 10-month progression-free survival was described in one patient [125]. Furthermore, 19 AR+ SDC patients were treated with abiraterone plus prednisone and an LHRH agonist in another phase II trial. The results showed that this combined therapy represented an active and safe second-line treatment option for AR+ SDC with a 12-month overall survival rate of 74.5% [126]. Likewise, the first prospective trials that evaluated CAB with apalutamide and gorselin as a GnRH agonist (NCT04325828) as well as with darolutamide and gorselin for AR+ SDC treatment are currently ongoing (NCT05694819). Although SDC can be classified as an androgen-dependent tumor, the growth of which can be controlled using ADT, the effectiveness and molecular mechanisms underlying the application of ADT in this type of tumor have yet to be investigated.

4. Conclusions

Disparities in HNSCC incidence and mortality rates between males and females are more than evident, and this could not be associated only with the different lifestyles. Although present levels of evidence are still quite limited and scarce, potential toxic effects of male sex hormones in HNSCC development and progression either alone or in combination with other well-known risk factors could be discovered by studying the role of ARs in the etiopathology of HNSCC. At present, nAR is the most studied AR in HNSCC. However, using the expression status of nAR as a prognostic or predictive biomarker for HNSCC is still far from a routine clinical practice. The only exception is AR+ salivary duct carcinomas, for which several clinical trials have shown an advantage in using the new generation of nonsteroidal antiandrogens in terms of a higher response rate and a better prognosis and clinical outcome. In that light, the substitution of traditional ADT with new-generation drugs or even using CAB might provide a better quality of life for all patients with AR+ HNSCC.

Author Contributions

Conceptualization, J.Č. and P.O.; writing—original draft preparation, J.Č.; writing—review and editing, M.S. and P.O.; visualization, J.Č.; supervision, P.O.; funding acquisition, M.S. and P.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Croatian Science Foundation project “Differential regulation of the GLI code in BRAF/NRAS driven tumors” (IP-2018-01-4889) and by the Terry Fox Foundation with a donation collected during the Terry Fox Run in Zagreb in 2019. The work of doctoral student J.Č. was supported by the “Young Researchers’ Career Development Project-Training of Doctoral Students” of the Croatian Science Foundation (DOK-2021-02-5821).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank the Croatian League Against Cancer and all organizers and participants of the Terry Fox Run Croatia in 2019 for helping us to acquire the funding for this research.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the conceptualization or writing of this manuscript.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA A Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  2. Mukherjee, A.; Idigo, A.J.; Ye, Y.; Wiener, H.W.; Paluri, R.; Nabell, L.M.; Shrestha, S. Geographical and Racial Disparities in Head and Neck Cancer Diagnosis in South-Eastern United States: Using Real-World Electronic Medical Records Data. Health Equity 2020, 4. [Google Scholar] [CrossRef]
  3. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. CA A Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  4. Bagnardi, V.; Rota, M.; Botteri, E.; Tramacere, I.; Islami, F.; Fedirko, V.; Scotti, L.; Jenab, M.; Turati, F.; Pasquali, E.; et al. Alcohol Consumption and Site-Specific Cancer Risk: A Comprehensive Dose–Response Meta-Analysis. Br. J. Cancer 2015, 112, 580–593. [Google Scholar] [CrossRef]
  5. Lubin, J.H.; Purdue, M.; Kelsey, K.; Zhang, Z.-F.; Winn, D.; Wei, Q.; Talamini, R.; Szeszenia-Dabrowska, N.; Sturgis, E.M.; Smith, E.; et al. Meta-and Pooled Analyses Total Exposure and Exposure Rate Effects for Alcohol and Smoking and Risk of Head and Neck Cancer: A Pooled Analysis of Case-Control Studies. Am. J. Epidemiol. 2009, 170, 937–947. [Google Scholar] [CrossRef]
  6. Miranda-Galvis, M.; Loveless, R.; Paulo Kowalski, L.; Teng, Y. Impacts of Environmental Factors on Head and Neck Cancer Pathogenesis and Progression. Cells 2021, 10, 389. [Google Scholar] [CrossRef]
  7. Rampias, T.; Sasaki, C.; Weinberger, P.; Psyrri, A. E6 and E7 Gene Silencing and Transformed Phenotype of Human Papillomavirus 16–Positive Oropharyngeal Cancer Cells. JNCI J. Natl. Cancer Inst. 2009, 101, 412–423. [Google Scholar] [CrossRef]
  8. Krump, N.A.; You, J. Molecular Mechanisms of Viral Oncogenesis in Humans. Nat. Rev. Microbiol. 2018, 16, 684–698. [Google Scholar] [CrossRef]
  9. Mishra, A.; Meherotra, R. Head and Neck Cancer: Global Burden and Regional Trends in India. Asian Pac. J. Cancer Prev. APJCP 2014, 15, 537–550. [Google Scholar] [CrossRef]
  10. Alsahafi, E.; KatherynBegg; Amelio, I.; Raulf1, N.; Lucarelli, P.; Sauter, T.; Tavassoli, M. Clinical Update on Head and Neck Cancer: Molecular Biology and Ongoing Challenges Elham. Cell Death Dis. 2019, 10, 540. [Google Scholar] [CrossRef]
  11. Johnson, D.E.; Burtness, B.; Leemans, C.R.; Lui, V.W.Y.; Bauman, J.E.; Grandis, J.R. Head and Neck Squamous Cell Carcinoma. Nat. Rev. Dis. Prim. 2020, 6, 92. [Google Scholar] [CrossRef]
  12. Hsieh, J.C.H.; Wang, H.M.; Wu, M.H.; Chang, K.P.; Chang, P.H.; Liao, C.T.; Liau, C.T. Review of Emerging Biomarkers in Head and Neck Squamous Cell Carcinoma in the Era of Immunotherapy and Targeted Therapy. Head Neck 2019, 41, 19–45. [Google Scholar] [CrossRef]
  13. Goel, B.; Tiwari, A.K.; Pandey, R.K.; Singh, A.P.; Kumar, S.; Sinha, A.; Jain, S.K.; Khattri, A. Therapeutic Approaches for the Treatment of Head and Neck Squamous Cell Carcinoma–An Update on Clinical Trials. Transl. Oncol. 2022, 21, 101426. [Google Scholar] [CrossRef]
  14. Basheeth, N.; Patil, N. Biomarkers in Head and Neck Cancer an Update. Indian J. Otolaryngol. Head Neck Surg. 2019, 71, 1002–1011. [Google Scholar] [CrossRef]
  15. Trigo, J.; García-Cosío, M.; García-Castaño, A.; Gomà, M.; Mesia-Nin, R.; Ruiz-Bravo, E.; Soria-Rivas, A.; Castillo, P.; Braña-García, I.; Alberola-Ferranti, M. Recommendations for the Use of Biomarkers for Head and Neck Cancer, Including Salivary Gland Tumours: A Consensus of the Spanish Society of Medical Oncology and the Spanish Society of Pathology. Clin. Transl. Oncol. 2022, 24, 1890–1902. [Google Scholar] [CrossRef]
  16. Dong, M.; Cioffi, G.; Wang, J.; Waite, K.A.; Ostrom, Q.T.; Kruchko, C.; Lathia, J.D.; Rubin, J.B.; Berens, M.E.; Connor, J.; et al. Sex Differences in Cancer Incidence and Survival: A Pan-Cancer Analysis. Cancer Epidemiol. Biomark. Prev. 2020, 29, 1389–1397. [Google Scholar] [CrossRef]
  17. Suba, Z. Gender-Related Hormonal Risk Factors for Oral Cancer. Pathol. Oncol. Res. 2007, 13, 195–202. [Google Scholar] [CrossRef]
  18. McGuire, W.L.; Chamness, G.C.; Costlow, M.E.; Shepherd, R.E. Hormone Dependence in Breast Cancer. Metabolism 1974, 23, 75–100. [Google Scholar] [CrossRef]
  19. Modugno, F.; Laskey, R.; Smith, A.L.; Andersen, C.L.; Haluska, P.; Oesterreich, S. Hormone Response in Ovarian Cancer: Time to Reconsider as a Clinical Target? Endocr. Relat. Cancer 2012, 19, 255–279. [Google Scholar] [CrossRef]
  20. Heinlein, C.A.; Chang, C. Androgen Receptor in Prostate Cancer. Endocr. Rev. 2004, 25, 276–308. [Google Scholar] [CrossRef]
  21. Imai, Y.; Yamakawa, M.; Matsuda, M.; Kasajima, T. Endogenous Sex Hormone and Estrogen Binding Activity in Thyroid Cancer. Histol. Histopathol. 1989, 4, 39–45. [Google Scholar] [PubMed]
  22. Fénichel, P.; Chevalier, N. Is Testicular Germ Cell Cancer Estrogen Dependent? The Role of Endocrine Disrupting Chemicals. Endocrinology 2019, 160, 2981–2989. [Google Scholar] [CrossRef] [PubMed]
  23. Rodriguez, A.C.; Blanchard, Z.; Maurer, K.A.; Gertz, J. Estrogen Signaling in Endometrial Cancer: A Key Oncogenic Pathway with Several Open Questions. Horm. Cancer 2019, 10, 51–63. [Google Scholar] [CrossRef] [PubMed]
  24. Henderson, B.E.; Feigelson, H.S. Hormonal Carcinogenesis. Carcinogenesis 2000, 21, 427–433. [Google Scholar] [CrossRef]
  25. Hu, J.; Zhang, Z.; Shen, W.J.; Azhar, S. Cellular Cholesterol Delivery, Intracellular Processing and Utilization for Biosynthesis of Steroid Hormones. Nutr. Metab. 2010, 7, 47. [Google Scholar] [CrossRef]
  26. Handa, R.J.; Weiser, M.J. Gonadal Steroid Hormones and the Hypothalamo-Pituitary-Adrenal Axis. Front. Neuroendocrinol. 2014, 35, 197–220. [Google Scholar] [CrossRef]
  27. Melmed, S.; Polonsky, K.S.; Larsen, P.R.; Kronenberg, H.M. WILLIAMS Textbook of Endocrinology, 12th ed.; Elsevier: Amsterdam, The Netherlands, 1994; Volume 308, ISBN 9781437703245. [Google Scholar]
  28. Ing, H.N. Steroid Hormones Regulate Gene Expression Posttranscriptionally by Altering the Stabilities of Messenger RNAs. Biol. Reprod. 2005, 72, 1290–1296. [Google Scholar] [CrossRef]
  29. Schwartz, N.; Verma, A.; Bivens, C.B.; Schwartz, Z.; Boyan, B.D. Rapid Steroid Hormone Actions via Membrane Receptors. Biochim. Biophys. Acta 2016, 1863, 2289–2298. [Google Scholar] [CrossRef]
  30. Treviño, L.S.; Gorelick, D.A. The Interface of Nuclear and Membrane Steroid Signaling. Endocrinology 2021, 162, bqab107. [Google Scholar] [CrossRef]
  31. Masi, M.; Racchi, M.; Travelli, C.; Corsini, E.; Buoso, E. Molecular Characterization of Membrane Steroid Receptors in Hormone-Sensitive Cancers. Cells 2021, 10, 2999. [Google Scholar] [CrossRef]
  32. Neto, C.P.D.O.; Brito, H.O.; Da Costa, R.M.G.; Brito, L.M.O. Is There a Role for Sex Hormone Receptors in Head-and-Neck Cancer? Links with HPV Infection and Prognosis. Anticancer Res. 2021, 41, 3707–3716. [Google Scholar] [CrossRef]
  33. Park, J.O.; Nam, I.C.; Kim, C.S.; Park, S.J.; Lee, D.H.; Kim, H.B.; Han, K.D.; Joo, Y.H. Sex Differences in the Prevalence of Head and Neck Cancers: A 10-Year Follow-Up Study of 10 Million Healthy People. Cancers 2022, 14, 2521. [Google Scholar] [CrossRef]
  34. Singh, H.; Kumar, P.; Goel, R.; Kumar, A. Sex Hormones in Head and Neck Cancer: Current Knowledge and Perspectives. Clin. Cancer Investig. J. 2012, 1, 2–5. [Google Scholar] [CrossRef]
  35. Evans, R.M. The Steroid and Thyroid Hormone Receptor Superfamily. Science 1988, 240, 889–895. [Google Scholar] [CrossRef]
  36. Eisermann, K.; Wang, D.; Jing, Y.; Pascal, L.E.; Wang, Z. Androgen Receptor Gene Mutation, Rearrangement, Polymorphism. Transl. Androl. Urol. 2013, 2, 137–147. [Google Scholar] [CrossRef]
  37. Anbalagan, M.; Huderson, B.; Murphy, L.; Rowan, B.G. Post-Translational Modifications of Nuclear Receptors and Human Disease. Nucl. Recept. Signal. 2012, 10, e001. [Google Scholar] [CrossRef] [PubMed]
  38. Lavery, D.N.; McEwan, I.J. Structure and Function of Steroid Receptor AF1 Transactivation Domains: Induction of Active Conformations. Biochem. J. 2005, 391, 449–464. [Google Scholar] [CrossRef] [PubMed]
  39. Davey, R.A.; Grossmann, M. Androgen Receptor Structure, Function and Biology: From Bench to Bedside. Clin. Biochemist. Rev. 2016, 37, 3–15. [Google Scholar]
  40. Wärnmark, A.; Treuter, E.; Wright, A.P.H.; Gustafsson, J.-A. Activation Functions 1 and 2 of Nuclear Receptors: Molecular Strategies for Transcriptional Activation. Mol. Endocrinol. 2003, 17, 1901–1909. [Google Scholar] [CrossRef] [PubMed]
  41. Zhou, Z.X.; Lane, M.V.; Kemppainen, J.A.; French, F.S.; Wilson, E.M. Specificity of Ligand-Dependent Androgen Receptor Stabilization: Receptor Domain Interactions Influence Ligand Dissociation and Receptor Stability. Mol. Endocrinol. 1995, 9, 208–218. [Google Scholar] [CrossRef] [PubMed]
  42. Saporita, A.J.; Zhang, Q.; Navai, N.; Dincer, Z.; Hahn, J.; Cai, X.; Wang, Z. Identification and Characterization of a Ligand-Regulated Nuclear Export Signal in Androgen Receptor*. J. Biol. Chem. 2003, 278, 41998–42005. [Google Scholar] [CrossRef] [PubMed]
  43. Dehm, S.M.; Tindall, D.J. Androgen Receptor Structural and Functional Elements: Role and Regulation in Prostate Cancer. Mol. Endocrinol. 2007, 21, 2855–2863. [Google Scholar] [CrossRef]
  44. Liao, R.S.; Ma, S.; Miao, L.; Li, R.; Yin, Y.; Raj, G.V. Androgen Receptor-Mediated Non-Genomic Regulation of Prostate Cancer Cell Proliferation. Transl. Androl. Urol. 2013, 2, 187–196. [Google Scholar] [CrossRef] [PubMed]
  45. Trnski, D.; Sabol, M.; Tomić, S.; Štefanac, I.; Mrčela, M.; Musani, V.; Rinčić, N.; Kurtović, M.; Petrić, T.; Levanat, S.; et al. SHH-N Non-Canonically Sustains Androgen Receptor Activity in Androgen-Independent Prostate Cancer Cells. Sci. Rep. 2021, 11, 14880. [Google Scholar] [CrossRef] [PubMed]
  46. Laffitte, B.A.; Kast, H.R.; Nguyen, C.M.; Zavacki, A.M.; Moore, D.D.; Edwards, P.A. Identification of the DNA Binding Specificity and Potential Target Genes for the Farnesoid X-Activated Receptor*. J. Biol. Chem. 2000, 275, 10638–10647. [Google Scholar] [CrossRef] [PubMed]
  47. Haelens, A.; Verrijdt, G.; Callewaert, L.; Peeters, B.; Rombauts, W.; Claessens, F. Androgen-Receptor-Specific DNA Binding to an Element in the First Exon of the Human Secretory Component Gene. Biochem. J. 2001, 353, 611–620. [Google Scholar] [CrossRef] [PubMed]
  48. Castro-Mondragon, J.A.; Riudavets-Puig, R.; Rauluseviciute, I.; Lemma, R.B.; Turchi, L.; Blanc-Mathieu, R.; Lucas, J.; Boddie, P.; Khan, A.; Manosalva Pérez, N.; et al. JASPAR 2022: The 9th Release of the Open-Access Database of Transcription Factor Binding Profiles. Nucleic Acids Res. 2022, 50, D165–D173. [Google Scholar] [CrossRef]
  49. Jiang, M.; Ma, Y.; Chen, C.; Fu, X.; Yang, S.; Li, X.; Yu, G.; Mao, Y.; Xie, Y.; Li, Y. Androgen-Responsive Gene Database: Integrated Knowledge on Androgen-Responsive Genes. Mol. Endocrinol. 2009, 23, 1927–1933. [Google Scholar] [CrossRef]
  50. Jin, H.-J.; Kim, J.; Yu, J. Androgen Receptor Genomic Regulation. Transl. Androl. Urol. 2013, 2, 158–177. [Google Scholar] [CrossRef]
  51. Liberzon, A.; Birger, C.; Thorvaldsdóttir, H.; Ghandi, M.; Mesirov, J.P.; Tamayo, P. The Molecular Signatures Database (MSigDB) Hallmark Gene Set Collection. Cell Syst. 2015, 1, 417–425. [Google Scholar] [CrossRef]
  52. Kaarbø, M.; Klokk, T.I.; Saatcioglu, F. Androgen Signaling and Its Interactions with Other Signaling Pathways in Prostate Cancer. BioEssays 2007, 29, 1227–1238. [Google Scholar] [CrossRef]
  53. Fujita, K.; Nonomura, N. Role of Androgen Receptor in Prostate Cancer: A Review. World J. Mens Health 2019, 37, 288. [Google Scholar] [CrossRef]
  54. Anestis, A.; Zoi, I.; Papavassiliou, A.G.; Karamouzis, M.V. Androgen Receptor in Breast Cancer-Clinical and Preclinical Research Insights. Molecules 2020, 25, 358. [Google Scholar] [CrossRef]
  55. Lombard, A.P.; Mudryj, M. The Emerging Role of the Androgen Receptor in Bladder Cancer. Endocr. Relat. Cancer 2015, 22, R265–R277. [Google Scholar] [CrossRef]
  56. Zhou, H.-C.; Liu, C.-X.; Pan, W.-D.; Shang, L.-R.; Zheng, J.-L.; Huang, B.-Y.; Chen, J.-Y.; Zheng, L.; Fang, J.-H.; Zhuang, S.-M. Dual and Opposing Roles of the Androgen Receptor in VETC-Dependent and Invasion-Dependent Metastasis of Hepatocellular Carcinoma. J. Hepatol. 2021, 75, 900–911. [Google Scholar] [CrossRef] [PubMed]
  57. Chung, W.-M.; Chen, L.; Chang, W.-C.; Su, S.-Y.; Hung, Y.-C.; Ma, W.-L. Androgen/Androgen Receptor Signaling in Ovarian Cancer: Molecular Regulation and Therapeutic Potentials. Int. J. Mol. Sci. 2021, 22, 7748. [Google Scholar] [CrossRef] [PubMed]
  58. Tangen, I.L.; Onyango, T.B.; Kopperud, R.; Berg, A.; Halle, M.K.; Øyan, A.M.; Werner, H.M.J.; Trovik, J.; Kalland, K.H.; Salvesen, H.B.; et al. Androgen Receptor as Potential Therapeutic Target in Metastatic Endometrial Cancer. Oncotarget 2016, 7, 49289–49298. [Google Scholar] [CrossRef] [PubMed]
  59. Choi, J.P.; Desai, R.; Zheng, Y.; Yao, M.; Dong, Q.; Watson, G.; Handelsman, D.J.; Simanainen, U. Androgen Actions via Androgen Receptor Promote PTEN Inactivation Induced Uterine Cancer. Endocr. Relat. Cancer 2015, 22, 687–701. [Google Scholar] [CrossRef]
  60. Munoz, J.; Wheler, J.J.; Kurzrock, R. Androgen Receptors beyond Prostate Cancer: An Old Marker as a New Target. Oncotarget 2014, 6, 592–603. [Google Scholar] [CrossRef]
  61. Atef, A.; El-Rashidy, M.A.; Elzayat, S.; Kabel, A.M. The Prognostic Value of Sex Hormone Receptors Expression in Laryngeal Carcinoma. Tissue Cell 2019, 57, 84–89. [Google Scholar] [CrossRef]
  62. Fei, M.; Zhang, J.; Zhou, J.; Xu, Y.; Wang, J. Sex-Related Hormone Receptor in Laryngeal Squamous Cell Carcinoma: Correlation with Androgen Estrogen-ɑ and Prolactin Receptor Expression and Influence of Prognosis. Acta Oto-Laryngol. 2018, 138, 66–72. [Google Scholar] [CrossRef]
  63. Tomasovic-Loncaric, C.; Fucic, A.; Andabak, A.; Andabak, M.; Ceppi, M.; Bruzzone, M.; Vrdoljak, D.; Vucicevic-Boras, V. Androgen Receptor as a Biomarker of Oral Squamous Cell Carcinoma Progression Risk. Anticancer Res. 2019, 39, 4285–4289. [Google Scholar] [CrossRef] [PubMed]
  64. Batelja-Vuletic, L.; Tomasovic-Loncaric, C.; Ceppi, M.; Bruzzone, M.; Fucic, A.; Krstanac, K.; Boras Vucicevic, V. Comparison of Androgen Receptor, VEGF, HIF-1, Ki67 and MMP9 Expression between Non-Metastatic and Metastatic Stages in Stromal and Tumor Cells of Oral Squamous Cell Carcinoma. Life 2021, 11, 336. [Google Scholar] [CrossRef]
  65. Rahimi, S.; Bradley, K.L.; Akaev, I.; Marani, C.; Yeoh, C.C.; Brennan, P.A. Androgen Receptor Expression in Primary Nonsquamous Cell Rare-Variant of the Head and Neck. OTO Open 2017, 1, 2473974X17715633. [Google Scholar] [CrossRef] [PubMed]
  66. Wu, T.F.; Luo, F.J.; Chang, Y.L.; Huang, C.M.; Chiu, W.J.; Weng, C.F.; Hsu, Y.K.; Yuan, T.C. The Oncogenic Role of Androgen Receptors in Promoting the Growth of Oral Squamous Cell Carcinoma Cells. Oral Dis. 2015, 21, 320–327. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, X.; Qing, S.; Che, K.; Li, L.; Liao, X. Androgen Receptor Promotes Oral Squamous Cell Carcinoma Cell Migration by Increasing EGFR Phosphorylation. OncoTargets Ther. 2019, 12, 4245–4252. [Google Scholar] [CrossRef] [PubMed]
  68. Colella, G.; Izzo, G.; Carinci, F.; Campisi, G.; Lo Muzio, L.; D’Amato, S.; Mazzotta, M.; Cannavale, R.; Ferrara, D.; Minucci, S. Expression of Sexual Hormones Receptors in Oral Squamous Cell Carcinoma. Int. J. Immunopathol. Pharmacol. 2011, 24, 129–132. [Google Scholar] [CrossRef] [PubMed]
  69. Rosa, F.E.; dos Santos, R.M.; Poli-Frederico, R.C.; de Azevedo Canevari, R.; Nishimoto, I.N.; Magrin, J.; Rainho, C.A.; Kowalski, L.P.; Rogatto, S.R. Shorter CAG Repeat Length in the AR Gene Is Associated with Poor Outcome in Head and Neck Cancer. Arch. Oral Biol. 2007, 52, 732–739. [Google Scholar] [CrossRef]
  70. Okamoto, Y.; Ohkubo, T.; Ikebe, T.; Yamazaki, J. Blockade of TRPM8 Activity Reduces the Invasion Potential of Oral Squamous Carcinoma Cell Lines. Int. J. Oncol. 2012, 40, 1431–1440. [Google Scholar] [CrossRef]
  71. Pan, F.; Wang, K.; Zheng, M.; Ren, Y.; Hao, W.; Yan, J. A TRP Family Based Signature for Prognosis Prediction in Head and Neck Squamous Cell Carcinoma. J. Oncol. 2022, 2022, 8757656. [Google Scholar] [CrossRef]
  72. Nomura, H.; Uzawa, K.; Yamano, Y.; Fushimi, K.; Ishigami, T.; Kato, Y.; Saito, K.; Nakashima, D.; Higo, M.; Kouzu, Y.; et al. Network-Based Analysis of Calcium-Binding Protein Genes Identifies Grp94 as a Target in Human Oral Carcinogenesis. Br. J. Cancer 2007, 97, 792–801. [Google Scholar] [CrossRef]
  73. Khammanivong, A.; Anandharaj, A.; Qian, X.; Song, J.M.; Upadhyaya, P.; Balbo, S.; Bandyopadhyay, D.; Dickerson, E.B.; Hecht, S.S.; Kassie, F. Transcriptome Profiling in Oral Cavity and Esophagus Tissues from (S)-N′-Nitrosonornicotine-Treated Rats Reveals Candidate Genes Involved in Human Oral Cavity and Esophageal Carcinogenesis. Mol. Carcinog. 2016, 55, 2168–2182. [Google Scholar] [CrossRef]
  74. Tomar, S.; Graves, C.A.; Altomare, D.; Kowli, S.; Kassler, S.; Sutkowski, N.; Gillespie, M.B.; Creek, K.E.; Pirisi, L. Human Papillomavirus Status and Gene Expression Profiles of Oropharyngeal and Oral Cancers from European American and African American Patients. Head Neck 2016, 38, E694–E704. [Google Scholar] [CrossRef]
  75. López-Ozuna, V.M.; Gupta, I.; Kiow, R.L.C.; Matanes, E.; Kheraldine, H.; Yasmeen, A.; Khalil, A.; Vranic, S.; Al Moustafa, A.-E.; Farsi, H.F.A. Water-Pipe Smoking Exposure Deregulates a Set of Genes Associated with Human Head and Neck Cancer Development and Prognosis. Toxics 2020, 8, 73. [Google Scholar] [CrossRef]
  76. Benten, W.P.M.; 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]
  77. Peier, A.M.; Moqrich, A.; Hergarden, A.C.; Reeve, A.J.; Andersson, D.A.; Story, G.M.; Earley, T.J.; Dragoni, I.; McIntyre, P.; Bevan, S.; et al. A TRP Channel That Senses Cold Stimuli and Menthol. Cell 2002, 108, 705–715. [Google Scholar] [CrossRef]
  78. Asuthkar, S.; Elustondo, P.A.; Demirkhanyan, L.; Sun, X.; Baskaran, P.; Velpula, K.K.; Thyagarajan, B.; Pavlov, E.V.; Zakharian, E. The TRPM8 Protein Is a Testosterone Receptor: I. Biochemical Evidence for Direct TRPM8-Testosterone Interactions. J. Biol. Chem. 2015, 290, 2659–2669. [Google Scholar] [CrossRef]
  79. Asuthkar, S.; Demirkhanyan, L.; Sun, X.; Elustondo, P.A.; Krishnan, V.; Baskaran, P.; Velpula, K.K.; Thyagarajan, B.; Pavlov, E.V.; Zakharian, E. The TRPM8 Protein Is a Testosterone Receptor: II. Functional Evidence for an Ionotropic Effect of Testosterone on TRPM8. J. Biol. Chem. 2015, 290, 2670–2688. [Google Scholar] [CrossRef] [PubMed]
  80. Henshall, S.M.; Afar, D.E.H.; Hiller, J.; Horvath, L.G.; Quinn, D.I.; Rasiah, K.K.; Gish, K.; Willhite, D.; Kench, J.G.; Gardiner-Garden, M.; et al. Survival Analysis of Genome-Wide Gene Expression Profiles of Prostate Cancers Identifies New Prognostic Targets of Disease Relapse. Cancer Res. 2003, 63, 4196–4203. [Google Scholar] [PubMed]
  81. Liu, J.; Chen, Y.; Shuai, S.; Ding, D.; Li, R.; Luo, R. TRPM8 Promotes Aggressiveness of Breast Cancer Cells by Regulating EMT via Activating AKT/GSK-3β Pathway. Tumour Biol. 2014, 35, 8969–8977. [Google Scholar] [CrossRef] [PubMed]
  82. Huang, Y.; Li, S.; Jia, Z.; Zhao, W.; Zhou, C.; Zhang, R.; Ali, D.W.; Michalak, M.; Chen, X.Z.; Tang, J. Transient Receptor Potential Melastatin 8 (TRPM8) Channel Regulates Proliferation and Migration of Breast Cancer Cells by Activating the AMPK-ULK1 Pathway to Enhance Basal Autophagy. Front. Oncol. 2020, 10, 2645. [Google Scholar] [CrossRef] [PubMed]
  83. Monet, M.; Lehen’kyi, V.; Gackiere, F.; Firlej, V.; Vandenberghe, M.; Roudbaraki, M.; Gkika, D.; Pourtier, A.; Bidaux, G.; Slomianny, C.; et al. Role of Cationic Channel TRPV2 in Promoting Prostate Cancer Migration and Progression to Androgen Resistance. Cancer Res. 2010, 70, 1225–1235. [Google Scholar] [CrossRef] [PubMed]
  84. Hofmann, F.; Flockerzi, V.; Kahl, S.; Wegener, J.W. L-Type CaV1.2 Calcium Channels: From In Vitro Findings to In Vivo Function. Physiol. Rev. 2014, 94, 303–326. [Google Scholar] [CrossRef] [PubMed]
  85. Zamponi, G.W.; Striessnig, J.; Koschak, A.; Dolphin, A.C. The Physiology, Pathology, and Pharmacology of Voltage-Gated Calcium Channels and Their Future Therapeutic Potential. Pharmacol. Rev. 2015, 67, 821–870. [Google Scholar] [CrossRef]
  86. Splawski, I.; Timothy, K.W.; Sharpe, L.M.; Decher, N.; Kumar, P.; Bloise, R.; Napolitano, C.; Schwartz, P.J.; Joseph, R.M.; Condouris, K.; et al. Ca(V)1.2 Calcium Channel Dysfunction Causes a Multisystem Disorder Including Arrhythmia and Autism. Cell 2004, 119, 19–31. [Google Scholar] [CrossRef]
  87. Fukuyama, M.; Wang, Q.; Kato, K.; Ohno, S.; Ding, W.-G.; Toyoda, F.; Itoh, H.; Kimura, H.; Makiyama, T.; Ito, M.; et al. Long QT Syndrome Type 8: Novel CACNA1C Mutations Causing QT Prolongation and Variant Phenotypes. Europace 2014, 16, 1828–1837. [Google Scholar] [CrossRef]
  88. Thiel, G.; Schmidt, T.; Rössler, O.G. Ca2+ Microdomains, Calcineurin and the Regulation of Gene Transcription. Cells 2021, 10, 875. [Google Scholar] [CrossRef]
  89. Déliot, N.; Constantin, B. Plasma Membrane Calcium Channels in Cancer: Alterations and Consequences for Cell Proliferation and Migration. Biochim. Biophys. Acta 2015, 1848, 2512–2522. [Google Scholar] [CrossRef]
  90. Grant, G.E.; Rokach, J.; Powell, W.S. 5-Oxo-ETE and the OXE Receptor. Prostaglandins Other Lipid Mediat. 2009, 89, 98–104. [Google Scholar] [CrossRef]
  91. Sarveswaran, S.; Ghosh, J. OXER1, a G Protein-Coupled Oxoeicosatetraenoid Receptor, Mediates the Survival-Promoting Effects of Arachidonate 5-Lipoxygenase in Prostate Cancer Cells. Cancer Lett. 2013, 336, 185–195. [Google Scholar] [CrossRef]
  92. Kalyvianaki, K.; Gebhart, V.; Peroulis, N.; Panagiotopoulou, C.; Kiagiadaki, F.; Pediaditakis, I.; Aivaliotis, M.; Moustou, E.; Tzardi, M.; Notas, G.; et al. Antagonizing Effects of Membrane-Acting Androgens on the Eicosanoid Receptor OXER1 in Prostate Cancer. Sci. Rep. 2017, 7, 44418. [Google Scholar] [CrossRef] [PubMed]
  93. Pi, M.; Wu, Y.; Quarles, L.D. GPRC6A Mediates Responses to Osteocalcin in β-Cells in Vitro and Pancreas in Vivo. J. Bone Miner. Res. 2011, 26, 1680–1683. [Google Scholar] [CrossRef] [PubMed]
  94. Pi, M.; Parrill, A.L.; Quarles, L.D. GPRC6A Mediates the Non-Genomic Effects of Steroids. J. Biol. Chem. 2010, 285, 39953–39964. [Google Scholar] [CrossRef] [PubMed]
  95. Thomas, P. Membrane Androgen Receptors Unrelated to Nuclear Steroid Receptors. Endocrinology 2019, 160, 772–781. [Google Scholar] [CrossRef]
  96. Oury, F.; Sumara, G.; Sumara, O.; Ferron, M.; Chang, H.; Smith, C.E.; Hermo, L.; Suarez, S.; Roth, B.L.; Ducy, P.; et al. Endocrine Regulation of Male Fertility by the Skeleton. Cell 2011, 144, 796–809. [Google Scholar] [CrossRef]
  97. De Toni, L.; Guidolin, D.; De Filippis, V.; Tescari, S.; Strapazzon, G.; Rocca, M.S.; Ferlin, A.; Plebani, M.; Foresta, C. Osteocalcin and Sex Hormone Binding Globulin Compete on a Specific Binding Site of GPRC6A. Endocrinology 2016, 157, 4473–4486. [Google Scholar] [CrossRef] [PubMed]
  98. Pi, M.; Kapoor, K.; Wu, Y.; Ye, R.; Senogles, S.E.; Nishimoto, S.K.; Hwang, D.J.; Miller, D.D.; Narayanan, R.; Smith, J.C.; et al. Structural and Functional Evidence for Testosterone Activation of GPRC6A in Peripheral Tissues. Mol. Endocrinol. 2015, 29, 1759–1773. [Google Scholar] [CrossRef]
  99. Liu, M.; Zhao, Y.-Y.; Yang, F.; Wang, J.-Y.; Shi, X.-H.; Zhu, X.-Q.; Xu, Y.; Wei, D.; Sun, L.; Zhang, Y.-G.; et al. Evidence for a Role of GPRC6A in Prostate Cancer Metastasis Based on Case-Control and in Vitro Analyses. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 2235–2248. [Google Scholar]
  100. Pi, M.; Nishimoto, S.K.; Quarles, L.D. GPRC6A: Jack of All Metabolism (or Master of None). Mol. Metab. 2017, 6, 185–193. [Google Scholar] [CrossRef]
  101. Ye, R.; Pi, M.; Nooh, M.M.; Bahout, S.W.; Quarles, L.D. Human GPRC6A Mediates Testosterone-Induced Mitogen-Activated Protein Kinases and MTORC1 Signaling in Prostate Cancer Cells. Mol. Pharmacol. 2019, 95, 563–572. [Google Scholar] [CrossRef]
  102. Thomas, P.; Converse, A.; Berg, H.A. ZIP9, a Novel Membrane Androgen Receptor and Zinc Transporter Protein. Gen. Comp. Endocrinol. 2018, 257, 130–136. [Google Scholar] [CrossRef] [PubMed]
  103. Berg, A.H.; Rice, C.D.; Rahman, M.S.; Dong, J.; Thomas, P. Identification and Characterization of Membrane Androgen Receptors in the ZIP9 Zinc Transporter Subfamily: I. Discovery in Female Atlantic Croaker and Evidence ZIP9 Mediates Testosterone-Induced Apoptosis of Ovarian Follicle Cells. Endocrinology 2014, 155, 4237–4249. [Google Scholar] [CrossRef] [PubMed]
  104. Thomas, P.; Pang, Y.; Dong, J.; Berg, A.H. Identification and Characterization of Membrane Androgen Receptors in the ZIP9 Zinc Transporter Subfamily: II. Role of Human ZIP9 in Testosterone-Induced Prostate and Breast Cancer Cell Apoptosis. Endocrinology 2014, 155, 4250–4265. [Google Scholar] [CrossRef] [PubMed]
  105. Bulldan, A.; Bartsch, J.W.; Konrad, L.; Scheiner-Bobis, G. ZIP9 but Not the Androgen Receptor Mediates Testosterone-Induced Migratory Activity of Metastatic Prostate Cancer Cells. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 1857–1868. [Google Scholar] [CrossRef]
  106. Fessas, P.; Lee, H.; Ikemizu, S.; Janowitz, T. A Molecular and Preclinical Comparison of the PD-1–Targeted T-Cell Checkpoint Inhibitors Nivolumab and Pembrolizumab. Semin. Oncol. 2017, 44, 136–140. [Google Scholar] [CrossRef]
  107. Burtness, B.; Harrington, K.J.; Greil, R.; Soulières, D.; Tahara, M.; De Castro, G.; Psyrri, A.; Basté, N.; Neupane, P.; Bratland, Å.; et al. Pembrolizumab Alone or with Chemotherapy versus Cetuximab with Chemotherapy for Recurrent or Metastatic Squamous Cell Carcinoma of the Head and Neck (KEYNOTE-048): A Randomised, Open-Label, Phase 3 Study. Lancet 2019, 394, 1915–1928. [Google Scholar] [CrossRef]
  108. Leonard, B.; Brand, T.M.; O’keefe, R.A.; Lee, E.D.; Zeng, Y.; Kemmer, J.D.; Li, H.; Grandis, J.R.; Bhola, N.E. BET Inhibition Overcomes Receptor Tyrosine Kinase–Mediated Cetuximab Resistance in HNSCC. Cancer Res. 2018, 78, 4331–4343. [Google Scholar] [CrossRef]
  109. Viscuse, P.V.; Price, K.A.; Garcia, J.J.; Schembri-Wismayer, D.J.; Chintakuntlawar, A. V First Line Androgen Deprivation Therapy vs. Chemotherapy for Patients with Androgen Receptor Positive Recurrent or Metastatic Salivary Gland Carcinoma-A Retrospective Study. Front. Oncol. 2019, 9, 701. [Google Scholar] [CrossRef] [PubMed]
  110. 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. [Google Scholar] [CrossRef] [PubMed]
  111. Williams, L.; Thompson, L.D.R.; Seethala, R.R.; Weinreb, I.; Assaad, A.M.; Tuluc, M.; Ud Din, N.; 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] [CrossRef] [PubMed]
  112. Williams, M.D.; Roberts, D.; Blumenschein, G.R., Jr.; Temam, S.; Kies, M.S.; Rosenthal, D.I.; Weber, R.S.; El-Naggar, A.K. Differential Expression of Hormonal and Growth Factor Receptors in Salivary Duct Carcinomas: Biologic Significance and Potential Role in Therapeutic Stratification of Patients. Am. J. Surg. Pathol. 2007, 31, 1645–1652. [Google Scholar] [CrossRef]
  113. Boon, E.; van Boxtel, W.; Buter, J.; Baatenburg de Jong, R.J.; van Es, R.J.J.; Bel Bsc, 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] [CrossRef] [PubMed]
  114. Kamata, Y.; Sumida, T.; Murase, R.; Nakano, H.; Yamada, T.; Mori, Y. Blockade of Androgen-Induced Malignant Phenotypes by Flutamide Administration in Human Salivary Duct Carcinoma Cells. Anticancer. Res. 2016, 36, 6071–6075. [Google Scholar] [CrossRef]
  115. Locati, L.D.; Perrone, F.; Cortelazzi, B.; Imbimbo, M.; Bossi, P.; Potepan, P.; Civelli, E.; Rinaldi, G.; Quattrone, P.; Licitra, L.; et al. Activity of Abiraterone in Rechallenging Two AR-Expressing Salivary Gland Adenocarcinomas, Resistant to Androgen-Deprivation Therapy. Cancer Biol. Ther. 2014, 15, 678–682. [Google Scholar] [CrossRef] [PubMed]
  116. Soper, M.S.; Iganej, S.; Thompson, L.D.R. Definitive Treatment of Androgen Receptor-Positive Salivary Duct Carcinoma with Androgen Deprivation Therapy and External Beam Radiotherapy. Head Neck 2014, 36, E4–E7. [Google Scholar] [CrossRef] [PubMed]
  117. van Boxtel, W.; Verhaegh, G.W.; van Engen-van Grunsven, I.A.; van Strijp, D.; Kroeze, L.I.; Ligtenberg, M.J.; van Zon, H.B.; Hendriksen, Y.; Keizer, D.; van de Stolpe, A.; et al. Prediction of Clinical Benefit from Androgen Deprivation Therapy in Salivary Duct Carcinoma Patients. Int. J. Cancer 2020, 146, 3196–3206. [Google Scholar] [CrossRef] [PubMed]
  118. Jaspers, H.C.J.; Verbist, B.M.; Schoffelen, R.; Mattijssen, V.; Slootweg, P.J.; Van Der Graaf, W.T.A.; Van Herpen, C.M.L. Androgen Receptor-Positive Salivary Duct Carcinoma: A Disease Entity with Promising New Treatment Options. J. Clin. Oncol. 2011, e473–e476. [Google Scholar] [CrossRef]
  119. Locati, L.D.; Perrone, F.; Cortelazzi, B.; Lo Vullo, S.; Bossi, P.; Dagrada, G.; Quattrone, P.; Bergamini, C.; Potepan, P.; Civelli, E.; et al. Clinical Activity of Androgen Deprivation Therapy in Patients with Metastatic/Relapsed Androgen Receptor-Positive Salivary Gland Cancers. Head Neck 2016, 38, 724–731. [Google Scholar] [CrossRef] [PubMed]
  120. 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] [CrossRef] [PubMed]
  121. Rades, D.; Seibold, N.D.; Schild, S.E.; Gebhard, M.P.; Noack, F. Androgen Receptor Expression: Prognostic Value in Locally Advanced Squamous Cell Carcinoma of the Head and Neck. Strahlenther. Onkol. 2013, 189, 849–855. [Google Scholar] [CrossRef] [PubMed]
  122. Schmitt, N.C.; Kang, H.; Sharma, A. Salivary Duct Carcinoma: An Aggressive Salivary Gland Malignancy with Opportunities for Targeted Therapy. Oral Oncol. 2017, 74, 40–48. [Google Scholar] [CrossRef] [PubMed]
  123. Student, S.; Hejmo, T.; Poterała-Hejmo, A.; Leśniak, A.; Bułdak, R. Anti-Androgen Hormonal Therapy for Cancer and Other Diseases. Eur. J. Pharmacol. 2020, 866, 172783. [Google Scholar] [CrossRef] [PubMed]
  124. 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] [CrossRef]
  125. Urban, D.; Rischin, D.; Angel, C.; D’Costa, I.; Solomon, B. Abiraterone in Metastatic Salivary Duct Carcinoma. J. Natl. Compr. Canc. Netw. 2015, 13, 288–290. [Google Scholar] [CrossRef] [PubMed]
  126. Locati, L.D.; Cavalieri, S.; Bergamini, C.; Resteghini, C.; Colombo, E.; Calareso, G.; Mariani, L.; Quattrone, P.; Alfieri, S.; Bossi, P.; et al. Abiraterone Acetate in Patients with Castration-Resistant, Androgen Receptor–Expressing Salivary Gland Cancer: A Phase II Trial. J. Clin. Oncol. 2021, 39, 4061–4068. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The primary risk factors and annual incidence and mortality rates in both males and females for different subtypes of head and neck squamous cell carcinoma. The data are from Global Cancer Statistics 2020 [1]. Created with BioRender.com.
Figure 1. The primary risk factors and annual incidence and mortality rates in both males and females for different subtypes of head and neck squamous cell carcinoma. The data are from Global Cancer Statistics 2020 [1]. Created with BioRender.com.
Ijms 24 03766 g001
Figure 2. Nuclear and membrane androgen receptors and different types of antiandrogen therapy. The nuclear form of the androgen receptor (AR) is activated by binding of dihydrotestosterone (DHT), the 5α-reduced metabolite of testosterone (T), in the cytoplasm and then translocation into the nucleus. There, it serves as a transcription factor of many target genes that contain the androgen response element (ARE) in their promoters and enhancers. Many AR-regulated genes are related to carcinogenesis. Nonsteroidal antiandrogens block androgen-induced activation of AR and prevent its nuclear translocation and thus inhibit AR-mediated transcription in both the hypothalamus and target tissues. Steroidal antiandrogens, like abiraterone acetate, block enzymes involved in androgen biosynthesis, thereby stimulating the negative feedback loop in the hypothalamus, which results in lowering the plasma T concentration. Antiandrogens are often used in combination with a gonadotropin-releasing hormone (GnRH) agonist or luteinizing hormone-releasing hormone (LHRH). The currently known membrane androgen receptors are a group of five proteins: Ca2+ ion channels TRPM8 and CaV1.2, G-protein-coupled receptors OXER1 and GPRC6A, and zinc transporter ZIP9, which activate different signaling pathways through the second messengers. FSH—follicle-stimulating hormone; LH—luteinizing hormone. Created with BioRender.com.
Figure 2. Nuclear and membrane androgen receptors and different types of antiandrogen therapy. The nuclear form of the androgen receptor (AR) is activated by binding of dihydrotestosterone (DHT), the 5α-reduced metabolite of testosterone (T), in the cytoplasm and then translocation into the nucleus. There, it serves as a transcription factor of many target genes that contain the androgen response element (ARE) in their promoters and enhancers. Many AR-regulated genes are related to carcinogenesis. Nonsteroidal antiandrogens block androgen-induced activation of AR and prevent its nuclear translocation and thus inhibit AR-mediated transcription in both the hypothalamus and target tissues. Steroidal antiandrogens, like abiraterone acetate, block enzymes involved in androgen biosynthesis, thereby stimulating the negative feedback loop in the hypothalamus, which results in lowering the plasma T concentration. Antiandrogens are often used in combination with a gonadotropin-releasing hormone (GnRH) agonist or luteinizing hormone-releasing hormone (LHRH). The currently known membrane androgen receptors are a group of five proteins: Ca2+ ion channels TRPM8 and CaV1.2, G-protein-coupled receptors OXER1 and GPRC6A, and zinc transporter ZIP9, which activate different signaling pathways through the second messengers. FSH—follicle-stimulating hormone; LH—luteinizing hormone. Created with BioRender.com.
Ijms 24 03766 g002
Table 1. Currently known androgen receptors and their significance in head and neck squamous cell carcinoma.
Table 1. Currently known androgen receptors and their significance in head and neck squamous cell carcinoma.
Type of
Androgen
Receptor
ReceptorReceptor FunctionSignificance in HNSCCReference
NuclearARTranscription factorExpressed in 64% of LSCC cases; higher in well-differentiated, lower in poorly differentiated and those with lymphatic invasion[61]
AR mRNA upregulated in LSCC; AR+ nuclear staining observed in 77.1% of samples[62]
AR+ cytoplasmic staining in OSCC epithelium greater than 20% correlated with metastases as well as the frequency of VEGF positive lymphocytes and Ki67 expression in metastases[63,64]
DHT treatment increased proliferation, migration, and invasion of OSCC cells by increasing EGFR phosphorylation, which was reversed by AR knockdown or observed only in AR+ cells[66,67]
Lower expression in OSCC samples[68]
In male oral cavity or oropharyngeal cancer patients, a lower number of AR gene [CAG]n repeats (≤20) correlated with shorter disease-free survival and more frequent recurrence or metastasis[69]
MembraneTRPM8Calcium ion channelMenthol-induced Ca2+ influx through TRPM8 enhanced migration and invasion of OSCC cells via gelatinase activity of MMP-9[70]
Higher TRPM8 mRNA expression was a negative prognostic biomarker for HNSCC[71]
CaV1.2Calcium ion channelDecreased CACNA1C mRNA expression observed in OSCC cells; CACNA1C amplification and/or mRNA upregulation observed in 10% of HNSCC cases[72,73]
OXER1G-protein-coupled receptorUpregulated OXER1 mRNA expression observed in HPV+ oropharyngeal and oral cancers[74]
Upregulated OXER1 mRNA expression observed in human normal oral epithelial cells exposed to water-pipe smoking[75]
GPRC6AG-protein-coupled receptorUnknown/
ZIP9Zinc transporterUnknown/
AR+—androgen receptor-positive; DHT—dihydrotestosterone; HNSCC—head and neck squamous cell carcinoma; HPV+—human papillomavirus-positive; LSCC—laryngeal squamous cell carcinoma; OSCC—oral squamous cell carcinoma.
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

Čonkaš, J.; Sabol, M.; Ozretić, P. ‘Toxic Masculinity’: What Is Known about the Role of Androgen Receptors in Head and Neck Squamous Cell Carcinoma. Int. J. Mol. Sci. 2023, 24, 3766. https://doi.org/10.3390/ijms24043766

AMA Style

Čonkaš J, Sabol M, Ozretić P. ‘Toxic Masculinity’: What Is Known about the Role of Androgen Receptors in Head and Neck Squamous Cell Carcinoma. International Journal of Molecular Sciences. 2023; 24(4):3766. https://doi.org/10.3390/ijms24043766

Chicago/Turabian Style

Čonkaš, Josipa, Maja Sabol, and Petar Ozretić. 2023. "‘Toxic Masculinity’: What Is Known about the Role of Androgen Receptors in Head and Neck Squamous Cell Carcinoma" International Journal of Molecular Sciences 24, no. 4: 3766. https://doi.org/10.3390/ijms24043766

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