Next Article in Journal
Redox Implications of Extreme Task Performance: The Case in Driver Athletes
Next Article in Special Issue
Analyzing the Androgen Receptor Interactome in Prostate Cancer: Implications for Therapeutic Intervention
Previous Article in Journal
PTH-Induced Bone Regeneration and Vascular Modulation Are Both Dependent on Endothelial Signaling
Previous Article in Special Issue
Interaction between Non-Coding RNAs and Androgen Receptor with an Especial Focus on Prostate Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Androgen Receptor-Mediated Transcription in Prostate Cancer

by
Doğancan Özturan
1,2,†,
Tunç Morova
3,† and
Nathan A. Lack
1,2,3,*
1
School of Medicine, Koç University, Istanbul 34450, Turkey
2
Koç University Research Centre for Translational Medicine (KUTTAM), Koç University, Istanbul 34450, Turkey
3
Vancouver Prostate Centre, Department of Urologic Sciences, University of British Columbia, Vancouver, BC V6H 3Z6, Canada
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2022, 11(5), 898; https://doi.org/10.3390/cells11050898
Submission received: 17 January 2022 / Revised: 25 February 2022 / Accepted: 1 March 2022 / Published: 5 March 2022

Abstract

:
Androgen receptor (AR)-mediated transcription is critical in almost all stages of prostate cancer (PCa) growth and differentiation. This process involves a complex interplay of coregulatory proteins, chromatin remodeling complexes, and other transcription factors that work with AR at cis-regulatory enhancer regions to induce the spatiotemporal transcription of target genes. This enhancer-driven mechanism is remarkably dynamic and undergoes significant alterations during PCa progression. In this review, we discuss the AR mechanism of action in PCa with a focus on how cis-regulatory elements modulate gene expression. We explore emerging evidence of genetic variants that can impact AR regulatory regions and alter gene transcription in PCa. Finally, we highlight several outstanding questions and discuss potential mechanisms of this critical transcription factor.

1. Androgen Receptor in Prostate Cancer

Prostate cancer (PCa) is one of the leading causes of cancer-related death in men [1]. In almost all PCa patients, the androgen receptor (AR) is the primary driver of growth and differentiation [2]. Given this critical role, AR pathway inhibitors (ARPI) are the standard of care for treating patients with recurrent or metastatic forms of the disease [3,4]. However, while treatment is initially successful, ~20% of patients develop resistance and progress to a castration-resistant prostate cancer (CRPC) [5]. This aggressive form of the disease is invariably lethal. Intriguingly, the AR still remains active in the majority of resistant patients through various mechanisms, including AR point mutations [6], constitutively active AR variants [7], and most commonly, AR gene and enhancer amplification [8,9,10,11]. Despite the importance of AR-mediated transcription in PCa, fundamental aspects of how this nuclear receptor drives gene expression are only now being revealed. In this review, we summarize the mechanism of AR-mediated transcription in PCa and discuss outstanding questions.

2. Androgen Receptor-Mediated Gene Transcription

The AR is a 919 amino acids (110 kDa) protein that contains an N-terminal domain (NTD), DNA-binding domain (DBD), and C-terminal ligand-binding domain (LBD) [12,13]. The inactive apo-form of AR primarily resides in the cytoplasm, where it is stabilized by chaperone proteins. When activated by androgens, most commonly testosterone or the more potent metabolite 5α-dihydrotestosterone (DHT), the AR undergoes an allosteric modification, homodimerizes, and then translocates into the nucleus where it binds to DNA at AR binding sites (ARBS) [14]. The location of these ARBS is influenced by numerous features including the DNA primary sequence or motif, protein–protein interactions, transcription factor (TF) occupancy, and chromatin accessibility. Once bound to enhancer cis-regulatory elements (CREs), the AR recruits coregulators [15], remodeling complexes [16], and other TFs [17,18,19,20] to create a transcriptional hub that initiates an AR-dependent transcriptional program, which impacts the expression of several hundred target genes [21]. These genes contribute to proliferation, cellular differentiation, and potentially metastasis. The specific genes associated with these complex cellular processes are controversial, though several have been proposed, including c-Myc, EVT1, and EIF5A2 (eukaryotic translation initiation factor) [22,23,24]. When looking at essentiality from published genome-wide CRISPR screens of AR-regulated genes in PCa cells (LNCaP) (Figure 1), we find many known and novel essential genes, including coactivators such as GRHL2 (grainyhead-like transcription factor 2) [25]; metabolic genes such as DNM1L (dynamin-related protein 1) [26], SREBP (sterol regulatory element-binding protein) cleavage activating protein SCAP [27] and mTOR (the protein kinase mammalian target of rapamycin) [28]; and transcriptional regulators such as NFKBIA (NFKB inhibitor alpha), an inhibitor protein of NF-κB and p53 [29].

3. Pioneer Factors and DNA Binding

Before the AR binds to DNA, the ARBS is first “primed” by pioneer factor (PF) proteins that interact with heterochromatin and increase chromatin accessibility [30]. Given that PFs determine where the AR can potentially bind, these proteins strongly influence ARBS locations [31,32]. PFs engage with nucleosomes [33] that are dynamically transitioning between fully wrapped and transient exposure states [34,35]. The window of exposure is sufficient to allow binding of PFs as well as other TFs [36]. PFs typically bind to poised/active enhancers containing histone marks such as histone 3 lysine 4 mono/di-methylation (H3K4me1/me2) [37,38] and histone 3 lysine 9/27 acetylation (H3K9/27ac) [39,40]. There is limited PF binding at regions with strong silencing/repressive histone marks and DNA methylation [41,42]. While not reported with AR, other steroid receptors have been shown to recruit ATP-dependent modifiers that can interact with closed chromatin independent of PFs [43,44] and recruit these to target sites [45]. PFs are classified based on their protein domains and mechanism of action [45,46,47]. FOXA1 (forkhead box transcription factor a1) plays a critical role in the activity of AR and other nuclear receptors [48]. FOX proteins contain a winged helix DBD domain that allows recognition of full/partial DNA motifs in the nucleosome [49,50,51]. The C-terminal of FOXA1 is necessary for both unwrapping the chromatin through an ATP-independent mechanism and recruiting ATP-dependent chromatin modifiers [52]. In AR-mediated transcription, FOXA1 both opens chromatin for direct AR binding and also acts as an anchor for AR to indirectly bind CREs [53]. Demonstrating its critical role, knockdown of FOXA1 causes a decrease in overall AR binding with a massive redistribution of ARBS at tens of thousands of new sites [32]. This is strongly influenced by AR itself, as those sites that are unaffected following FOXA1 knockdown generally have higher AR occupancy [54]. GATA2 (GATA-binding factor 2) is another well-characterized PF that increases accessibility at ARBS [36,55]. Although GATA2 chromatin accessibility induction is less effective than FOXA1 [52], GATA proteins facilitate binding of AR and estrogen receptor (ER) in prostate [56] and breast cancer [57], respectively. Different from FOXA1, the GATA family of TFs requires other chromatin remodeling proteins, such as the SWI/SNF complex, to alter accessibility [45,58,59]. Lastly, HOXB13 (Homeobox b13), a member of the HOX family of proteins, has been speculated to have potential PF activity due to its preference to bind methylated DNA that is found at heterochromatin [60,61]. However, further work is needed to demonstrate this potential function. While the hierarchy of PFs is not clearly defined during AR activation, >70% of ARBS overlap with either GATA2, FOXA1, or HOXB13 binding sites [53,56,59,62]. Given that PF activity is critical for the oncogenic transformation of several cancers, including prostate and breast, there is ongoing research to develop therapeutics that target these PFs [63,64].

4. Plasticity of the AR Cistrome in Prostate Cancer Progression

AR induces gene transcription by binding to specific CREs in the genome. Identifying the location of ARBS is therefore critical to understand how AR functions. The first ARBS identified were found at the promoter of the rat probasin [65,66] and KLK3/PSA (kallikrein related peptidase 3/prostate specific antigen) gene [67,68,69]. The identification of subsequent ARBS, most prominently the AREIII (AR regulated enhancer III) upstream of KLK3/PSA, demonstrated that AR primarily drives transcription through enhancers rather than promoters [70,71]. Enhancers are non-coding regulatory elements that are required for gene expression, as they “enhance” transcription of target genes [72,73]. AR-driven enhancer activity has been confirmed with various genes including PSMA (prostate-specific membrane antigen) [74] and p21 [75]. Large-scale functional genomic studies, first with ChIP-on-chip (chromatin immunoprecipitation followed by microarray) and then ChIP-seq (chromatin immunoprecipitation followed by sequencing), provided additional support that AR activity occurred through enhancer CREs as these studies demonstrated that the vast majority of ARBS (>95%) are located in non-coding intronic or intergenic regions, with few binding sites found at promoters (<2%) [76,77,78].
The AR cistrome, or genome-wide binding sites, is not static, and clinical ARBS display remarkable plasticity and significant reprogramming during both tumor initiation and disease progression [79]. During neoplastic development, there is a dramatic expansion (3×) of ARBS in primary PCa compared to normal prostate [53]. Similarly, in metastatic CRPC samples, the AR gains an additional > 17,000 distinct binding sites that are associated with prostate development [79]. This suggests that CRPC regains an AR-driven early developmental transcription signature to potentially increase survival and proliferation during ARPI treatment. Interestingly, the changes in the AR cistrome seem unlikely to be solely due to chromatin accessibility as in both normal prostate and primary PCa most gained sites are already accessible euchromatin bound by PFs [79]. The reason why AR does not bind to these accessible regions in primary PCa remains an active area of research. While speculative, the increased expression of AR or coregulators in advanced stages of the PCa may influence where and for how long AR binds to chromatin. Supporting this potential mechanism, overexpression of AR in vitro has been shown to sensitize binding and alter ARBS [80]. Further, changes in FOXA1 expression alter the global chromatin accessibility and generate both pseudo-AR hypersensitivity and an increase in open chromatin ARBS [54]. This correlates with the broad increase in chromatin accessibility that is observed during the progression from primary PCa to CRPC [81,82]. Potentially, these newly accessible regions primed by FOXA1 or other coregulatory proteins may require increased expression of AR to compensate for the gained regions.

5. Impact of Motifs on AR Binding and Activity

ARBS are enriched for a conserved androgen response element (ARE) binding motif that is made up of two 6-bp asymmetrical elements separated by a 3-bp spacer (5′-AGAACAnnnTGTTCT-3′) [12,69,83]. This specific motif has been extensively validated by various studies, including an in vitro SELEX-seq (systematic evolution of ligands by exponential enrichment followed by sequencing) with recombinant AR-DBD protein [84]. Nonetheless, it remains unclear how important an ARE motif is to AR binding in situ [85]. Whole-genome AR ChIP-seq shows that only 8–30% of ARBS contain a canonical ARE, while most binding sites have a more common ARE half-motif (5′ -AGAACA-3′) [53,56,78,79,86]. However, the importance of ARE half-motifs is itself subject to discussion. A previous study found no evidence for binding at half-motif sites [87], whereas other work has proposed two modes of AR binding that include both half- and full-motif stabilized by FOXA1 interactions [30,88]. Further, there is contradicting data on the impact of motifs on enhancer activity. Studies with glucocorticoid receptor (GR) observed higher enhancer activity and target gene expression at those regions with GR motifs [89]. However, this was not binary, and many enhancers harboring weak GR motifs also had similar enhancer activity as those with strong motifs [89]. Our recent work did not report a strong correlation between any specific ARE motifs and AR enhancer activity [90]. These findings suggest that there is mixed evidence supporting the prerequisite of an ARE motif for AR binding and enhancer activity. Incorporating datasets such as chromosome accessibility (ATAC-seq, DNA-seq, FAIRE-seq) or alternative methodologies such as CUT&RUN may potentially help to reduce noise in defining in situ AR motifs.

6. Impact of Chromatin Modifying Enzymes on AR Activity

Once bound to DNA, the AR recruits chromatin-modifying enzymes that stabilize accessibility and provide a platform for coregulators and TFs [91]. This includes histone methyltransferases such as EZH2 [92,93,94], SET9 [95], and MLL complex (MLL, MLL4, WDR5, ASH2L) [96]; histone acetyltransferases (HAT) such as p160, SRC-1, TIF2/GRIP1-1, ACTR/AIB1/RAC3/pCIP [97,98,99,100,101], CBP [102], p300 [103], and pCAF [104]; and histone deacetylases (HDAC) such as HDAC1-3 (class I), HDAC4-10 (class II), and SIRT1-7 (class III) and HDAC11 (class IV) [105]. Characterization of the AREIII enhancer demonstrated the sequential recruitment of p160 and p300, was then followed by CBP and pCAF [21]. HAT and HDAC work antagonistically to acetylate the lysines on histone N-terminal tails that promote the formation of heterochromatin through electrostatic interactions [106]. Further, BRM (SMARCA2), a member of the SWI/SNF chromatin remodeling complex, has been shown to be essential for KLK3/PSA and rat probasin expression [107]. Overexpression of the related BRG1 (SMARCA4) in BRM/BRG1 mutant cell lines showed limited AR activity at KLK3/PSA and rat probasin enhancers [108]. In ER-mediated transcription, FOXA1 has also been shown to recruit MLL3 to promote H3K4me3 at ER binding sites in breast cancer cell lines [109]. The kinetic hierarchy of regulation still needs further characterization given the intricacies of this dynamic transcriptional complex.

7. AR-Coregulators and Gene Transcription

Once bound to chromatin, AR forms a complex with numerous coregulatory proteins to activate enhancers and alter the expression of target genes (Figure 2) [89,110,111,112]. This process involves many different proteins, including coactivators such as CBP/p300 (CREB binding protein) and SRC-1 [21]; chromatin modifying enzymes that alter accessibility [108]; and proteins that stabilize AR binding [16,113]. Overall, this is an extremely dynamic process, and to date >250 proteins have been identified that interact with the AR and have potential coregulatory activity [114]. Further, AR interacts with MED1 (mediator complex subunit 1) and other members of the Mediator complex to stabilize AR-mediated enhancer–promoter interactions [17]. Recent work has suggested that these AR–MED1 interactions may induce the formation of phase condensates at super-enhancer regulatory elements [115]. These super-enhancers have higher transcriptional output of target genes than individual enhancers and play a pivotal role in cell identity and tumor progression [116]. In CRPC, several gained super-enhancers were proposed to activate oncogenes such as CHPT1 (choline phosphotransferase 1) and drive resistance [117]. In addition to these coregulatory proteins, long non-coding RNAs (lncRNAs) and enhancer RNAs have also been reported to impact AR-mediated gene transcription [8,118]. Although this is a current subject of research, these RNA species are proposed to recruit protein complexes to the transcription target sites [119]. For example, the PCA3 (prostate cancer antigen 3) lncRNA interacts with AR and stabilizes androgen-induced transcription [120]. However, there is considerable diversity in the mechanism of action. AR-upregulated ARLNC1 (androgen receptor regulated long noncoding RNA 1) has been shown to stabilize AR mRNA transcripts and alter expression through a transcriptional feedback loop [121]. Further, SChLAP1 (second chromosome locus-associated with prostate-1) is highly expressed in CRPC and induces proliferation through interactions with the SWI/SNF chromatin-modifying complex [122].
While poorly understood, AR-mediated gene downregulation has been proposed to occur through both direct repression and indirect coactivator sequestering, also known as squelching [123]. In direct repression, corepressor proteins such as homologous proteins SMRT and NCoR (nuclear receptor corepressor) interact with AR and recruit histone deacetylases such as HDAC4 that cause chromatin compaction and transcriptional repression (Figure 3) [124]. Specific corepressors include RIP140 (receptor-interacting protein 140), which directly binds to C-terminus of AR protein [125]. Others such as LCoR (ligand dependent corepressor), inhibit AR-mediated transcription by interacting with HDACs and CtBP (C-terminal binding protein), which suppress tumor growth in vivo [126,127]. Calcium-binding protein calreticulin inhibits AR activity through its DBD [128,129]. In contrast with these mechanisms, the squelching model of repression proposes that the activated AR, which previously resided in the cytoplasm, binds to numerous coregulatory proteins that impact the activity of non-AR TFs by limiting access to these critical proteins. AR and other nuclear receptors are also the subject of auto-squelching, which can repress target genes [130].
Regardless of the mechanism, gene downregulation is believed to be important for the growth and progression of advanced stage PCa [131]. Importantly, the activation of AR leads to the downregulation of c-Myc [132], which has an antagonistic transcriptional network with AR [133]. c-Myc repression by AR is largely independent of AR binding to its target sites and primarily occurs via the redistribution of AR coactivators [134]. Further, c-Myc regulation by histone methyltransferase, DOT1L (disruptor of telomeric silencing 1-like), and AR through an enhancer has also been reported [135]. Decreased AR expression upon the inhibition of DOT1L, which coregulates AR and MYC pathways, leads to increased expression of AR-target genes by other TFs such as c-Myc. AR has also been shown to alter the expression of other known tumor-suppressors such as p53, PTEN (phosphatase and tensin homolog deleted on chromosome 10), and LRIG (leucine-rich repeats and immunoglobulin-like domains). AR inhibits p53 expression, while p53 directly represses the expression of AR by binding to target promoters [136,137]. PI3K (phosphoinositide3-kinase) signaling is altered in PCa through loss of PTEN and is associated with aggressive PCa prognosis [138]. Expression of PTEN is inversely correlated with AR in PCa tumors, and AR is reported to directly inhibit PTEN expression [139]. Finally, elevated expression of the AR-stimulated tumor-suppressor LRIG is associated with increased overall survival in PCa cohorts [140,141]. LRIG expression is also affected by SUMOylation of AR in which small ubiquitin-related modifiers (SUMOs) covalently bind to the AR and alter the downstream transcription events [142]. Interestingly, motif analysis of corepressor-bound AR and coactivator-bound AR showed a similar binding motif, which suggests there may be competition between these two complexes for gene activation/repression [143].

8. AR Enhancers in Gene Transcription

AR primarily drives gene expression through enhancer CREs. Located at euchromatin [144], enhancers commonly correlate with histone marks such as H3K27Ac and H3K4me1 [145]. Enhancers are also typically bound by multiple TFs [89,90,146,147,148], RNA polymerase II [149], transcriptional coactivators [150,151,152], and CEBP/p300 [153]. Nonetheless, these features only broadly correlate with activity, and there are numerous enhancer CREs that do not contain some or any of these specific modifications [90,154,155,156]. Functional annotation is needed to understand how ARBS work together to drive gene transcription. This is particularly important in AR-mediated transcription, as there are 10–100× more ARBS (tens of thousands) than differentially expressed genes (hundreds). The development of novel high-throughput enhancer assays such as STARR-seq (self-transcribing active regulatory region sequencing) have enabled researchers to test the enhancer activities of thousands of genomic regions in a single experiment [157,158,159,160]. Recently, all high-confidence clinical ARBS were tested using STARR-seq and revealed three different classes of binding sites—named as inducible, inactive, and constitutive enhancers. Only a fraction of the regions showed AR-activated or inducible enhancer activity (7%), and instead the majority of ARBS did not demonstrate any enhancer activity (81%). Further, approximately 12% of ARBS exhibited constitutive enhancer activity that was independent of AR binding. Inducible AR enhancers were found to correlate with both high-AR occupancy and an increase in chromatin loops to other CREs and gene promoters. While it could be argued that these differences in activity are either contextually or temporally dependent, a strong correlation was observed between these in vitro annotations and H3K27Ac in clinical PCa samples. Unexpectedly, when these ARBS classes were functionally tested, both inactive and constitutively active enhancers, in addition to inducible enhancers, were frequently required for AR-mediated gene transcription. Supporting this functional role, each ARBS class demonstrated equivalent evolutionary conservation, suggesting that each enhancer type is required for gene transcription [90]. Apart from AR inducible enhancers, different mechanisms have been proposed for constitutive and inactive ARBS enhancers. Inactive sites could support long-range chromatin interactions or increase the local AR concentration to trigger gene transcription. A similar stabilization may occur with constitutive enhancers where multiple TFs support looping following AR binding. In this model, these genomic regions would produce enhancer reporter signals in an episomal assay with non-AR TF binding but would only contribute to gene transcription when the AR is bound. Large-scale functional testing of these distinct ARBS enhancer classes is needed to interrogate their role in AR-mediated gene transcription.
There is increasing evidence that multiple ARBS work together to drive gene transcription. More than 60% of the AR-regulated genes have >2 ARBS within 200 kb proximity (Figure 4a). For example, TMPRSS2 (transmembrane serine protease 2) is regulated by several enhancers bound by AR/FOXA1/p300 that loop to the gene promoter and induce transcription [161]. How this occurs is poorly understood, but several distinct biological models have been proposed to explain the collaborative mechanism between multiple bound TFs [162,163]. Recent studies of ER binding sites have shown both hierarchical and synergistic interactions between enhancers [164]. The hierarchical model suggests that a dominant motif-containing enhancer can activate gene expression by itself, and a nearby weak motif-containing enhancer only contributes secondarily to gene activation. In contrast, the synergistic model proposes that a motif-containing binding site only contributes to gene expression if a neighboring binding site is also bound. However, it remains to be determined whether a single model can explain all interactions between different ARBS enhancers.

9. 3D Genome Organization

AR-regulated enhancers are brought in close physical proximity to the target gene promoter by chromatin looping [165]. These ARBS enhancer–promoter loops occur within topologically associated domains (TADs) that are formed by both insulator protein CTCF (CCCTC-binding factor) and cohesin [166]. TADs segment the genome into regions that contain high contact frequency loops and similar histone modification patterns [167,168]. Within each TAD, most ARBS loops are distributed between 10 kb to 1 mb (Figure 4b), though there are notable exceptions, including the gene STEAP4 (six-transmembrane epithelial antigen of prostate 4) that interacts with an ARBS found >2 MB from the target promoter [169]. Enhancer–promoter loops in PCa cell lines are enriched for numerous binding motifs, including AR, FOXA1, and the coregulator GRHL2 [170]. These enhancer–promoter interactions are proposed to either be pre-existing or formed de novo [171,172]. Pre-existing links are convenient for rapid transcriptional activation [173], whereas de novo loops can be formed through TF interacting structural proteins such as YY1 (Yin Yang 1) [174,175]. Gene expression and loop strength are independent of the distance between an enhancer and its target promoter [176]. Work from the ENCODE project demonstrated that the average distance of enhancer–promoter loops is around 120 kb with almost four enhancers for any given active gene [177]. However, these regulatory networks are complicated by the significant alterations found in PCa at almost all hierarchical levels of chromosomal organization [178]. A recent study conducted HiC, a whole genome chromosome conformation capture assay, in multiple PCa cell lines, including RWPE-1, LNCaP, DU145, 22Rv1, VCaP, PC3, MDAPCa2a, MDAPCa2b, and C4-2B, identified 387 TAD gene compartments that were distinct for each cell line [179]. Similarly, in situ HiC maps of RWPE1, C4-2B, and 22Rv1 cell lines showed that common TADs found in all cell lines were much smaller than those TADs unique to a single cell line [170]. Using “normal” prostate epithelial cells and models of PCa (LNCaP and PC3), those so-called “normal” TADs were much larger, higher in number, and located in distinct positions [180]. However, low-input HiC from both primary PCa (n = 12) and benign prostate tissues (n = 5) demonstrated that, unlike cell lines, there was no significant difference in the number of TADs called or in TAD borders between samples [181]. Combining these data with whole-genome sequencing, they found only one structural variant (out of 260) with altered gene expression in an intra-TAD region. While spatial organization of the genome affects gene transcription, it remains to be determined how these changes in chromatin looping affect AR-mediated gene expression.

10. Enhancer CRE Mutations

PCa has a relatively low somatic mutation frequency compared with other cancer types [184]. Common oncogenic drivers include TP53 (tumor protein 53) and PTEN (phosphatase and tensin homolog), as well as prostate-specific recurring mutations such as SPOP (speckle type BTB/POZ protein) and FOXA1 [185]. Given the critical role of AR signaling in CRPC, late-stage PCa commonly harbors AR somatic mutations including gene duplications, single nucleotide variants (SNV), or structural variants (SV) [10]. Several excellent reviews have discussed protein coding mutations in PCa and their role to stratify patients for treatment [186,187,188]. However, protein coding regions make up only ~1% of the whole genome [189,190,191], and there is increasing evidence that non-coding mutations at enhancer CREs contribute to PCa progression. This is particularly important as while the mutational burden is relatively low in PCa, there is a high frequency of SVs that can cause enhancer-driven dysregulation of transcriptional networks [192]. For instance, duplication of an upstream enhancer that regulates the AR gene is commonly found in most advanced PCa patients (81%) and can act as the sole driver of ARPI-resistance in CRPC [193,194]. Further, there are several common fusion events where AR-driven regulatory elements induce transcription of oncogenic driver genes through enhancer hijacking [195,196,197]. This was first reported with TMPRSS2-ERG fusions [198] and then later TMPRSS2-ETV1/ETV4 fusions [199]. TMPRSS2 is a prostate-specific AR-regulated gene, whereas ERG is a critical regulator of proliferation, differentiation, and apoptosis [198,200]. As a result of the fusion, the ERG gene expression becomes regulated by AR signaling and is highly expressed in PCa. Similar complex rearrangements between the AR-regulated gene NRF1 and BRAF have also been observed [201]. n-Myc and c-Myc expression have also been attributed to enhancer hijacking of distal enhancers in neuroblastoma [202,203]. Changes in gene expression can also occur by mutations that alter TAD structures. A commonly found deletion in the 17p13.1 locus that contains the tumor-suppressor p53 gene separates a well-defined TAD that occurs in normal cells into two distinct TADs, with significant changes in CRE usage [180]. Further, a recent study demonstrated that disruption of a single CTCF binding site in the KLK locus alters transcription of the gene cluster [204]. However, it is not known exactly which enhancer regions play a role in this activation. Further, this potentially may be locus-specific, as there is evidence that CTCF depletion is not affecting enhancer–promoter connections [205]. Large-scale chromosomal alterations can also cause circular extrachromosomal DNA (ecDNA) that leads to the expression of numerous oncogenes through changes in enhancer usage [206,207]. Given the highly unstable genomic landscape of PCa, ecDNA is increasingly being found [208,209]. Further, enhancer retargeting caused by promoter somatic mutations can also lead to gene reactivation [210].
There is conflicting evidence for the role of non-coding SNVs in PCa initiation and progression. The vast majority of PCa point mutations are non-coding and found in intergenic (46%), intronic (44%), and promoter (9%) regions [211]. Most of these SNVs are likely passenger mutations. In the large-scale Pan-Cancer Analysis of Whole Genomes (PCAWG), only 0.3% (986 of 276,892) of patients had recurrent non-coding mutations, suggesting that there is little selective pressure [211]. However, this interpretation is complicated by the nature of gene transcription, where multiple CREs commonly work together and a mutation at independent enhancer regions could potentially cause the same alteration in gene expression [212]. Therefore, instead of recurrent individual mutations causing transcriptional dysregulation, they could occur at multiple sites in a local regulatory network (plexuses) and alter the expression of critical genes [211]. Supporting a potential role of CRE SNVs in PCa development, the majority of single nucleotide polymorphisms (SNPs) identified from PCa-associated genome-wide association studies (GWAS) occur in non-coding regulatory regions and are proposed to alter TF binding at enhancer regions [213,214,215]. While the impact of these mutations is unclear, we and others have demonstrated that the binding sites of lineage-specific TFs have an increased rate of somatic mutations [211,216]. However, interpretation of these SNVs is limited by poor understanding of enhancer “grammar” that prevents the identification of potential pathogenic variants.

11. Targeting AR and Coregulators

Current PCa therapies target AR through either direct antagonism (bicalutamide, enzalutamide, apalutamide, etc.) or by reducing the synthesis of androgenic steroids (LHRH agonists, abiraterone, etc.) [217]. However, while initially effective, almost all tumors eventually develop resistance to treatment [5]. While a subset of these resistant tumors differentiates into a neuroendocrine state (<15%), the vast majority of CRPC tumors still remain dependent on AR signaling. Given their critical function, AR-coactivator interactions have been proposed as an alternative pharmacological target for overcoming many common mechanisms of resistance [218].

12. Conclusions

In this brief review, we discuss the mechanism of AR-regulated gene expression in PCa. Numerous studies characterizing the AR cistrome have revealed that the vast majority of ARBS are located at enhancer CREs that regulate the transcriptional activity via chromatin looping. However, while these ARBS are well characterized, there are still many outstanding questions, particularly, related to the expansion of ARBS in CRPC where there is a broad reactivation of early developmental transcriptional processes. There is emerging evidence that ARBS can influence gene transcription even without episomal enhancer activity, suggesting that AR directly or via other TFs can potentially stabilize CRE chromatin interactions and influence transcription. From this perspective, there is a need for additional chromosomal genome organization datasets to improve our understanding of phenotypic events in the different stages of PCa. With high-throughput dataset initiatives such as ENCODE [191,219] and 4D-Nucleome Project [220], as well as several large-scale clinical projects, these datasets will help to contribute to our knowledge of this complex process. By understanding AR-mediated gene transcription, we can both begin to stratify potential non-coding driver mutations and identify therapeutic vulnerabilities to better treat late-stage PCa patients.

Author Contributions

D.Ö., T.M., N.A.L. have contributed to the structuring, writing, and editing of this article. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Scientific and Technological Research Council of Turkey (TUBITAK 1001; 119Z279) and Department of Defense Prostate Cancer Research Program Idea Development Award (PC200674).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge the use of the services and facilities of the Koç University Research Center for Translational Medicine (KUTTAM), funded by the Presidency of Turkey, Presidency of Strategy and Budget.

Conflicts of Interest

The authors declare no conflict of interest.

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 Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. Lamont, K.R.; Tindall, D.J. Androgen Regulation of Gene Expression. Adv. Cancer Res. 2010, 107, 137–162. [Google Scholar] [CrossRef] [PubMed]
  3. 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. Cancer Res. 1941, 1, 293–297. [Google Scholar]
  4. Crawford, E.D.; Hou, A.H. The role of LHRH antagonists in the treatment of prostate cancer. Oncology 2009, 23, 626–630. [Google Scholar]
  5. Kirby, M.; Hirst, C.; Crawford, E.D. Characterising the castration-resistant prostate cancer population: A systematic review. Int. J. Clin. Pract. 2011, 65, 1180–1192. [Google Scholar] [CrossRef] [PubMed]
  6. Hay, C.W.; McEwan, I.J. The Impact of Point Mutations in the Human Androgen Receptor: Classification of Mutations on the Basis of Transcriptional Activity. PLoS ONE 2012, 7, e32514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. 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] [CrossRef]
  8. Zhang, A.; Zhao, J.C.; Kim, J.; Fong, K.-W.; Yang, Y.A.; Chakravarti, D.; Mo, Y.-Y.; Yu, J. LncRNA HOTAIR Enhances the Androgen-Receptor-Mediated Transcriptional Program and Drives Castration-Resistant Prostate Cancer. Cell Rep. 2015, 13, 209–221. [Google Scholar] [CrossRef] [Green Version]
  9. Barbieri, C.; Baca, S.C.; Lawrence, M.S.; Demichelis, F.; Blattner, M.; Theurillat, J.-P.; White, T.A.; Stojanov, P.; Van Allen, E.; Stransky, N.; et al. Exome sequencing identifies recurrent SPOP, FOXA1 and MED12 mutations in prostate cancer. Nat. Genet. 2012, 44, 685–689. [Google Scholar] [CrossRef] [Green Version]
  10. Robinson, D.; Van Allen, E.M.; Wu, Y.-M.; Schultz, N.; Lonigro, R.J.; Mosquera, J.-M.; Montgomery, B.; Taplin, M.-E.; Pritchard, C.C.; Attard, G.; et al. Integrative Clinical Genomics of Advanced Prostate Cancer. Cell 2015, 162, 454. [Google Scholar] [CrossRef] [Green Version]
  11. Taylor, B.S.; Schultz, N.; Hieronymus, H.; Gopalan, A.; Xiao, Y.; Carver, B.S.; Arora, V.K.; Kaushik, P.; Cerami, E.; Reva, B.; et al. Integrative Genomic Profiling of Human Prostate Cancer. Cancer Cell 2010, 18, 11–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Mangelsdorf, D.J.; Thummel, C.; Beato, M.; Herrlich, P.; Schütz, 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]
  13. Chang, C.; Saltzman, A.; Yeh, S.; Young, W.; Keller, E.; Lee, H.-J.; Wang, C.; Mizokami, A. Androgen Receptor: An Overview. Crit. Rev. Eukaryot. Gene Expr. 1995, 5, 97–125. [Google Scholar] [CrossRef] [PubMed]
  14. Davey, R.A.; Grossmann, M. Androgen Receptor Structure, Function and Biology: From Bench to Bedside. Clin. Biochem. Rev. 2016, 37, 3–15. [Google Scholar]
  15. Heemers, H.V.; Tindall, D.J. Androgen Receptor (AR) Coregulators: A Diversity of Functions Converging on and Regulating the AR Transcriptional Complex. Endocr. Rev. 2007, 28, 778–808. [Google Scholar] [CrossRef] [Green Version]
  16. Menon, T.; Yates, J.A.; Bochar, D.A. Regulation of Androgen-Responsive Transcription by the Chromatin Remodeling Factor CHD8. Mol. Endocrinol. 2010, 24, 1165–1174. [Google Scholar] [CrossRef] [Green Version]
  17. Jin, F.; Claessens, F.; Fondell, J.D. Regulation of Androgen Receptor-dependent Transcription by Coactivator MED1 Is Mediated through a Newly Discovered Noncanonical Binding Motif. J. Biol. Chem. 2012, 287, 858–870. [Google Scholar] [CrossRef] [Green Version]
  18. Tan, P.Y.; Chang, C.W.; Chng, K.R.; Wansa, K.S.A.; Sung, W.-K.; Cheung, E. Integration of Regulatory Networks by NKX3-1 Promotes Androgen-Dependent Prostate Cancer Survival. Mol. Cell. Biol. 2012, 32, 399–414. [Google Scholar] [CrossRef] [Green Version]
  19. Meyer, R.; Wolf, S.S.; Obendorf, M. PRMT2, a member of the protein arginine methyltransferase family, is a coactivator of the androgen receptor. J. Steroid Biochem. Mol. Biol. 2007, 107, 1–14. [Google Scholar] [CrossRef]
  20. Vélot, L.; Lessard, F.; Bérubé-Simard, F.-A.; Tav, C.; Neveu, B.; Teyssier, V.; Boudaoud, I.; Dionne, U.; Lavoie, N.; Bilodeau, S.; et al. Proximity-dependent Mapping of the Androgen Receptor Identifies Kruppel-like Factor 4 as a Functional Partner. Mol. Cell. Proteom. 2021, 20, 100064. [Google Scholar] [CrossRef]
  21. Shang, Y.; Myers, M.; Brown, M. Formation of the Androgen Receptor Transcription Complex. Mol. Cell 2002, 9, 601–610. [Google Scholar] [CrossRef]
  22. Gao, L.; Schwartzman, J.; Gibbs, A.; Lisac, R.; Kleinschmidt, R.; Wilmot, B.; Bottomly, D.; Coleman, I.; Nelson, P.; McWeeney, S.; et al. Androgen Receptor Promotes Ligand-Independent Prostate Cancer Progression through c-Myc Upregulation. PLoS ONE 2013, 8, e63563. [Google Scholar] [CrossRef] [Green Version]
  23. Cai, C.; Hsieh, C.-L.; Omwancha, J.; Zheng, Z.; Chen, S.-Y.; Baert, J.-L.; Shemshedini, L. ETV1 Is a Novel Androgen Receptor-Regulated Gene that Mediates Prostate Cancer Cell Invasion. Mol. Endocrinol. 2007, 21, 1835–1846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zheng, Y.; Li, P.; Huang, H.; Ye, X.; Chen, W.; Xu, G.; Zhang, F. Androgen receptor regulates eIF5A2 expression and promotes prostate cancer metastasis via EMT. Cell Death Discov. 2021, 7, 373. [Google Scholar] [CrossRef] [PubMed]
  25. Paltoglou, S.; Das, R.; Townley, S.L.; Hickey, T.; Tarulli, G.; Coutinho, I.; Fernandes, R.; Hanson, A.R.; Denis, I.; Carroll, J.; et al. Novel Androgen Receptor Coregulator GRHL2 Exerts Both Oncogenic and Antimetastatic Functions in Prostate Cancer. Cancer Res. 2017, 77, 3417–3430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Lee, Y.G.; Nam, Y.; Shin, K.J.; Yoon, S.; Park, W.S.; Joung, J.Y.; Seo, J.K.; Jang, J.; Lee, S.; Nam, D.; et al. Androgen-induced expression of DRP1 regulates mitochondrial metabolic reprogramming in prostate cancer. Cancer Lett. 2020, 471, 72–87. [Google Scholar] [CrossRef] [PubMed]
  27. Yu, I.-C.; Lin, H.-Y.; Sparks, J.D.; Yeh, S.; Chang, C. Androgen Receptor Roles in Insulin Resistance and Obesity in Males: The Linkage of Androgen-Deprivation Therapy to Metabolic Syndrome. Diabetes 2014, 63, 3180–3188. [Google Scholar] [CrossRef] [Green Version]
  28. Wu, Y.; Chhipa, R.R.; Cheng, J.; Zhang, H.; Mohler, J.L.; Ip, C. Androgen receptor-mTOR crosstalk is regulated by testosterone availability: Implication for prostate cancer cell survival. Anticancer Res. 2010, 30, 3895–3901. [Google Scholar]
  29. Carter, S.L.; Centenera, M.M.; Tilley, W.D.; Selth, L.A.; Butler, L.M. IκBα mediates prostate cancer cell death induced by combinatorial targeting of the androgen receptor. BMC Cancer 2016, 16, 141. [Google Scholar] [CrossRef] [Green Version]
  30. Tewari, A.K.; Yardimci, G.G.; Shibata, Y.; Sheffield, N.C.; Song, L.; Taylor, B.S.; Georgiev, S.G.; Coetzee, G.A.; Ohler, U.; Furey, T.S.; et al. Chromatin accessibility reveals insights into androgen receptor activation and transcriptional specificity. Genome Biol. 2012, 13, R88. [Google Scholar] [CrossRef] [Green Version]
  31. Lupien, M.; Eeckhoute, J.; Meyer, C.A.; Wang, Q.; Zhang, Y.; Li, W.; Carroll, J.S.; Liu, X.S.; Brown, M. FoxA1 Translates Epigenetic Signatures into Enhancer-Driven Lineage-Specific Transcription. Cell 2008, 132, 958–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Sahu, B.; Laakso, M.; Ovaska, K.; Mirtti, T.; Lundin, J.; Rannikko, A.; Sankila, A.; Turunen, J.-P.; Lundin, M.; Konsti, J.; et al. Dual role of FoxA1 in androgen receptor binding to chromatin, androgen signalling and prostate cancer. EMBO J. 2011, 30, 3962–3976. [Google Scholar] [CrossRef]
  33. Garcia, M.F.; Moore, C.D.; Schulz, K.N.; Alberto, O.; Donague, G.; Harrison, M.M.; Zhu, H.; Zaret, K.S. Structural Features of Transcription Factors Associating with Nucleosome Binding. Mol. Cell 2019, 75, 921–932.e6. [Google Scholar] [CrossRef] [PubMed]
  34. Polach, K.; Widom, J. Mechanism of Protein Access to Specific DNA Sequences in Chromatin: A Dynamic Equilibrium Model for Gene Regulation. J. Mol. Biol. 1995, 254, 130–149. [Google Scholar] [CrossRef] [Green Version]
  35. Anderson, J.; Widom, J. Sequence and position-dependence of the equilibrium accessibility of nucleosomal DNA target sites. J. Mol. Biol. 2000, 296, 979–987. [Google Scholar] [CrossRef]
  36. Sunkel, B.D.; Stanton, B.Z. Pioneer factors in development and cancer. iScience 2021, 24, 103132. [Google Scholar] [CrossRef] [PubMed]
  37. Strahl, B.D.; Ohba, R.; Cook, R.G.; Allis, C.D. Methylation of histone H3 at lysine 4 is highly conserved and correlates with transcriptionally active nuclei in Tetrahymena. Proc. Natl. Acad. Sci. USA 1999, 96, 14967–14972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Wang, J.; Zhuang, J.; Iyer, S.; Lin, X.; Whitfield, T.W.; Greven, M.C.; Pierce, B.G.; Dong, X.; Kundaje, A.; Cheng, Y.; et al. Sequence features and chromatin structure around the genomic regions bound by 119 human transcription factors. Genome Res. 2012, 22, 1798–1812. [Google Scholar] [CrossRef] [Green Version]
  39. Struhl, K. Histone acetylation and transcriptional regulatory mechanisms. Genes Dev. 1998, 12, 599–606. [Google Scholar] [CrossRef] [Green Version]
  40. Creyghton, M.P.; Cheng, A.W.; Welstead, G.G.; Kooistra, T.; Carey, B.W.; Steine, E.J.; Hanna, J.; Lodato, M.A.; Frampton, G.M.; Sharp, P.A.; et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc. Natl. Acad. Sci. USA 2010, 107, 21931–21936. [Google Scholar] [CrossRef] [Green Version]
  41. Rougeulle, C.; Chaumeil, J.; Sarma, K.; Allis, C.D.; Reinberg, D.; Avner, P.; Heard, E. Differential Histone H3 Lys-9 and Lys-27 Methylation Profiles on the X Chromosome. Mol. Cell. Biol. 2004, 24, 5475–5484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Iwafuchi-Doi, M.; Zaret, K.S. Cell fate control by pioneer transcription factors. Development 2016, 143, 1833–1837. [Google Scholar] [CrossRef] [Green Version]
  43. Ballare, C.; Castellano, G.; Gaveglia, L.; Althammer, S.; González-Vallinas, J.; Eyras, E.; Le Dily, F.; Zaurin, R.; Soronellas, D.; Vicent, G.P.; et al. Nucleosome-Driven Transcription Factor Binding and Gene Regulation. Mol. Cell 2013, 49, 67–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Swinstead, E.E.; Miranda, T.B.; Paakinaho, V.; Baek, S.; Goldstein, I.; Hawkins, M.; Karpova, T.; Ball, D.; Mazza, D.; Lavis, L.; et al. Steroid Receptors Reprogram FoxA1 Occupancy through Dynamic Chromatin Transitions. Cell 2016, 165, 593–605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Swinstead, E.E.; Paakinaho, V.; Presman, D.M.; Hager, G.L. Pioneer factors and ATP-dependent chromatin remodeling factors interact dynamically: A new perspective: Multiple transcription factors can effect chromatin pioneer functions through dynamic interactions with ATP-dependent chromatin remodeling factors. Bioessays 2016, 38, 1150–1157. [Google Scholar] [CrossRef] [PubMed]
  46. Zaret, K.S.; Carroll, J.S. Pioneer transcription factors: Establishing competence for gene expression. Genes Dev. 2011, 25, 2227–2241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Hankey, W.; Chen, Z.; Wang, Q. Shaping Chromatin States in Prostate Cancer by Pioneer Transcription Factors. Cancer Res. 2020, 80, 2427–2436. [Google Scholar] [CrossRef] [Green Version]
  48. Teng, M.; Zhou, S.; Cai, C.; Lupien, M.; He, H.H. Pioneer of prostate cancer: Past, present and the future of FOXA1. Protein Cell 2021, 12, 29–38. [Google Scholar] [CrossRef]
  49. Clark, K.L.; Halay, E.D.; Lai, E.; Burley, S.K. Co-crystal structure of the HNF-3/fork head DNA-recognition motif resembles histone H5. Nature 1993, 364, 412–420. [Google Scholar] [CrossRef]
  50. Cirillo, L.A.; McPherson, C.E.; Bossard, P.; Stevens, K.; Cherian, S.; Shim, E.Y.; Clark, K.L.; Burley, S.; Zaret, K.S. Binding of the winged-helix transcription factor HNF3 to a linker histone site on the nucleosome. EMBO J. 1998, 17, 244–254. [Google Scholar] [CrossRef] [Green Version]
  51. Iwafuchi, M.; Donahue, G.; Kakumanu, A.; Watts, J.A.; Mahony, S.; Pugh, B.F.; Lee, D.; Kaestner, K.H.; Zaret, K.S. The Pioneer Transcription Factor FoxA Maintains an Accessible Nucleosome Configuration at Enhancers for Tissue-Specific Gene Activation. Mol. Cell 2016, 62, 79–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Cirillo, L.A.; Lin, F.R.; Cuesta, I.; Friedman, D.; Jarnik, M.; Zaret, K.S. Opening of Compacted Chromatin by Early Developmental Transcription Factors HNF3 (FoxA) and GATA-4. Mol. Cell 2002, 9, 279–289. [Google Scholar] [CrossRef]
  53. Pomerantz, M.M.; Li, F.; Takeda, D.Y.; Lenci, R.; Chonkar, A.; Chabot, M.S.; Cejas, P.; Vazquez, F.; Cook, J.; Shivdasani, R.A.; et al. The androgen receptor cistrome is extensively reprogrammed in human prostate tumorigenesis. Nat. Genet. 2015, 47, 1346–1351. [Google Scholar] [CrossRef] [PubMed]
  54. Jin, H.; Zhao, J.C.; Wu, L.; Kim, J.; Yu, J. Cooperativity and equilibrium with FOXA1 define the androgen receptor transcriptional program. Nat. Commun. 2014, 5, 3972. [Google Scholar] [CrossRef]
  55. Casciello, F.; Al-Ejeh, F.; Kelly, G.; Brennan, D.J.; Ngiow, S.F.; Young, A.; Stoll, T.; Windloch, K.; Hill, M.M.; Smyth, M.J.; et al. G9a drives hypoxia-mediated gene repression for breast cancer cell survival and tumorigenesis. Proc. Natl. Acad. Sci. USA 2017, 114, 7077–7082. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, Q.; Li, W.; Liu, X.S.; Carroll, J.; Jänne, O.A.; Keeton, E.K.; Chinnaiyan, A.M.; Pienta, K.; Brown, M. A Hierarchical Network of Transcription Factors Governs Androgen Receptor-Dependent Prostate Cancer Growth. Mol. Cell 2007, 27, 380–392. [Google Scholar] [CrossRef] [Green Version]
  57. Krum, S.A.; Miranda-Carboni, G.A.; Lupien, M.; Eeckhoute, J.; Carroll, J.S.; Brown, M. Unique ERα Cistromes Control Cell Type-Specific Gene Regulation. Mol. Endocrinol. 2008, 22, 2393–2406. [Google Scholar] [CrossRef] [Green Version]
  58. Sanalkumar, R.; Johnson, K.D.; Gao, X.; Boyer, M.E.; Chang, Y.-I.; Hewitt, K.J.; Zhang, J.; Bresnick, E.H. Mechanism governing a stem cell-generating cis-regulatory element. Proc. Natl. Acad. Sci. USA 2014, 111, E1091–E1100. [Google Scholar] [CrossRef] [Green Version]
  59. Wu, D.; Sunkel, B.; Chen, Z.; Liu, X.; Ye, Z.; Li, Q.; Grenade, C.; Ke, J.; Zhang, C.; Chen, H.; et al. Three-tiered role of the pioneer factor GATA2 in promoting androgen-dependent gene expression in prostate cancer. Nucleic Acids Res. 2014, 42, 3607–3622. [Google Scholar] [CrossRef]
  60. Mann, R.S.; Affolter, M. Hox proteins meet more partners. Curr. Opin. Genet. Dev. 1998, 8, 423–429. [Google Scholar] [CrossRef]
  61. Yin, Y.; Morgunova, E.; Jolma, A.; Kaasinen, E.; Sahu, B.; Khund-Sayeed, S.; Das, P.K.; Kivioja, T.; Dave, K.; Zhong, F.; et al. Impact of cytosine methylation on DNA binding specificities of human transcription factors. Science 2017, 356, eaaj2239. [Google Scholar] [CrossRef]
  62. Norris, J.D.; Chang, C.-Y.; Wittmann, B.M.; Kunder, R.S.; Cui, H.; Fan, D.; Joseph, J.D.; McDonnell, D.P. The Homeodomain Protein HOXB13 Regulates the Cellular Response to Androgens. Mol. Cell 2009, 36, 405–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. He, B.; Lanz, R.B.; Fiskus, W.; Geng, C.; Yi, P.; Hartig, S.M.; Rajapakshe, K.; Shou, J.; Wei, L.; Shah, S.S.; et al. GATA2 facilitates steroid receptor coactivator recruitment to the androgen receptor complex. Proc. Natl. Acad. Sci. USA 2014, 111, 18261–18266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gao, S.; Chen, S.; Han, D.; Wang, Z.; Li, M.; Han, W.; Besschetnova, A.; Liu, M.; Zhou, F.; Barrett, D.; et al. Chromatin binding of FOXA1 is promoted by LSD1-mediated demethylation in prostate cancer. Nat. Genet. 2020, 52, 1011–1017. [Google Scholar] [CrossRef]
  65. Claessens, F.; Rushmere, N.; Davies, P.; Celis, L.; Peeters, B.; Rombauts, W. Sequence-specific binding of androgen-receptor complexes to prostatic binding protein genes. Mol. Cell. Endocrinol. 1990, 74, 203–212. [Google Scholar] [CrossRef]
  66. Rushmere, N.; Parker, M.; Davies, P. Androgen receptor-binding regions of an androgen-responsive gene. Mol. Cell. Endocrinol. 1987, 51, 259–265. [Google Scholar] [CrossRef]
  67. Rennie, P.S.; Bruchovsky, N.; Leco, K.J.; Sheppard, P.C.; McQueen, S.A.; Cheng, H.; Snoek, R.; Hamel, A.; Bock, M.E.; Macdonald, B.S.; et al. Characterization of two cis-acting DNA elements involved in the androgen regulation of the probasin gene. Mol. Endocrinol. 1993, 7, 23–36. [Google Scholar] [CrossRef] [Green Version]
  68. Luke, M.C.; Coffey, D.S. Human androgen receptor binding to the androgen response element of prostate specific antigen. J. Androl. 1994, 15, 41–51. [Google Scholar] [PubMed]
  69. Riegman, P.H.J.; Vlietstra, R.J.; Van Der Korput, J.A.G.M.; Brinkmann, A.O.; Trapman, J. The Promoter of the Prostate-Specific Antigen Gene Contains a Functional Androgen Responsive Element. Mol. Endocrinol. 1991, 5, 1921–1930. [Google Scholar] [CrossRef]
  70. Yamamoto, K.R.; Alberts, B. On the Specificity of the Binding of the Estradiol Receptor Protein to Deoxyribonucleic Acid. J. Biol. Chem. 1974, 249, 7076–7086. [Google Scholar] [CrossRef]
  71. Gronemeyer, H.; Pongs, O. Localization of ecdysterone on polytene chromosomes of Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 1980, 77, 2108–2112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Benoist, C.; Chambon, P. In vivo sequence requirements of the SV40 early promoter region. Nature 1981, 290, 304–310. [Google Scholar] [CrossRef] [PubMed]
  73. Cleutjens, K.B.; van der Korput, H.A.; van Eekelen, C.C.; van Rooij, H.C.; Faber, P.W.; Trapman, J. An androgen response element in a far upstream enhancer region is essential for high, androgen-regulated activity of the prostate-specific antigen promoter. Mol. Endocrinol. 1997, 11, 148–161. [Google Scholar] [CrossRef]
  74. Watt, F.; Martorana, A.; Brookes, D.E.; Ho, T.; Kingsley, E.; O’Keefe, D.S.; Russell, P.J.; Hestond, W.D.W.; Molloy, P.L. A tissue-specific enhancer of the prostate-specific membrane antigen gene, FOLH1. Genomics 2001, 73, 243–254. [Google Scholar] [CrossRef] [PubMed]
  75. Lu, S.; Liu, M.; Epner, D.E.; Tsai, S.Y.; Tsai, M.J. Androgen regulation of the cyclin-dependent kinase inhibitor p21 gene through an androgen response element in the proximal promoter. Mol. Endocrinol. 1999, 13, 376–384. [Google Scholar] [CrossRef]
  76. Yu, J.; Yu, J.; Mani, R.; Cao, Q.; Brenner, C.J.; Cao, X.; Wang, X.; Wu, L.; Li, J.; Hu, M.; et al. An Integrated Network of Androgen Receptor, Polycomb, and TMPRSS2-ERG Gene Fusions in Prostate Cancer Progression. Cancer Cell 2010, 17, 443–454. [Google Scholar] [CrossRef] [Green Version]
  77. Massie, C.L.E.; Adryan, B.; Barbosa-Morais, N.; Lynch, A.; Tran, M.G.B.; Neal, D.; Mills, I.G. New androgen receptor genomic targets show an interaction with the ETS1 transcription factor. EMBO Rep. 2007, 8, 871–878. [Google Scholar] [CrossRef] [Green Version]
  78. Wilson, S.; Qi, J.; Filipp, F.V. Refinement of the androgen response element based on ChIP-Seq in androgen-insensitive and androgen-responsive prostate cancer cell lines. Sci. Rep. 2016, 6, 32611. [Google Scholar] [CrossRef]
  79. Pomerantz, M.M.; Qiu, X.; Zhu, Y.; Takeda, D.Y.; Pan, W.; Baca, S.C.; Gusev, A.; Korthauer, K.; Severson, T.M.; Ha, G.; et al. Prostate cancer reactivates developmental epigenomic programs during metastatic progression. Nat. Genet. 2020, 52, 790–799. [Google Scholar] [CrossRef]
  80. Urbanucci, A.; Sahu, B.; Seppälä, J.; Larjo, A.; Latonen, L.M.; Waltering, K.K.; Tammela, T.L.J.; Vessella, R.L.; Lähdesmäki, H.; Jänne, O.A.; et al. Overexpression of androgen receptor enhances the binding of the receptor to the chromatin in prostate cancer. Oncogene 2012, 31, 2153–2163. [Google Scholar] [CrossRef] [Green Version]
  81. Urbanucci, A.; Barfeld, S.J.; Kytölä, V.; Itkonen, H.M.; Coleman, I.M.; Vodák, D.; Sjöblom, L.; Sheng, X.; Tolonen, T.; Minner, S.; et al. Androgen Receptor Deregulation Drives Bromodomain-Mediated Chromatin Alterations in Prostate Cancer. Cell Rep. 2017, 19, 2045–2059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Stelloo, S.; Nevedomskaya, E.; Poel, H.G.; Jong, J.; Leenders, G.J.; Jenster, G.; Wessels, L.F.A.; Bergman, A.M.; Zwart, W. Androgen receptor profiling predicts prostate cancer outcome. EMBO Mol. Med. 2015, 7, 1450–1464. [Google Scholar] [CrossRef] [PubMed]
  83. Roche, P.J.; Hoare, S.A.; Parker, M.G. A consensus DNA-binding site for the androgen receptor. Mol. Endocrinol. 1992, 6, 2229–2235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhang, L.; Martini, G.D.; Rube, H.T.; Kribelbauer, J.F.; Rastogi, C.; Fitzpatrick, V.D.; Houtman, J.; Bussemaker, H.J.; Pufall, M.A. SelexGLM differentiates androgen and glucocorticoid receptor DNA-binding preference over an extended binding site. Genome Res. 2018, 28, 111–121. [Google Scholar] [CrossRef] [Green Version]
  85. Deblois, G.; Giguère, V. Nuclear Receptor Location Analyses in Mammalian Genomes: From Gene Regulation to Regulatory Networks. Mol. Endocrinol. 2008, 22, 1999–2011. [Google Scholar] [CrossRef]
  86. Bolton, E.C.; So, A.Y.; Chaivorapol, C.; Haqq, C.M.; Li, H.; Yamamoto, K.R. Cell- and gene-specific regulation of primary target genes by the androgen receptor. Genes Dev. 2007, 21, 2005–2017. [Google Scholar] [CrossRef] [Green Version]
  87. Denayer, S.; Helsen, C.; Thorrez, L.; Haelens, A.; Claessens, F. The Rules of DNA Recognition by the Androgen Receptor. Mol. Endocrinol. 2010, 24, 898–913. [Google Scholar] [CrossRef] [Green Version]
  88. Chen, Z.; Lan, X.; Thomas-Ahner, J.; Wu, D.; Liu, X.; Ye, Z.; Wang, L.; Sunkel, B.; Grenade, C.; Chen, J.; et al. Agonist and antagonist switch DNA motifs recognized by human androgen receptor in prostate cancer. EMBO J. 2015, 34, 502–516. [Google Scholar] [CrossRef] [Green Version]
  89. McDowell, I.C.; Barrera, A.; D’Ippolito, A.M.; Vockley, C.M.; Hong, L.K.; Leichter, S.M.; Bartelt, L.C.; Majoros, W.H.; Song, L.; Safi, A.; et al. Glucocorticoid receptor recruits to enhancers and drives activation by motif-directed binding. Genome Res. 2018, 28, 1272–1284. [Google Scholar] [CrossRef] [Green Version]
  90. Huang, C.-C.F.; Lingadahalli, S.; Morova, T.; Ozturan, D.; Hu, E.; Yu, I.P.L.; Linder, S.; Hoogstraat, M.; Stelloo, S.; Sar, F.; et al. Functional mapping of androgen receptor enhancer activity. Genome Biol. 2021, 22, 149. [Google Scholar] [CrossRef]
  91. Cucchiara, V.; Yang, J.C.; Mirone, V.; Gao, A.C.; Rosenfeld, M.G.; Evans, C.P. Epigenomic Regulation of Androgen Receptor Signaling: Potential Role in Prostate Cancer Therapy. Cancers 2017, 9, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Varambally, S.; Dhanasekaran, S.M.; Zhou, M.; Barrette, T.R.; Kumar-Sinha, C.; Sanda, M.G.; Ghosh, D.; Pienta, K.J.; Sewalt, R.G.; Otte, A.P.; et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature 2002, 419, 624–629. [Google Scholar] [CrossRef] [PubMed]
  93. Yu, J.; Yu, J.; Rhodes, D.R.; Tomlins, S.; Cao, X.; Chen, G.; Mehra, R.; Wang, X.; Ghosh, D.; Shah, R.B.; et al. A Polycomb Repression Signature in Metastatic Prostate Cancer Predicts Cancer Outcome. Cancer Res. 2007, 67, 10657–10663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Davies, A.; Nouruzi, S.; Ganguli, D.; Namekawa, T.; Thaper, D.; Linder, S.; Karaoğlanoğlu, F.; Omur, M.E.; Kim, S.; Kobelev, M.; et al. An androgen receptor switch underlies lineage infidelity in treatment-resistant prostate cancer. Nat. Cell Biol. 2021, 23, 1023–1034. [Google Scholar] [CrossRef] [PubMed]
  95. Gaughan, L.; Stockley, J.; Wang, N.; McCracken, S.R.; Treumann, A.; Armstrong, K.; Shaheen, F.; Watt, K.; McEwan, I.J.; Wang, C.; et al. Regulation of the androgen receptor by SET9-mediated methylation. Nucleic Acids Res. 2011, 39, 1266–1279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Malik, R.; Khan, A.P.; Asangani, I.; Cieślik, M.; Prensner, J.; Wang, X.; Iyer, M.K.; Jiang, X.; Borkin, D.; Escara-Wilke, J.; et al. Targeting the MLL complex in castration-resistant prostate cancer. Nat. Med. 2015, 21, 344–352. [Google Scholar] [CrossRef] [PubMed]
  97. Anzick, S.L.; Kononen, J.; Walker, R.L.; Azorsa, D.O.; Tanner, M.M.; Guan, X.-Y.; Sauter, G.; Kallioniemi, O.-P.; Trent, J.M.; Meltzer, P.S. AIB1, a Steroid Receptor Coactivator Amplified in Breast and Ovarian Cancer. Science 1997, 277, 965–968. [Google Scholar] [CrossRef] [Green Version]
  98. Chen, H.; Lin, R.J.; Schiltz, R.; Chakravarti, D.; Nash, A.; Nagy, L.; Privalsky, M.L.; Nakatani, Y.; Evans, R.M. Nuclear Receptor Coactivator ACTR Is a Novel Histone Acetyltransferase and Forms a Multimeric Activation Complex with P/CAF and CBP/p300. Cell 1997, 90, 569–580. [Google Scholar] [CrossRef] [Green Version]
  99. Hong, H.; Kohli, K.; Trivedi, A.; Johnson, D.L.; Stallcup, M.R. GRIP1, a novel mouse protein that serves as a transcriptional coactivator in yeast for the hormone binding domains of steroid receptors. Proc. Natl. Acad. Sci. USA 1996, 93, 4948–4952. [Google Scholar] [CrossRef] [Green Version]
  100. Li, H.; Gomes, P.J.; Chen, J.D. RAC3, a steroid/nuclear receptor-associated coactivator that is related to SRC-1 and TIF2. Proc. Natl. Acad. Sci. USA 1997, 94, 8479–8484. [Google Scholar] [CrossRef] [Green Version]
  101. Oñate, S.A.; Tsai, S.Y.; Tsai, M.-J.; O’Malley, B.W. Sequence and Characterization of a Coactivator for the Steroid Hormone Receptor Superfamily. Science 1995, 270, 1354–1357. [Google Scholar] [CrossRef] [PubMed]
  102. Chakravarti, D.; Lamorte, V.J.; Nelson, M.C.; Nakajima, T.; Schulman, I.G.; Juguilon, H.; Montminy, M.; Evans, R. Role of CBP/P300 in nuclear receptor signalling. Nature 1996, 383, 99–103. [Google Scholar] [CrossRef] [PubMed]
  103. Hanstein, B.; Eckner, R.; DiRenzo, J.; Halachmi, S.; Liu, H.; Searcy, B.; Kurokawa, R.; Brown, M. p300 is a component of an estrogen receptor coactivator complex. Proc. Natl. Acad. Sci. USA 1996, 93, 11540–11545. [Google Scholar] [CrossRef] [Green Version]
  104. Blanco, J.C.; Minucci, S.; Lu, J.; Yang, X.-J.; Walker, K.K.; Chen, H.; Evans, R.M.; Nakatani, Y.; Ozato, K. The histone acetylase PCAF is a nuclear receptor coactivator. Genes Dev. 1998, 12, 1638–1651. [Google Scholar] [CrossRef] [Green Version]
  105. Abbas, A.; Gupta, S. The role of histone deacetylases in prostate cancer. Epigenetics 2008, 3, 300–309. [Google Scholar] [CrossRef] [PubMed]
  106. Rana, Z.; Diermeier, S.; Hanif, M.; Rosengren, R.J. Understanding Failure and Improving Treatment Using HDAC Inhibitors for Prostate Cancer. Biomedecines 2020, 8, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Huang, Z.-Q.; Li, J.; Sachs, L.M.; Cole, P.A.; Wong, J. A role for cofactor-cofactor and cofactor-histone interactions in targeting p300, SWI/SNF and Mediator for transcription. EMBO J. 2003, 22, 2146–2155. [Google Scholar] [CrossRef] [Green Version]
  108. Marshall, T.W.; Link, K.A.; Petre-Draviam, C.E.; Knudsen, K.E. Differential Requirement of SWI/SNF for Androgen Receptor Activity. J. Biol. Chem. 2003, 278, 30605–30613. [Google Scholar] [CrossRef] [Green Version]
  109. Jozwik, K.M.; Chernukhin, I.; Serandour, A.A.; Nagarajan, S.; Carroll, J.S. FOXA1 Directs H3K4 Monomethylation at Enhancers via Recruitment of the Methyltransferase MLL3. Cell Rep. 2016, 17, 2715–2723. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, Q.; Carroll, J.; Brown, M. Spatial and Temporal Recruitment of Androgen Receptor and Its Coactivators Involves Chromosomal Looping and Polymerase Tracking. Mol. Cell 2005, 19, 631–642. [Google Scholar] [CrossRef]
  111. Green, K.A.; Carroll, J. Oestrogen-receptor-mediated transcription and the influence of co-factors and chromatin state. Nat. Cancer 2007, 7, 713–722. [Google Scholar] [CrossRef] [PubMed]
  112. Cato, L.; Neeb, A.; Sharp, A.; Buzón, V.; Ficarro, S.B.; Yang, L.; Muhle-Goll, C.; Kuznik, N.C.; Riisnaes, R.; Rodrigues, D.N.; et al. Development of Bag-1L as a therapeutic target in androgen receptor-dependent prostate cancer. eLife 2017, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Coffey, K.; Rogerson, L.; Ryan-Munden, C.; Alkharaif, D.; Stockley, J.; Heer, R.; Sahadevan, K.; O’Neill, D.; Jones, D.; Darby, S.; et al. The lysine demethylase, KDM4B, is a key molecule in androgen receptor signalling and turnover. Nucleic Acids Res. 2013, 41, 4433–4446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. DePriest, A.D.; Fiandalo, M.; Schlanger, S.; Heemers, F.; Mohler, J.L.; Liu, S.; Heemers, H.V. Regulators of Androgen Action Resource: A one-stop shop for the comprehensive study of androgen receptor action. Database 2016, 2016. [Google Scholar] [CrossRef] [Green Version]
  115. Zhang, F.; Wong, S.; Lee, J.; Lingadahalli, S.; Wells, C.; Saxena, N.; Sanchez, C.; Sun, B.; Parra-Nuñez, A.K.; Chan, N.; et al. Dynamic phase separation of the androgen receptor and its coactivators to regulate gene expression. bioRxiv 2021. [Google Scholar] [CrossRef]
  116. Hnisz, D.; Abraham, B.; Lee, T.I.; Lau, A.; Saint-André, V.; Sigova, A.A.; Hoke, H.A.; Young, R.A. Super-Enhancers in the Control of Cell Identity and Disease. Cell 2013, 155, 934–947. [Google Scholar] [CrossRef] [Green Version]
  117. Wen, S.; He, Y.; Wang, L.; Zhang, J.; Quan, C.; Niu, Y.; Huang, H. Aberrant activation of super enhancer and choline metabolism drive antiandrogen therapy resistance in prostate cancer. Oncogene 2020, 39, 6556–6571. [Google Scholar] [CrossRef]
  118. Hsieh, C.-L.; Fei, T.; Chen, Y.; Li, T.; Gao, Y.; Wang, X.; Sun, T.; Sweeney, C.J.; Lee, G.-S.M.; Chen, S.; et al. Enhancer RNAs participate in androgen receptor-driven looping that selectively enhances gene activation. Proc. Natl. Acad. Sci. USA 2014, 111, 7319–7324. [Google Scholar] [CrossRef] [Green Version]
  119. Lee, N.; Steitz, J.A. Noncoding RNA-guided recruitment of transcription factors: A prevalent but undocumented mechanism? BioEssays 2015, 37, 936–941. [Google Scholar] [CrossRef]
  120. Ferreira, L.B.; Palumbo, A.; De Mello, K.D.; Sternberg, C.; Caetano, M.S.; De Oliveira, F.L.; Neves, A.F.; Nasciutti, L.E.; Goulart, L.R.; Gimba, E.R.P. PCA3 noncoding RNA is involved in the control of prostate-cancer cell survival and modulates androgen receptor signaling. BMC Cancer 2012, 12, 507. [Google Scholar] [CrossRef] [Green Version]
  121. Zhang, Y.; Pitchiaya, S.; Cieślik, M.; Niknafs, Y.S.; Tien, J.C.-Y.; Hosono, Y.; Iyer, M.K.; Yazdani, S.; Subramaniam, S.; Shukla, S.; et al. Analysis of the androgen receptor–regulated lncRNA landscape identifies a role for ARLNC1 in prostate cancer progression. Nat. Genet. 2018, 50, 814–824. [Google Scholar] [CrossRef]
  122. Prensner, J.R.; Iyer, M.K.; Sahu, A.; Asangani, I.A.; Cao, Q.; Patel, L.; Vergara, I.A.; Davicioni, E.; Erho, N.; Ghadessi, M.; et al. The long noncoding RNA SChLAP1 promotes aggressive prostate cancer and antagonizes the SWI/SNF complex. Nat. Genet. 2013, 45, 1392–1398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Simon, T.W.; Budinsky, R.A.; Rowlands, J.C. A Model for Aryl Hydrocarbon Receptor-Activated Gene Expression Shows Potency and Efficacy Changes and Predicts Squelching Due to Competition for Transcription Co-Activators. PLoS ONE 2015, 10, e0127952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Yang, Y.; Tse, A.K.-W.; Li, P.; Ma, Q.; Xiang, S.; Nicosia, S.V.; Seto, E.; Zhang, X.; Bai, W. Inhibition of androgen receptor activity by histone deacetylase 4 through receptor SUMOylation. Oncogene 2011, 30, 2207–2218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Carascossa, S.; Gobinet, J.; Georget, V.; Lucas, A.; Badia, E.; Castet, A.; White, R.; Nicolas, J.-C.; Cavailles, V.; Jalaguier, S.S. Receptor-Interacting Protein 140 Is a Repressor of the Androgen Receptor Activity. Mol. Endocrinol. 2006, 20, 1506–1518. [Google Scholar] [CrossRef]
  126. Fernandes, I.; Bastien, Y.; Wai, T.; Nygard, K.; Lin, R.; Cormier, O.; Lee, H.S.; Eng, F.; Bertos, N.R.; Pelletier, N.; et al. Ligand-Dependent Nuclear Receptor Corepressor LCoR Functions by Histone Deacetylase-Dependent and -Independent Mechanisms. Mol. Cell 2003, 11, 139–150. [Google Scholar] [CrossRef]
  127. Asim, M.; Bin Hafeez, B.; Siddiqui, I.A.; Gerlach, C.; Patz, M.; Mukhtar, H.; Baniahmad, A. Ligand-dependent Corepressor Acts as a Novel Androgen Receptor Corepressor, Inhibits Prostate Cancer Growth, and Is Functionally Inactivated by the Src Protein Kinase. J. Biol. Chem. 2011, 286, 37108–37117. [Google Scholar] [CrossRef] [Green Version]
  128. Nguyen, M.M.; Dincer, Z.; Wade, J.R.; Alur, M.; Michalak, M.; DeFranco, D.B.; Wang, Z. Cytoplasmic localization of the androgen receptor is independent of calreticulin. Mol. Cell. Endocrinol. 2009, 302, 65–72. [Google Scholar] [CrossRef] [Green Version]
  129. Dedhar, S.; Rennie, P.S.; Shago, M.; Hagesteijn, C.-Y.L.; Yang, H.; Filmus, J.; Hawley, R.G.; Bruchovsky, N.; Cheng, H.; Matusik, R.J.; et al. Inhibition of nuclear hormone receptor activity by calreticulin. Nature 1994, 367, 480–483. [Google Scholar] [CrossRef]
  130. Simons, S.S. Structure and Function of the Steroid and Nuclear Receptor Ligand Binding Domain. In Molecular Biology of Steroid and Nuclear Hormone Receptors; Freedman, L.P., Ed.; Birkhäuser Boston: Boston, MA, USA, 1998; pp. 35–104. [Google Scholar]
  131. Chen, C.; Bhalala, H.V.; Vessella, R.L.; Dong, J.-T. KLF5 is frequently deleted and down-regulated but rarely mutated in prostate cancer. Prostate 2003, 55, 81–88. [Google Scholar] [CrossRef]
  132. Ni, M.; Chen, Y.; Fei, T.; Li, D.; Lim, E.; Liu, X.S.; Brown, M. Amplitude modulation of androgen signaling by c-MYC. Genes Dev. 2013, 27, 734–748. [Google Scholar] [CrossRef] [Green Version]
  133. Barfeld, S.J.; Urbanucci, A.; Itkonen, H.; Fazli, L.; Hicks, J.L.; Thiede, B.; Rennie, P.S.; Yegnasubramanian, S.; DeMarzo, A.M.; Mills, I.G. c-Myc Antagonises the Transcriptional Activity of the Androgen Receptor in Prostate Cancer Affecting Key Gene Networks. EBioMedicine 2017, 18, 83–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Guo, H.; Wu, Y.; Nouri, M.; Spisak, S.; Russo, J.W.; Sowalsky, A.G.; Pomerantz, M.M.; Wei, Z.; Korthauer, K.; Seo, J.-H.; et al. Androgen receptor and MYC equilibration centralizes on developmental super-enhancer. Nat. Commun. 2021, 12, 7308. [Google Scholar] [CrossRef] [PubMed]
  135. Vatapalli, R.; Sagar, V.; Rodriguez, Y.; Zhao, J.C.; Unno, K.; Pamarthy, S.; Lysy, B.; Anker, J.; Han, H.; Yoo, Y.A.; et al. Histone methyltransferase DOT1L coordinates AR and MYC stability in prostate cancer. Nat. Commun. 2020, 11, 4153. [Google Scholar] [CrossRef] [PubMed]
  136. Alimirah, F.; Panchanathan, R.; Chen, J.; Zhang, X.; Ho, S.-M.; Choubey, D. Expression of Androgen Receptor Is Negatively Regulated By p53. Neoplasia 2007, 9, 1152–1159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Rokhlin, O.W.; Taghiyev, A.F.; Guseva, N.V.; Glover, R.A.; Chumakov, P.; Kravchenko, J.E.; Cohen, M.B. Androgen regulates apoptosis induced by TNFR family ligands via multiple signaling pathways in LNCaP. Oncogene 2005, 24, 6773–6784. [Google Scholar] [CrossRef] [Green Version]
  138. Wise, H.M.; Hermida, M.A.; Leslie, N.R. Prostate cancer, PI3K, PTEN and prognosis. Clin. Sci. 2017, 131, 197–210. [Google Scholar] [CrossRef]
  139. Wang, Y.; Romigh, T.; He, X.; Tan, M.-H.; Orloff, M.S.; Silverman, R.H.; Heston, W.D.; Eng, C. Differential regulation of PTEN expression by androgen receptor in prostate and breast cancers. Oncogene 2011, 30, 4327–4338. [Google Scholar] [CrossRef] [Green Version]
  140. Li, Q.; Liu, B.; Chao, H.-P.; Ji, Y.; Lu, Y.; Mehmood, R.; Jeter, C.; Chen, T.; Moore, J.R.; Li, W.; et al. LRIG1 is a pleiotropic androgen receptor-regulated feedback tumor suppressor in prostate cancer. Nat. Commun. 2019, 10, 5495. [Google Scholar] [CrossRef]
  141. Thomasson, M.; Wang, B.; Hammarsten, P.; Dahlman, A.; Persson, J.L.; Josefsson, A.; Stattin, P.; Granfors, T.; Egevad, L.; Henriksson, R.; et al. LRIG1 and the liar paradox in prostate cancer: A study of the expression and clinical significance of LRIG1 in prostate cancer. Int. J. Cancer 2011, 128, 2843–2852. [Google Scholar] [CrossRef]
  142. Gareau, J.R.; Lima, C.D. The SUMO pathway: Emerging mechanisms that shape specificity, conjugation and recognition. Nat. Rev. Mol. Cell Biol. 2010, 11, 861–871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Jones, P.L.; Shi, Y.B. N-CoR-HDAC corepressor complexes: Roles in transcriptional regulation by nuclear hormone receptors. Curr. Top Microbiol. Immunol. 2003, 274, 237–268. [Google Scholar] [PubMed]
  144. Shashikant, T.; Ettensohn, C.A. Genome-wide analysis of chromatin accessibility using ATAC-seq. Methods Cell Biol. 2019, 151, 219–235. [Google Scholar] [CrossRef]
  145. Wang, Q.; Li, W.; Zhang, Y.; Yuan, X.; Xu, K.; Yu, J.; Chen, Z.; Beroukhim, R.; Wang, H.; Lupien, M.; et al. Androgen Receptor Regulates a Distinct Transcription Program in Androgen-Independent Prostate Cancer. Cell 2009, 138, 245–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Barozzi, I.; Simonatto, M.; Bonifacio, S.; Yang, L.; Rohs, R.; Ghisletti, S.; Natoli, G. Coregulation of Transcription Factor Binding and Nucleosome Occupancy through DNA Features of Mammalian Enhancers. Mol. Cell 2014, 54, 844–857. [Google Scholar] [CrossRef] [Green Version]
  147. Hallikas, O.; Palin, K.; Sinjushina, N.; Rautiainen, R.; Partanen, J.; Ukkonen, E.; Taipale, J. Genome-wide Prediction of Mammalian Enhancers Based on Analysis of Transcription-Factor Binding Affinity. Cell 2006, 124, 47–59. [Google Scholar] [CrossRef] [Green Version]
  148. Hah, N.; Murakami, S.; Nagari, A.; Danko, C.G.; Kraus, W.L. Enhancer transcripts mark active estrogen receptor binding sites. Genome Res. 2013, 23, 1210–1223. [Google Scholar] [CrossRef] [Green Version]
  149. Koch, F.; Jourquin, F.; Ferrier, P.; Andrau, J.-C. Genome-wide RNA polymerase II: Not genes only! Trends Biochem. Sci. 2008, 33, 265–273. [Google Scholar] [CrossRef]
  150. Cheng, J.-H.; Pan, D.Z.-C.; Tsai, Z.T.-Y.; Tsai, H.-K. Genome-wide analysis of enhancer RNA in gene regulation across 12 mouse tissues. Sci. Rep. 2015, 5, 12648. [Google Scholar] [CrossRef] [Green Version]
  151. Arner, E.; Daub, C.O.; Vitting-Seerup, K.; Andersson, R.; Lilje, B.; Drabløs, F.; Lennartsson, A.; Rönnerblad, M.; Hrydziuszko, O.; Vitezic, M.; et al. Transcribed enhancers lead waves of coordinated transcription in transitioning mammalian cells. Science 2015, 347, 1010–1014. [Google Scholar] [CrossRef] [Green Version]
  152. Kouno, T.; Moody, J.; Kwon, A.T.-J.; Shibayama, Y.; Kato, S.; Huang, Y.; Böttcher, M.; Motakis, E.; Mendez, M.; Severin, J.; et al. C1 CAGE detects transcription start sites and enhancer activity at single-cell resolution. Nat. Commun. 2019, 10, 360. [Google Scholar] [CrossRef] [Green Version]
  153. Xie, S.; Duan, J.; Li, B.; Zhou, P.; Hon, G.C. Multiplexed Engineering and Analysis of Combinatorial Enhancer Activity in Single Cells. Mol. Cell 2017, 66, 285–299.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Zacher, B.; Michel, M.; Schwalb, B.; Cramer, P.; Tresch, A.; Gagneur, J. Accurate Promoter and Enhancer Identification in 127 ENCODE and Roadmap Epigenomics Cell Types and Tissues by GenoSTAN. PLoS ONE 2017, 12, e0169249. [Google Scholar] [CrossRef]
  155. Zhang, T.; Zhang, Z.; Dong, Q.; Xiong, J.; Zhu, B. Histone H3K27 acetylation is dispensable for enhancer activity in mouse embryonic stem cells. Genome Biol. 2020, 21, 45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Inoue, F.; Kircher, M.; Martin, B.; Cooper, G.M.; Witten, D.M.; McManus, M.T.; Ahituv, N.; Shendure, J. A systematic comparison reveals substantial differences in chromosomal versus episomal encoding of enhancer activity. Genome Res. 2017, 27, 38–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Arnold, C.D.; Gerlach, D.; Stelzer, C.; Boryń, L.M.; Rath, M.; Stark, A. Genome-Wide Quantitative Enhancer Activity Maps Identified by STARR-seq. Science 2013, 339, 1074–1077. [Google Scholar] [CrossRef] [PubMed]
  158. Neumayr, C.; Pagani, M.; Stark, A.; Arnold, C.D. STARR-seq and UMI-STARR-seq: Assessing Enhancer Activities for Genome-Wide-, High-, and Low-Complexity Candidate Libraries. Curr. Protoc. Mol. Biol. 2019, 128, e105. [Google Scholar] [CrossRef] [Green Version]
  159. Liu, Y.; Yuwen, L.; Dhiman, V.K.; Brunetti, T.; Eckart, H.; White, K.P. Functional assessment of human enhancer activities using whole-genome STARR-sequencing. Genome Biol. 2017, 18, 219. [Google Scholar] [CrossRef] [Green Version]
  160. Muerdter, F.; Boryń, Ł.M.; Arnold, C.D. STARR-seqPrinciples and applications. Genomics 2015, 106, 145–150. [Google Scholar] [CrossRef] [Green Version]
  161. Chen, Z.; Song, X.; Li, Q.; Xie, L.; Guo, T.; Su, T.; Tang, C.; Chang, X.; Liang, B.; Huang, D.; et al. Androgen Receptor-Activated Enhancers Simultaneously Regulate Oncogene TMPRSS2 and lncRNA PRCAT38 in Prostate Cancer. Cells 2019, 8, 864. [Google Scholar] [CrossRef] [Green Version]
  162. Carleton, J.B.; Berrett, K.C.; Gertz, J. Multiplex Enhancer Interference Reveals Collaborative Control of Gene Regulation by Estrogen Receptor α-Bound Enhancers. Cell Syst. 2017, 5, 333–344.e5. [Google Scholar] [CrossRef] [Green Version]
  163. Proudhon, C.; Snetkova, V.; Raviram, R.; Lobry, C.; Badri, S.; Jiang, T.; Hao, B.; Trimarchi, T.; Kluger, Y.; Aifantis, I.; et al. Active and Inactive Enhancers Cooperate to Exert Localized and Long-Range Control of Gene Regulation. Cell Rep. 2016, 15, 2159–2169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Carleton, J.B.; Ginley-Hidinger, M.; Berrett, K.C.; Layer, R.M.; Quinlan, A.R.; Gertz, J. Regulatory sharing between estrogen receptor α bound enhancers. Nucleic Acids Res. 2020, 48, 6597–6610. [Google Scholar] [CrossRef] [PubMed]
  165. Souaid, C.; Bloyer, S.; Noordermeer, D. 19—Promoter–Enhancer Looping and Regulatory Neighborhoods: Gene Regulation in the Framework of Topologically Associating Domains. In Nuclear Architecture and Dynamics; Lavelle, C., Victor, J.-M., Eds.; Academic Press: Boston, MA, USA, 2018; pp. 435–456. [Google Scholar]
  166. Rowley, M.J.; Corces, V.G. Organizational principles of 3D genome architecture. Nat. Rev. Genet. 2018, 19, 789–800. [Google Scholar] [CrossRef] [PubMed]
  167. Dixon, J.R.; Selvaraj, S.; Yue, F.; Kim, A.; Li, Y.; Shen, Y.; Hu, M.; Liu, J.S.; Ren, B. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 2012, 485, 376–380. [Google Scholar] [CrossRef] [Green Version]
  168. Sexton, T.; Yaffe, E.; Kenigsberg, E.; Bantignies, F.; Leblanc, B.; Hoichman, M.; Parrinello, H.; Tanay, A.; Cavalli, G. Three-Dimensional Folding and Functional Organization Principles of the Drosophila Genome. Cell 2012, 148, 458–472. [Google Scholar] [CrossRef] [Green Version]
  169. Zhang, Z.; Chng, K.R.; Lingadahalli, S.; Chen, Z.; Liu, M.H.; Do, H.H.; Cai, S.; Rinaldi, N.; Poh, H.M.; Li, G.; et al. An AR-ERG transcriptional signature defined by long-range chromatin interactomes in prostate cancer cells. Genome Res. 2019, 29, 223–235. [Google Scholar] [CrossRef] [Green Version]
  170. Rhie, S.K.; Perez, A.; Lay, F.D.; Schreiner, S.; Shi, J.; Polin, J.; Farnham, P.J. A high-resolution 3D epigenomic map reveals insights into the creation of the prostate cancer transcriptome. Nat. Commun. 2019, 10, 4154. [Google Scholar] [CrossRef] [Green Version]
  171. Palstra, R.-J.; Tolhuis, B.; Splinter, E.; Nijmeijer, R.; Grosveld, F.; de Laat, W. The β-globin nuclear compartment in development and erythroid differentiation. Nat. Genet. 2003, 35, 190–194. [Google Scholar] [CrossRef]
  172. Lettice, L.A.; Heaney, S.J.; Purdie, L.A.; Li, L.; De Beer, P.; Oostra, B.A.; Goode, D.; Elgar, G.; Hill, R.E.; De Graaff, E. A long-range Shh enhancer regulates expression in the developing limb and fin and is associated with preaxial polydactyly. Hum. Mol. Genet. 2003, 12, 1725–1735. [Google Scholar] [CrossRef]
  173. Andrey, G.; Montavon, T.; Mascrez, B.; Gonzalez, F.; Noordermeer, D.; Leleu, M.; Trono, D.; Spitz, F.; Duboule, D. A Switch Between Topological Domains Underlies HoxD Genes Collinearity in Mouse Limbs. Science 2013, 340, 1234167. [Google Scholar] [CrossRef] [PubMed]
  174. Vernimmen, D.; De Gobbi, M.; Sloane-Stanley, J.A.; Wood, W.G.; Higgs, D.R. Long-range chromosomal interactions regulate the timing of the transition between poised and active gene expression. EMBO J. 2007, 26, 2041–2051. [Google Scholar] [CrossRef] [PubMed]
  175. Verheul, T.C.J.; Van Hijfte, L.; Perenthaler, E.; Barakat, T.S. The Why of YY1: Mechanisms of Transcriptional Regulation by Yin Yang 1. Front. Cell Dev. Biol. 2020, 8, 592164. [Google Scholar] [CrossRef] [PubMed]
  176. Van de Werken, H.J.G.; Landan, G.; Holwerda, S.J.B.; Hoichman, M.; Klous, P.; Chachik, R.; Splinter, E.; Valdes-Quezada, C.; Öz, Y.; Bouwman, B.A.M.; et al. Robust 4C-seq data analysis to screen for regulatory DNA interactions. Nat. Methods 2012, 9, 969–972. [Google Scholar] [CrossRef]
  177. Sanyal, A.; Lajoie, B.R.; Jain, G.; Dekker, J. The long-range interaction landscape of gene promoters. Nature 2012, 489, 109–113. [Google Scholar] [CrossRef]
  178. Misteli, T. Higher-order Genome Organization in Human Disease. Cold Spring Harb. Perspect. Biol. 2010, 2, a000794. [Google Scholar] [CrossRef] [Green Version]
  179. Martin, R.S.; Das, P.; Marques, R.D.R.; Xu, Y.; McCord, R.P. Alterations in chromosome spatial compartmentalization classify prostate cancer progression. bioRxiv 2021. [Google Scholar] [CrossRef]
  180. Taberlay, P.; Achinger-Kawecka, J.; Lun, A.; Buske, F.A.; Sabir, K.; Gould, C.M.; Zotenko, E.; Bert, S.A.; Giles, K.; Bauer, D.; et al. Three-dimensional disorganization of the cancer genome occurs coincident with long-range genetic and epigenetic alterations. Genome Res. 2016, 26, 719–731. [Google Scholar] [CrossRef] [Green Version]
  181. Hawley, J.R.; Zhou, S.; Arlidge, C.; Grillo, G.; Kron, K.J.; Hugh-White, R.; van der Kwast, T.H.; Fraser, M.; Boutros, P.C.; Bristow, R.G.; et al. Reorganization of the 3D Genome Pinpoints Noncoding Drivers of Primary Prostate Tumors. Cancer Res. 2021, 81, 5833–5848. [Google Scholar] [CrossRef]
  182. Metzger, E.; Willmann, D.; McMillan, J.; Forne, I.; Metzger, P.; Gerhardt, S.; Petroll, K.; Von Maessenhausen, A.; Urban, S.; Schott, A.-K.; et al. Assembly of methylated KDM1A and CHD1 drives androgen receptor–dependent transcription and translocation. Nat. Struct. Mol. Biol. 2016, 23, 132–139. [Google Scholar] [CrossRef]
  183. Malinen, M.; Niskanen, E.; Kaikkonen, M.; Palvimo, J.J. Crosstalk between androgen and pro-inflammatory signaling remodels androgen receptor and NF-κB cistrome to reprogram the prostate cancer cell transcriptome. Nucleic Acids Res. 2017, 45, 619–630. [Google Scholar] [CrossRef] [Green Version]
  184. Martincorena, I.; Campbell, P.J. Somatic mutation in cancer and normal cells. Science 2015, 349, 1483–1489. [Google Scholar] [CrossRef] [PubMed]
  185. Pan-cancer analysis of whole genomes. Nature 2020, 578, 82–93. [CrossRef] [Green Version]
  186. Frank, S.; Nelson, P.; Vasioukhin, V. Recent advances in prostate cancer research: Large-scale genomic analyses reveal novel driver mutations and DNA repair defects. F1000Research 2018, 7, 1173. [Google Scholar] [CrossRef] [PubMed]
  187. Barbieri, C.; Bangma, C.H.; Bjartell, A.; Catto, J.; Culig, Z.; Grönberg, H.; Luo, J.; Visakorpi, T.; Rubin, M. The Mutational Landscape of Prostate Cancer. Eur. Urol. 2013, 64, 567–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Lang, S.H.; Swift, S.L.; White, H.; Misso, K.; Kleijnen, J.; Quek, R.G. A systematic review of the prevalence of DNA damage response gene mutations in prostate cancer. Int. J. Oncol. 2019, 55, 597–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Elliott, K.; Larsson, E. Non-coding driver mutations in human cancer. Nat. Cancer 2021, 21, 500–509. [Google Scholar] [CrossRef] [PubMed]
  190. Walavalkar, K.; Notani, D. Beyond the coding genome: Non-coding mutations and cancer. Front. Biosci. 2020, 25, 1828–1838. [Google Scholar]
  191. ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef]
  192. Li, Y.; PCAWG Structural Variation Working Group; Roberts, N.; Wala, J.A.; Shapira, O.; Schumacher, S.E.; Kumar, K.; Khurana, E.; Waszak, S.; Korbel, J.O.; et al. Patterns of somatic structural variation in human cancer genomes. Nature 2020, 578, 112–121. [Google Scholar] [CrossRef] [Green Version]
  193. Takeda, D.Y.; Spisák, S.; Seo, J.-H.; Bell, C.; O’Connor, E.; Korthauer, K.; Ribli, D.; Csabai, I.; Solymosi, N.; Szallasi, Z.; et al. A Somatically Acquired Enhancer of the Androgen Receptor Is a Noncoding Driver in Advanced Prostate Cancer. Cell 2018, 174, 422–432.e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Viswanathan, S.; Ha, G.; Hoff, A.M.; Wala, J.A.; Carrot-Zhang, J.; Whelan, C.; Haradhvala, N.J.; Freeman, S.; Reed, S.; Rhoades, J.; et al. Structural Alterations Driving Castration-Resistant Prostate Cancer Revealed by Linked-Read Genome Sequencing. Cell 2018, 174, 433–447.e19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Weischenfeldt, J.; Dubash, T.; Drainas, A.P.; Mardin, B.R.; Chen, Y.; Stütz, A.M.; Waszak, S.M.; Bosco, G.; Halvorsen, A.R.; Raeder, B.; et al. Pan-cancer analysis of somatic copy-number alterations implicates IRS4 and IGF2 in enhancer hijacking. Nat. Genet. 2017, 49, 65–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Spielmann, M.; Lupiáñez, D.G.; Mundlos, S. Structural variation in the 3D genome. Nat. Rev. Genet. 2018, 19, 453–467. [Google Scholar] [CrossRef] [Green Version]
  197. Wang, X.; Xu, J.; Zhang, B.; Hou, Y.; Song, F.; Lyu, H.; Yue, F. Genome-wide detection of enhancer-hijacking events from chromatin interaction data in rearranged genomes. Nat. Methods 2021, 18, 661–668. [Google Scholar] [CrossRef]
  198. Tomlins, S.A.; Rhodes, D.R.; Perner, S.; Dhanasekaran, S.M.; Mehra, R.; Sun, X.-W.; Varambally, S.; Cao, X.; Tchinda, J.; Kuefer, R.; et al. Recurrent Fusion of TMPRSS2 and ETS Transcription Factor Genes in Prostate Cancer. Science 2005, 310, 644–648. [Google Scholar] [CrossRef]
  199. St John, J.; Powell, K.; Conley-Lacomb, M.K.; Chinni, S.R. TMPRSS2-ERG Fusion Gene Expression in Prostate Tumor Cells and Its Clinical and Biological Significance in Prostate Cancer Progression. J. Cancer Sci. Ther. 2012, 4, 94–101. [Google Scholar] [CrossRef] [Green Version]
  200. Wang, Z.; Wang, Y.; Zhang, J.; Zhu, W.; Zhi, F.; Zhang, S.; Mao, D.; Zhang, Y.; Yuliang, W. Significance of the TMPRSS2:ERG gene fusion in prostate cancer. Mol. Med. Rep. 2017, 16, 5450–5458. [Google Scholar] [CrossRef] [Green Version]
  201. Baca, S.C.; Prandi, D.; Lawrence, M.S.; Mosquera, J.M.; Romanel, A.; Drier, Y.; Park, K.; Kitabayashi, N.; Macdonald, T.Y.; Ghandi, M.; et al. Punctuated Evolution of Prostate Cancer Genomes. Cell 2013, 153, 666–677. [Google Scholar] [CrossRef] [Green Version]
  202. Schuijers, J.; Manteiga, J.C.; Weintraub, A.S.; Day, D.S.; Zamudio, A.V.; Hnisz, D.; Lee, T.I.; Young, R.A. Transcriptional Dysregulation of MYC Reveals Common Enhancer-Docking Mechanism. Cell Rep. 2018, 23, 349–360. [Google Scholar] [CrossRef] [Green Version]
  203. Helmsauer, K.; Valieva, M.E.; Ali, S.; Chamorro González, R.; Schöpflin, R.; Röefzaad, C.; Bei, Y.; Dorado Garcia, H.; Rodriguez-Fos, E.; Puiggròs, M.; et al. Enhancer hijacking determines extrachromosomal circular MYCN amplicon architecture in neuroblastoma. Nat. Commun. 2020, 11, 5823. [Google Scholar] [CrossRef] [PubMed]
  204. Khoury, A.; Achinger-Kawecka, J.; Bert, S.A.; Smith, G.C.; French, H.J.; Luu, P.-L.; Peters, T.J.; Du, Q.; Parry, A.J.; Valdes-Mora, F.; et al. Constitutively bound CTCF sites maintain 3D chromatin architecture and long-range epigenetically regulated domains. Nat. Commun. 2020, 11, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Hsieh, T.H.S.; Cattoglio, C.; Slobodyanyuk, E.; Hansen, A.S. Enhancer-Promoter Interactions and Transcription Are Maintained upon Acute Loss of CTCF, Cohesin, WAPL, and YY1. bioRxiv 2021. Available online: https://www.biorxiv.org/content/10.1101/2021.07.14.452365.abstract (accessed on 16 January 2022).
  206. Yan, Y.; Guo, G.; Huang, J.; Gao, M.; Zhu, Q.; Zeng, S.; Gong, Z.; Xu, Z. Current understanding of extrachromosomal circular DNA in cancer pathogenesis and therapeutic resistance. J. Hematol. Oncol. 2020, 13, 124. [Google Scholar] [CrossRef] [PubMed]
  207. Kim, H.; Nguyen, N.-P.; Turner, K.; Wu, S.; Gujar, A.D.; Luebeck, J.; Liu, J.; Deshpande, V.; Rajkumar, U.; Namburi, S.; et al. Extrachromosomal DNA is associated with oncogene amplification and poor outcome across multiple cancers. Nat. Genet. 2020, 52, 891–897. [Google Scholar] [CrossRef] [PubMed]
  208. Hawley, J.R.; Zhou, S.; Arlidge, C.; Grillo, G.; Kron, K.J.; Hugh-White, R.; van der Kwast, T.H.; Fraser, M.; Boutros, P.C.; Bristow, R.G.; et al. Cis-regulatory Element Hijacking by Structural Variants Overshadows Higher-Order Topological Changes in Prostate Cancer. bioRxiv 2021. [Google Scholar] [CrossRef]
  209. Zhu, Y.; Gujar, A.D.; Wong, C.-H.; Tjong, H.; Ngan, C.Y.; Gong, L.; Chen, Y.-A.; Kim, H.; Liu, J.; Li, M.; et al. Oncogenic extrachromosomal DNA functions as mobile enhancers to globally amplify chromosomal transcription. Cancer Cell 2021, 39, 694–707.e7. [Google Scholar] [CrossRef] [PubMed]
  210. Oh, S.; Shao, J.; Mitra, J.; Xiong, F.; D’Antonio, M.; Wang, R.; Garcia-Bassets, I.; Ma, Q.; Zhu, X.; Lee, J.-H.; et al. Enhancer release and retargeting activates disease-susceptibility genes. Nature 2021, 595, 735–740. [Google Scholar] [CrossRef]
  211. Mazrooei, P.; Kron, K.J.; Zhu, Y.; Zhou, S.; Grillo, G.; Mehdi, T.; Ahmed, M.; Severson, T.M.; Guilhamon, P.; Armstrong, N.S.; et al. Cistrome Partitioning Reveals Convergence of Somatic Mutations and Risk Variants on Master Transcription Regulators in Primary Prostate Tumors. Cancer Cell 2019, 36, 674–689.e6. [Google Scholar] [CrossRef]
  212. Bailey, S.D.; Desai, K.; Kron, K.J.; Mazrooei, P.; Sinnott-Armstrong, N.A.; Treloar, A.E.; Dowar, M.; Thu, K.L.; Cescon, D.W.; Silvester, J.; et al. Noncoding somatic and inherited single-nucleotide variants converge to promote ESR1 expression in breast cancer. Nat. Genet. 2016, 48, 1260–1266. [Google Scholar] [CrossRef]
  213. Hua, J.T.; Ahmed, M.; Guo, H.; Zhang, Y.; Chen, S.; Soares, F.; Lu, J.; Zhou, S.; Wang, M.; Li, H.; et al. Risk SNP-Mediated Promoter-Enhancer Switching Drives Prostate Cancer through lncRNA PCAT19. Cell 2018, 174, 564–575.e18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Takayama, K.-I.; Suzuki, T.; Fujimura, T.; Urano, T.; Takahashi, S.; Homma, Y.; Inoue, S. CtBP2 Modulates the Androgen Receptor to Promote Prostate Cancer Progression. Cancer Res. 2014, 74, 6542–6553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Gao, P.; Xia, J.; Sipeky, C.; Dong, X.-M.; Zhang, Q.; Yang, Y.; Zhang, P.; Cruz, S.P.; Zhang, K.; Zhu, J.; et al. Biology and Clinical Implications of the 19q13 Aggressive Prostate Cancer Susceptibility Locus. Cell 2018, 174, 576–589.e18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Morova, T.; McNeill, D.R.; Lallous, N.; Gönen, M.; Dalal, K.; Wilson, D.M., 3rd; Gürsoy, A.; Keskin, Ö.; Lack, N.A. Androgen receptor-binding sites are highly mutated in prostate cancer. Nat Commun. 2020, 11, 832. [Google Scholar] [CrossRef] [Green Version]
  217. 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]
  218. Xu, D.; Chen, Q.; Liu, Y.; Wen, X. Baicalein suppresses the androgen receptor (AR)-mediated prostate cancer progression via inhibiting the AR N-C dimerization and AR-coactivators interaction. Oncotarget 2017, 8, 105561–105573. [Google Scholar] [CrossRef] [Green Version]
  219. ENCODE Project Consortium. The ENCODE (ENCyclopedia Of DNA Elements) Project. Science 2004, 306, 636–640. [Google Scholar] [CrossRef] [Green Version]
  220. Dekker, J.; Belmont, A.S.; Guttman, M.; Leshyk, V.O.; Lis, J.T.; Lomvardas, S.; Mirny, L.A.; O’Shea, C.C.; Park, P.J.; Ren, B.; et al. The 4D nucleome project. Nature 2017, 549, 219–226. [Google Scholar] [CrossRef]
Figure 1. Essential and AR-upregulated genes in androgen-sensitive and PSA-positive LNCaP cell line. Data taken from 5 LNCaP genome-wide CRISPR screens in DepMap database (DepMap 21Q4, DepMap 21Q3, GeCKO 19Q1, GeCKO CERES, Sanger CERES) and ranked based on their essentiality. AR-upregulated genes are taken from RNA-seq samples of androgen-treated LNCaP cells (GEO: GSE64529).
Figure 1. Essential and AR-upregulated genes in androgen-sensitive and PSA-positive LNCaP cell line. Data taken from 5 LNCaP genome-wide CRISPR screens in DepMap database (DepMap 21Q4, DepMap 21Q3, GeCKO 19Q1, GeCKO CERES, Sanger CERES) and ranked based on their essentiality. AR-upregulated genes are taken from RNA-seq samples of androgen-treated LNCaP cells (GEO: GSE64529).
Cells 11 00898 g001
Figure 2. Cartoon representation of ARBS enhancer activity on AR-mediated gene. Upon AR binding, coactivators, mediator complex, cohesin proteins, and transcriptional machinery are recruited to initiate gene expression.
Figure 2. Cartoon representation of ARBS enhancer activity on AR-mediated gene. Upon AR binding, coactivators, mediator complex, cohesin proteins, and transcriptional machinery are recruited to initiate gene expression.
Cells 11 00898 g002
Figure 3. AR-mediated gene repression. AR is bound by corepressors such as NCoR/SMRT, creating a corepressor complex and facilitating HDAC activity to suppress gene activation.
Figure 3. AR-mediated gene repression. AR is bound by corepressors such as NCoR/SMRT, creating a corepressor complex and facilitating HDAC activity to suppress gene activation.
Cells 11 00898 g003
Figure 4. ARBS interaction landscape. (a) Multiple ARBS reside in close proximity to AR-regulated genes. Frequency of ARBS were quantified using publicly available androgen-induced RNA-seq [182] and AR ChIP-seq [183] from LNCaP cells. (b) AR ChIA-PET cis-contacts are mostly concentrated around 10 kb to 1 mb. AR-mediated looping dataset were used calculate the frequency of chromatin loops in VCaP cells [169].
Figure 4. ARBS interaction landscape. (a) Multiple ARBS reside in close proximity to AR-regulated genes. Frequency of ARBS were quantified using publicly available androgen-induced RNA-seq [182] and AR ChIP-seq [183] from LNCaP cells. (b) AR ChIA-PET cis-contacts are mostly concentrated around 10 kb to 1 mb. AR-mediated looping dataset were used calculate the frequency of chromatin loops in VCaP cells [169].
Cells 11 00898 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Özturan, D.; Morova, T.; Lack, N.A. Androgen Receptor-Mediated Transcription in Prostate Cancer. Cells 2022, 11, 898. https://doi.org/10.3390/cells11050898

AMA Style

Özturan D, Morova T, Lack NA. Androgen Receptor-Mediated Transcription in Prostate Cancer. Cells. 2022; 11(5):898. https://doi.org/10.3390/cells11050898

Chicago/Turabian Style

Özturan, Doğancan, Tunç Morova, and Nathan A. Lack. 2022. "Androgen Receptor-Mediated Transcription in Prostate Cancer" Cells 11, no. 5: 898. https://doi.org/10.3390/cells11050898

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