Next Article in Journal
Comparing Outcomes of Oligometastases Treated with Hypofractionated Image-Guided Radiotherapy (HIGRT) with a Simultaneous Integrated Boost (SIB) Technique versus Metastasis Alone: A Multi-Institutional Analysis
Previous Article in Journal
Current Trend of Artificial Intelligence Patents in Digital Pathology: A Systematic Evaluation of the Patent Landscape
Previous Article in Special Issue
FOXA1: A Pioneer of Nuclear Receptor Action in Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nuclear Receptor Coregulators in Hormone-Dependent Cancers

by
Hedieh Jafari
1,2,†,
Shahid Hussain
2,† and
Moray J. Campbell
2,*
1
Department of Molecular Genetics, The Ohio State University, Columbus, OH 43210, USA
2
Department of Pharmaceutics and Pharmaceutical Chemistry, College of Pharmacy, The Ohio State University, Columbus, OH 43210, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2022, 14(10), 2402; https://doi.org/10.3390/cancers14102402
Submission received: 29 April 2022 / Accepted: 9 May 2022 / Published: 13 May 2022
(This article belongs to the Special Issue Nuclear Receptors and Cancer)

Abstract

:

Simple Summary

Altered nuclear receptor signaling is well-established to contribute to a range of hormone-dependent cancers and therefore they are therapeutic targets, for example in breast and prostate cancer. However, nuclear receptor signaling in the cancer, and in response to therapeutic targeting is frequently altered in part by changes to the expression and function of the large numbers of coregulators that modulate sensitivity of the receptor, change their protein-protein interaction and alter the transcriptional consequences. These coregulators perform numerous roles in terms of regulating nuclear receptor responses and the functional diversity of these interactions is an area of emerging insight. Coregulator interactions, functions, and impacts on nuclear receptor signaling across hormone dependent cancers are the topic of this review.

Abstract

Nuclear receptors (NRs) function collectively as a transcriptional signaling network that mediates gene regulatory actions to either maintain cellular homeostasis in response to hormonal, dietary and other environmental factors, or act as orphan receptors with no known ligand. NR complexes are large and interact with multiple protein partners, collectively termed coregulators. Coregulators are essential for regulating NR activity and can dictate whether a target gene is activated or repressed by a variety of mechanisms including the regulation of chromatin accessibility. Altered expression of coregulators contributes to a variety of hormone-dependent cancers including breast and prostate cancers. Therefore, understanding the mechanisms by which coregulators interact with and modulate the activity of NRs provides opportunities to develop better prognostic and diagnostic approaches, as well as novel therapeutic targets. This review aims to gather and summarize recent studies, techniques and bioinformatics methods used to identify distorted NR coregulator interactions that contribute as cancer drivers in hormone-dependent cancers.

1. Introduction

1.1. Nuclear Receptors and Hormone-Dependent Cancers

The healthy functioning of multicellular organisms requires highly coordinated gene regulation (both protein coding and non-coding RNA transcripts) to govern cell-fate decisions, sustain tissue organization, maintain homeostasis and respond to external stimuli, including towards dietary and xenobiotic compounds. These gene networks are directed by the combinatorial actions of cis-binding transcription factor (TF) complexes interacting with and impacting the underlying chromatin architecture. Transcriptional functioning of TF complexes is regulated further by trans-interacting events such as post-translational RNA and protein modifications and interactions with non-coding RNAs (ncRNAs). In this manner, TF complexes control gene expression by sensing and regulating chromatin architecture, and initiating RNA transcription, splicing, elongation and RNA modifications, as well as initiating feedback control of these very same processes. These regulatory steps result in precise and tunable protein expression in robust signaling networks that are central to homeostasis. Given these diverse functions, TF complexes are large, often being megadaltons in size, and the composition is highly dynamic. In addition, given the importance of TFs and their interacting proteins it is unsurprising that they comprise approximately 15% of the human protein-coding genome; ~1500 TFs, and ~1500 coregulators [1,2,3]. Clearly, all are not expressed and functioning in every cell.
To illustrate the impact of TF interactions with coregulators, it is informative to consider perhaps one of the best understood TF families, namely, the human nuclear hormone receptor (NR) superfamily, which consists of 48 members. The medical importance of NRs has been appreciated at least since rickets was first described in the 17th century by Daniel Whistler [4]; rickets arises from a lack of bone mineralization due to insufficient signaling via the NR, the vitamin D receptor (NR1I1/VDR) (reviewed in [5]). The expansive literature supports the concept that the NR network forms a major hormonal and environmental signaling conduit to the genome, the output of which is central to the development and homeostasis of many cells and tissues (reviewed in [6]).
An appreciation of the relationship between sex steroids and cancers of the reproductive system was pioneered by the work of Sir George Beatson in the 19th century, who began to define the relationship between estrogen and breast cancer (BRCA) risk [7]. Subsequently, in the 1940s this concept was echoed by the work of Dr. Charles Huggins and colleagues who established the endocrine synthesis of androgens and the relationship to prostate cancer (PRAD) [8]. Reflecting their biological importance and the appreciation of their functions, these TFs are major research targets in the arenas of hormone-dependent cancers, including not only BRCA (reviewed in [9]) and PRAD (reviewed in [10]), but also cancers of the ovaries (OVCA), endometrium (ECa), uterus (UCS), testis (TCGT), thyroid (THCA), pancreas (PAAD) and adrenal glands (ACC).
Outside of cancers, NRs impact other diseases including diabetes and metabolic syndrome; inflammation and inflammatory diseases; aging syndromes including bone health functions and cardiovascular diseases. Therefore, NR-targeting drugs include the well-established specific NR antagonists for BRCA and PRAD, and NR agonists for a wide range of inflammatory disorders and also to target specific leukemias. As a result, approximately 15% of prescription drugs target NRs; furthermore, NR-driven modulation of xenobiotic-metabolizing capacity is central to the clearance of an even bigger proportion of pharmacological drugs [11,12].
Across hormone-dependent cancers there is now a considerable volume of pre-clinical and clinical findings to define how NR-transcriptome mechanism functions are altered in cancer progression and central in response to therapeutic approaches. For example, prostate differentiation and homeostasis is intimately regulated by the androgen receptor (NR3C4/AR). Indeed, in 1972, AR control of cell-fate decisions was the catalyst for defining the field of programmed cell death [13]. Transcriptomic approaches, in the first case using microarrays and subsequently applying next generation sequencing (NGS) approaches, have established that PRAD cells display a remarkable diversity of AR transcriptional responses, and underscored that there are few examples of obligate and universal transcriptional responses regulated by the AR [14,15,16]. Widespread epigenetic events redirect AR genomic binding, including the recommissioning of embryonic enhancers [17]. As a result, in PRAD the AR no longer regulates luminal differentiation and alternative programs are enhanced, leading to novel lineages such as neuroendocrine (NE)-PRAD [16,18,19,20,21,22,23,24,25,26,27,28,29,30,31]. This finding is echoed by the diverse responses of estrogen receptors (NR3A1/ER⍺, NR3A2/ERβ) in BRCA. Many investigators have considered how NR coregulators are altered to explain how NR functions appear to be extremely context-dependent and are redirected through cancer progression [32,33,34,35,36,37,38,39].

1.2. A Primer on Nuclear Receptors and Their Interaction with Regulatory Complexes

The NR superfamily is structurally classified into seven families by sequence homology and, broadly, these receptors comprise three distinct functional groups with regard to their interaction with ligands (reviewed in [40,41]). Type I NRs, also called classical or steroid hormone receptors, include the ERs and AR and interact with high affinity ligands. Other receptors include the progesterone (NR3C3/PR), glucocorticoid (NR3C1/GR) and mineralocorticoid (NR3C2/MR) members. Type IIa NRs include binding non-steroidal ligands, as well as broad affinity ligands, and include all-trans retinoic acid (NR1B1/RARα, NR1B2/RARβ, NR1B3/RARγ), 9-cis retinoic acid (NR2B1/RXRα, NR2B2/RXRβ, NR2B3/RXRγ), peroxisome proliferator activated receptors (NR1C1/PPARα, NR1C2/PPARβ, NR1C3/PPARγ) and VDR. A third class of NRs referred to as Type IIb are orphan receptors with unidentified ligands.
To interact with the genome, Type I NRs form homodimers and Type IIb form heterodimers predominantly with RXRs, whereas Type IIb orphan receptors bind to the genome frequently as homodimers or monomers. The classical view of the cellular distribution for different classes of nuclear receptors is that Type I NRs are sequestered in the cytoplasm in the absence of ligands and not associated with cis regulatory elements (CREs) on target genes. In the presence of ligands, they shuttle into the nucleus to bind CREs and regulate gene expression. By contrast, the Type IIa NRs are nuclear-resident bound to CREs independent of ligands where they exert a repressive effect, and ligand activation generally redistributes them and leads to transactivation. Type IIb orphan NRs are also nuclear resident and can move across the genome in response to various stimuli [40,41]. However, with the development of chromatin immunoprecipitation coupled with NGS (ChIP-Seq) it has become apparent that these views are too rigid and genomic binding of all NRs is detectable in the absence of ligands, albeit to a lesser extent for Type I NRs.
Certainly, NR genomic localization sets the stage for a wide range of NR–protein interactions, many modulated by ligand and protein interactions, and allow NRs to participate in a wide variety of cis- and trans-regulatory mechanisms to positively and negatively regulate transcription [42,43,44,45]. NRs bind to CREs in a protein–DNA complex, whereas NRs can bind indirectly, in trans, to CREs and regulate transcription indirectly in a protein–protein–DNA interaction complex. Other than proteins, non-coding RNAs such as miRNA and long non-coding RNAs (lncRNAs) can alter NR-regulated gene expression by directly affecting the DNA environment or other RNA-binding proteins (RBPs) [46]. For example, the steroid receptor activator (SRA) was one of the first regulatory lncRNAs identified to coactivate a NR, namely the GR [47,48]. The SRA also coimmunoprecipitated with the coactivator NCOA1/SRC1, suggesting a scaffolding function for a wider range of NRs (reviewed in [47]).
Structurally, NRs contain a variable N-terminal domain (NTD), a DNA-binding domain (DBD), a hinge region, a conserved ligand-binding domain (LBD) and a variable C-terminal domain. The two most highly conserved domains amongst all NRs are the DNA-binding domain and the ligand-binding domain. The DBD contains two zinc fingers for genomic binding [49,50]. The LBD comprises a three-layered, antiparallel, α-helical sandwich and within this α-helical core, two additional layers of helices form the ligand-binding cavity. An additional helix for ligand-dependent transcriptional activation (AF2) resides at the C-terminal of the LBD and adopts different positions depending on the presence or absence of ligands [51,52,53,54,55,56,57,58]. This domain is mainly responsible for the interaction with transcriptional coregulators depending on conformational changes induced by different ligands [59,60], as ligand activation induces an allosteric change, thus inducing activation [6,61].
The concept that protein–protein interactions regulate TF functions was established in early studies of yeast by Ptashne and Gann, which identified roles for TF adapter proteins to bring about successful mRNA regulation. These proteins were not considered as constitutive members of RNA pol II, but indirectly associate with TFs and could positively modulate gene expression, and hence, gave rise to the term “activators” [62]. This discovery then led to the idea of protein interactions modulating TF functions.
The canonical view of NR function is that in the absence of ligands the receptor associates with corepressors. Structural analyses of corepressors, such as nuclear receptor corepressor 1 and 2 (NCOR1/NCOR2), revealed they contain an analogous extended helix with LXXH/IIXXXI/L residues forming a common motif called the CoRNR-box (corepressor-nuclear-receptor), which is required for interaction with RXR, TR and other NRs [63]. The CoRNR-box occupies the same hydrophobic pocket contacted by the NR box in the absence of ligands, due to displacement of the AF2 helix [64,65,66,67].
The presence of ligands induces structural LBD conformational changes, dissociates corepressor complexes and exposes protein surfaces for coactivator complexes to bind. These complexes contain chromatin modifier proteins, which further remodel chromatin and facilitate the interaction between the NR complex bound at enhancer/promoter regions of target genes to the Mediator complex [68] and the basal transcriptional machinery at the TSS. Coactivators such as NCOA1/SRC1 interact with NRs through helical motifs that contain an LXXLL-consensus peptide motif, also called the NR box.
Agonist binding reduces the affinity of nuclear receptors for CoRNR-box-containing corepressors and increases the affinity for LXXLL-containing coactivators. In vivo studies demonstrate a swing of agonistic or antagonistic activity for some BRCA treatments, such as tamoxifen. This effect is due to selective recruitment of coregulators to regulatory regions, depending on tissue specific expression level of either coactivator or corepressor [69].
Interactions between NRs and coregulators has been a topic of extensive investigation (reviewed in [70,71,72,73,74,75]). An intriguing question in the coregulator field is over how broad is the definition of their function. Clearly, coregulators that directly bind NRs are readily defined, but it is less clear if this extends to the cofactors that interact in a protein–protein manner with the NR complex, or function upstream of NR signaling. For example, the SWI/SNF complexes that remodel the nucleosome position may not directly inteact with NRs, but nonetheless regulate their function [76,77] and similarly downstream coregulators that facilitate the process of transcription, such as helicases are impactful [78,79]. In this manner, if the definitions of coregulators are extended, this results in a large number of proteins impacting NR coregulator proteins [80]. A working model of NR and coregulator functions is shown in Figure 1.

1.3. Functions Played by Nuclear Receptor Coregulators

Previously, we used bioinformatic approaches to develop [1] unique lists of coactivators, corepressors and mixed function coregulators that sum to more than 1500 proteins. Testing the enrichment of these components in NR complexes reveals the diversity of components. We have examined coregulator enrichment in NR-dependent RIME data [81], for example, to capture all VDR-interacting proteins, and identified at least 100 interacting proteins, many of which are classified as coregulators; thus, reflecting the findings of others [1,82,83,84,85]. This diversity of coregulator interactions, coupled with their dynamic sharing and sequestration, has been proposed to be a critical aspect of NR network behavior [86]. On the basis of function, coregulators can contribute to the control of the following events:
a. Nucleosome positioning. In humans the switch/sucrose non-fermenting (SWI/SNF) ATP-dependent chromatin remodeling complex, known as the BAF complex, includes coregulators that change nucleosome positioning through ATP-dependent helicases. These ATP helicases include various domains to bind chromatin; bromodomains (identify acetylated lysine), chromodomain helicase DNA-binding (CHD), plant homeodomain (PHD) domains (identify methylated lysine) and HAND-SANT-SLIDE domains (identify unmodified histone tails). Recently, another group of mixed function coregulators have been identified, which can act as a coactivator by associating with RNA pol II and corepressor by associating with nucleosome remodeling and the deacetylase (NuRD) complex [87,88].
b. Histone modifications. Histone modifying enzymes, such as acetyl transferases (e.g., KAT8) and histone deacetylases (HDACs) (e.g., HDAC1), alter the net charge of the specific residues on the histone tails within the nucleosome, thereby loosening or tightening the DNA–histone interactions. Activating modifications are typified by those on histone 3 (H3), such as acetylation of lysine 9 [89,90], which in turn results in gene transactivation. By contrast, depending on the site, histone methylation can both activate and repress transcriptional states of the chromatin. Coregulator components can recognize specific histone modifications, such as the SANT motif in the case of the NCOR2 complex [91]; thus, the genomic interactions of coregulators most likely reflect NR-targeted loci recruitment coupled with nucleosome affinity. Coactivator-associated arginine methyltransferase 1 (CARM-1), which is also known as protein arginine methyltransferase (PRMT-5), is a histone methyltransferase acting at H3R17, was identified initially by its ability to interact with NCOA coactivators. Due to its indirect recruitment to ligand-bound NRs, including the AR, CARM-1 has been classified as a secondary coactivator [92].
c. Post-translational protein modifications. More broadly, coregulators such as ubiquitin protein ligase E3A and CHY zinc finger domain-containing 1 demonstrate E3 ligase activity, which is implicated in the modulation of NR-driven regulation. E3A interacts with AR in a hormone-dependent manner and causes hormone-dependent recruitment to the promoter region of AR-target genes, leading to enhancing the transactivation function of the AR [93]. Similarly, SUMO homologs have been shown to affect AR-mediated transcription. SUMO-1 decreases, whereas SUMO-2 and -3 enhance AR transcriptional activity [94,95]. SUMO-3 interacts with the AR NTD and mutations in sumoylation sites of NTD are reported to increase AR transactivation. The positive effect of SUMO-3 on AR-transcriptional activity depends on stimulation by androgens and is independent of sumoylation sites of the AR or sumoylation by SUMO-3 [94].
d. RNA splicing and metabolism. The pre-mRNA splicing proteins polypyrimidine tract binding protein 1, paraspeckle protein 1 (PSP1) and PSP2 interact with AF-1 of the AR [96] and the non-POU domain-containing octamer-binding, another component of the U1snRNP pre-spliceosome, interacts with the AR NTD in a ligand-dependent manner and potentiates AF-1 function [96]. The metabolism of the cell can also impact NR function as powerfully illustrated by the significant literature on PPARs [97,98,99,100,101] and the PPAR coregulator PPARGC1a [1,102,103] in PRAD.
e. DNA repair. The protein kinase DNA-activated catalytic subunit (PRKDC) complex is involved in DNA repair and has emerged as a part of the transcriptional machinery. The AR-LBD interacts directly with the Ku70 and Ku80 regulatory subunits of PRKDC in a DNA-dependent manner. Ku70 and Ku80, as well as PRKDC, enhance AR activity in transactivation assays through the recycling of transcriptional factors [104].
f. Chaperones. Steroid hormone NRs bound to heat shock proteins (HSPs) are maintained in a stable and largely inactive configuration. For example, in the early stages of the AR activation process, AR-LBD interacts transiently with various HSPs, leading to an equilibrium in which the AR is maintained in an overall high-affinity ligand-binding state. Loss of the HSP70 cochaperone DNAJ HSP member A1 (DNAJA1) in a knockout (KO) mouse model leads to increases in AR protein levels and enhanced transcription of several androgen-responsive genes in Sertoli cells, and indeed, can be considered as a negative regulator of transactivation by the AR [105].
g. Signal integrators and transducers, scaffolds and adaptors. Several cellular scaffold proteins have coregulatory properties. For example, the endocytic coregulator protein huntingtin-associated protein 1 (HAP1) interacts with the AR-LBD [106]. In addition, caveolin-1 also interacts with the AR in the presence of a ligand [107]. Other proteins, such as the transforming growth factor beta 1 induced transcript 1 (TGFβ1I1), are involved in cell–cell contacts and also bind to AR-LBD and enhance transcriptional activity [108]. Indeed, protein tyrosine kinase 2 beta phosphorylates TGFβ1I1, which in turn represses AR transactivation [109]. The TGFβ1|1-related protein paxillin (PXN) also localizes within focal adhesions and can participate in AR signal transduction pathways. PXN directly interacts with the AR, and overexpression of PXN results in increased shuttling of the AR to the nuclear matrix [110]. Similarly, the TACC family paralogs are known to interact with histone acetyl transferases [111] and RAR complexes [112,113], and also participate in microtubule integrity [114]. Dysfunction of TACC1 is implicated in several cancers including BRCA [115,116,117], but is under-explored in PRAD, although it is associated with worse disease-free survival [118].

2. Examples of Nuclear Receptor Interactions with Coactivators

In Table 1 are examples of coactivator genes, their superfamily they belong to, NR interactions and the frequency of structural variation in several hormone dependent cancers.

2.1. SWI/SNF (BAF) Complex

Highly dense chromatin requires nucleosome remodeling to enable TF access. There are four families of nucleosome remodelers in eukaryotic cells, namely, SWI/SNF, ISWI, CHD and INO80 [138], and most eukaryotes contain at least one of these four. The SWI/SNF complexes are multi-subunit, ATP-dependent chromatin remodelers that mediate nucleosome dynamics. He et al. demonstrated that the catalytic subunit of SWI/SNF, SMARCA4, binds to the base of the nucleosomes and the ATPase modules [139]. Other core subunits are either the Brahma (BRM) or Brahma-related gene 1 (BRG1) that act as the ATPase core, bind to acetylated histones, have been known as tumor suppressors and potential binding partners for VDR [140]. More recently, we revealed that BAZ1A, a SMARCA5 binding partner, governs VDR gene regulation [83]. Many mutations and the functional disruption of the SWI/SNF complex have been reported from cancer genome sequencing in 25% of all cancer types [141] and mutations in SMARCA4 have been reported in BRCA, PRAD, OVCA and PAAD [142,143]. Both SMARCA4 and its paralog, SMARCA2, are identified as tumor suppressors or drivers of cancer and emerging therapeutic approaches are targeting these genes [144,145,146]. These SMARCA mutations appear to impact the expression of genes including NRAS, KRAS and TP53 [147].
Key questions are whether SWI/SNF subunits are recruited to specific loci through their direct or indirect interactions with NRs and if they are essential for all NR-mediated transcriptional activation [76,148]. For example, NR4A1/NUR77 interacts with SWI/SNF subunits, mediates an opposing role to PPARγ and promotes AR-dependent transcription [76,149]. In PRAD (LNCaP cells), the expression of AR-target genes requires the recruitment of SWI/SNF complex proteins to proximal and distal regions or AR CREs in the presence of DHT [150]. Moreover, studies on BRCA relapsed tumors highlight the importance of SWI/SNF-related genes, such as ARID1 and ARID2, in chemotherapy and survival outcome of patients with ERα positive BRCA [151,152]. Other studies have shown the importance of the SWI/SNF complex for ligand-dependent transcriptional activation of the glucocorticoid receptor (GR) using a hormone-responsive mouse mammary tumor virus (MMTV) promoter [153].

2.2. NCOA/p160/SRC

The earliest interaction discovered between NRs and regulatory proteins involved NCOA proteins [154], which belong to the P160/SRC family. The isolated cDNA of NCOA1/SRC1 was first identified through a yeast 2-hybrid assay for PGR [155] and was mapped to the 2p23 chromosome [156]. Similarly, in BRCA, frequent copy number gains were shown to encapsulate amplified in breast cancer 1 (AIB1), which was subsequently identified as NCOA3 [157,158].
Structurally, the NCOAs are composed of the basic helix–loop–helix-PER-ARNT-SIM (HLH-PAS) domain on N-terminus, a serine/threonine-rich (S/T) domain and a nuclear receptor interaction domain (NRID) in the middle and two transactivation domains (AD) on the C-terminus [159,160]. To function as coactivators, NCOA1-3 associates with histone acetyl transferase proteins such as p300/CBP, p/CAF and the Mediator complex [154].
Overexpression of NCOA1, NCOA2 and NCOA3 has been reported widely in hormone-dependent cancers, including in BRCA, PRAD and OVCA [161]. NCOA3 splice variants are also potent in BRCA. NCOA3/AIB1Δ4 with truncated exon 4 has recently been shown to drive aggressive BRCA, in part by altering enhancer usage [162]. RNAi-mediated depletion of NCOA3 revealed that it is important in the proliferation of hormone-dependent and -independent PRAD and BRCA cells [163,164,165,166,167]. Moreover, the opposing function of NCOA1 and NCOA2 were shown in KO animal studies, in which they differentially altered the PPARγ2 transcriptional capacity [168].
High expression of these coactivators has also been linked to relapse after hormone-deprivation therapies by activating the mitogen-activated protein (MAP) kinase pathway that, for example, results in both ERα and NCOA3 phosphorylation [166,169,170,171,172,173,174]. In turn, recent studies have suggested that NCOA1 phosphorylation initiates cascades that activate the STAT3 pathway leading to therapy-resistant BRCA [173,175,176,177]. Additionally, NCOA3 recruitment to ERα at specific enhancers/promoters is disrupted in MCF-7 cells that are resistant to tamoxifen [178,179].
Perhaps unsurprisingly, in PRAD cells, knockdown of NCOA1 impacts AR capacity to regulate target gene expression. Specifically, in androgen-responsive LNCaP cells and C4-2 cells, which are androgen deprivation therapy recurrent variants of LNCaP, depletion of NCOA1 results in reduced growth and lower expression of AR-regulated genes. On the other hand, AR-negative PRAD cell lines PC3 and DU145 display no impact of NCOA1 depletion, supporting a role for NCOA1 to control AR function [180]. Reflecting these cancer-driver effects, NCOA3 has emerged as an attractive target for novel cancer therapeutics [181].

2.3. Transforming Acidic Coiled-Coil Containing Protein 1 (TACC1)

TACC1 belongs to the TACC family and has two other paralogs in humans, TACC2 and TACC3. Genetically, all TACC family proteins have different splice variants; for example, the TACC1 gene has 10 different transcriptional splicing variants with isoforms showing tissue-specific differences. Initially, TACC1 and TACC2 were detected to be involved in the progression of solid tumors [182], whereasTACC3 KO in a mouse causes a lethal embryonic phenotype and increases apoptosis [183].
TACC1 is located on chromosome 8p, a common amplicon in BRCA [184] that also contains MYC, CCND1 and ERBB2 [116]. In metastatic BRCA, the binding of an RNA-binding protein, called muscle bind-like 1, to 3`UTR of TACC1 suppressed metastasis by increasing the stability of the mRNA [117]. TACC1 functions to stabilize microtubules, control mRNA maturation and chromatin remodeling for unliganded NRs (including TR and RARs) [185,186,187] and alter their target gene expression [113] such that depleting TACC1 causes a reduction of RARα and TR ligand-dependent transcriptional activity. This broad list of TACC functions reflects the large number of protein-binding domains and underscores their roles in gene expression regulation.
Early studies detected the oncogenic nature of the TACC family with specific splice variants that promote breast tumorigenesis in mouse models [188,189,190], and expression levels that differ in specific cancer types; for example, being frequently overexpressed in BRCA, but with both up- and down-regulated expression in PRAD [118,185]. In PRAD, reduced expression of RARγ leads to increased cell proliferation by unopposed AR signaling; thus, targeting RARγ is therapeutically attractive [191,192]. Our studies indicate miR-96 targets RARγ and TACC1 and associates with aggressive PRAD tumors [118].

2.4. PPARG Coactivator 1 Alpha (PPARGC1A)

PPARGC1A was first cloned in a mouse in 1998 and possesses a LxxLL motif and a serin- and arginine-rich domain with a PPARγ-binding site in between them and an RNA-binding site on its C-terminus [193]. This protein acts as coactivator for NRs such as PPARγ, AR, ERα, ERβ and TR, and controls a variety of mechanisms, including those involved with mitochondrial biogenesis [194]. The activity of PPARGC1A is regulated through post-translational modifications by acetyltransferase and deacetylase (GCN5 and SIRT1, respectively) with the acetylated form being inactive [195,196]. Generally, the expression level of PPARGC1A depends on the energy demand of the cell, and therefore, acts as master regulator of mitochondrial biogenesis. Other pathways such as adaptative thermogenesis, glucose and fatty acid metabolism and circadian rhythm are coordinated through the PPRGC1A coactivator [197].
Unsurprisingly, given the roles in energy regulation, PPARGC1A impacts tumor progression and growth, but the precise mechanisms are cancer-specific. PPARGC1A is overexpressed in BRCA, helping in survival, proliferation and stem cell maintenance, whereas in PRAD there are examples where the reduced expression of PPARGC1A is associated with aggressive features [1,103,198,199]. However, in other settings, such as PPARγ overexpression, cooperation with PPARGC1A increases AKT levels which in turn results in mitochondrial biogenesis that drives PRAD progression [97]. Furthermore, in the BRCA cell lines MCF7 and MDA-MB-231, overexpression of PPARGC1A induces apoptosis and inhibits glycolysis metabolism [200]. Additionally, PPARGC1A exerts its function through the regulation of estrogen-related receptor alpha (ERRα) activity and inhibits PRAD metastasis with metabolic reprogramming of the cells [103]. Interestingly, the genomic analysis on metastatic PRAD samples is in line with the idea of a selective pressure on PPARGC1A deletion as the disease progresses [201,202,203]. The conditional deletion of Pgc1α in a murine model’s prostate epithelium together with Pten, a tumor suppressor, increased the prostate mass and invasiveness of cancer. Rewiring the metabolic landscape of the PRAD cells is possible through the reprogramming of the signaling pathway and PPARGC has been reported to coordinate complexes, although the detailed mechanism has yet to be studied [103].

3. Examples of Nuclear Receptor Interactions with Corepressors

In Table 2 are examples of corepressor genes, their superfamily they belong to, NR interactions and the frequency of structural variation in several hormone dependent cancers.

3.1. RE1-Silencing Transcription Corepressor (REST) Complex 1 (RCOR1)

RCOR1 contains two SANT domains separated by 191 aa [226]. RCOR1 is part of a complex with HDAC1, -2, BHC80, high mobility group 20B (HMG20B) and lysine demethylase 1A (KDM1). KDM1A identifies and causes demethylation at histone 3 lysine 4 (H3K4), leading to gene repression, although demethylation of H3K27 leads to gene activation at H3K4 and K27 [227,228]. KDM1A’s assembly into the RCOR1 complex promotes demethylation of H3K4me2 at active loci and the subsequent recruitment of MECP2 to facilitate CpG island methylation and the generation of a repressive chromatin structure [228,229].
For example, this complex mediates AR actions on gene repression. Imaging has revealed REST and AR closely colocalize in vivo [206]. Another mechanism of REST-mediated transcriptional repression involves chromodomain on Y-like (CDYL), which acts as a REST corepressor that physically bridges REST and the histone methylase G9a to repress transcription [207]. RCOR1−/− mice are profoundly anemic and die in late gestation. Mutant erythroid progenitor cells derived from knockout mice form myeloid colonies instead of erythroid colonies, indicating a change to the myeloid phenotype [230].
Forkhead box (FOX) A1 recruits the KDM1A/RCOR1 complex and HDAC-1/2 to the AR-regulated enhancers and suppresses basal transcription in an AR-independent manner [25482560]. AR and REST repressive complexes may also contribute to repressing the transcriptional regulation of the serine peptidase inhibitor, Kazal type 1 (SPINK1) [204], and indeed, the AR and REST function as transcriptional-repressors of SPINK1 and AR antagonists alleviate this repression leading to SPINK1 upregulation, which in turn is linked to transdifferentiation, stemness and cellular-plasticity [204].

3.2. NCOR1 and NCOR2 Corepressor Complex

NCOR1 and its paralog NCOR2/SMRT were discovered separately in 1995 as part of efforts to explain gene repression by NRs. NCOR1 and NCOR2 bind to TR and RARs to inhibit the expression of their target genes in the absence of ligands [231,232]. Ncor1−/− mice with targeted deletion of the 5′ SIN3 interaction domain exhibit defects in the development of erythrocyte, thymocyte and neural events [233]. Tissue specific deletion of NCOR1 in muscles leads to enhanced exercise endurance due to an increase in muscle mass and number of mitochondria, indicating that loss enhances the functioning of muscles, as a result of its impact on PPARs signaling [234]. Ncor2−/− mice have an impaired development of the forebrain, and the corepressor is required for maintenance of the neural stem cell state [235].
Both NCOR1 and NCOR2 have N-terminal repressor domains (RDs) that recruit and associate with additional proteins, forming a large corepressor holocomplex. For example, RDs interact and recruit SANT-like domains (Swi3, Ada2, Ncor1, and TFIIB) to synergistically promote deacetylation through a deacetylase activation domain and histone interaction domains on histone tails through C-terminal RIDs that tether the corepressor holocomplex to the NR partners on target genes [236,237,238,239]. HSPA8 and TCP1 chaperones occlude the active site for HDAC3. NCOR1 and NCOR2 corepressor complexes interact with HDAC3, 4, 5 and 7 through the deacetylase domain of the NCOR1–NCOR2 complex, leading to the expulsion of two chaperone proteins and freeing the active site of the HDAC [240]. Mechanisms have emerged whereby activated TFs can recruit NCOR1 and NCOR2, leading to active gene silencing. For example, ligand-dependent sumoylation of the PPARγ ligand-binding domain recruits the NCOR1 and HDAC3 complex to drive transrepression [241]. In turn, this prevents NCOR1 recruitment to the ubiquitylation/19S proteasome and promoter clearing.
These corepressors undergo significant alternative splicing to generate protein variants that determine function [242,243]. For example, NCOR splice variants RIP-13α and RIP-13𝛿1 directly interact with TFIIB, TAFII32 and TAFII70 [244]. The most prominent alternative-splicing events occur at the receptor interaction domain (RID)-3 and alters the affinity for NR partners [242,245]. Furthermore, NCOR2 is regulated by mitogen-activated protein kinase kinase kinase 3 (MAP3K3) cascades that induce its release from its receptor partners, its export from nucleus to cytoplasm and its derepression of target gene expression. Cells can customize their transcriptional response to the MAPKKK cascade signaling by selective corepressor splice variant expression, and by selective exploitation of specific tiers of the MAPK cascade [246,247,248,249].

3.3. SIN3 Transcription Regulator Family Member A (SIN3A)

The SIN3 gene was first cloned in S. cerevisiae. SIN3A has four paired amphipathic alpha-helix (PAH) motifs which are imperfect repeats of ~100 aa residues [250,251,252]. Based on a primary sequence analysis, each PAH motif was proposed to contain two α-helices; one exclusively hydrophilic, the other hydrophobic, similar to the MYC family of transcriptional factors within the helix–loop–helix protein dimerization domains even without possessing significant homology at the amino acid level, indicating Sin3’s potential for a multitude of protein–protein interactions [253]. SIN3A+/− mice have normal expression of SIN3A, whereas SIN3A−/− do not survive post-embryonic day 6.5, indicating the essential role of SIN3A in early development at the peri-implantation stage [254].
SIN3A is capable of multiple protein–protein interactions. SIN3 is devoid of intrinsic DNA-binding abilities and its ability to target chromatin comes through direct interactions with the DNA-binding TFs, such as Mad1 or KLF repressors, and indirect interactions with adaptor proteins including NCOR1/2 [255]. It is thought that the SIN3 proteins serve as a scaffold on which corepressor complexes assemble and recruit HDAC activity. The TGIF interacts with the DBD of AR, facilitating the binding of SIN3A and HDAC1 to form a repression complex and resulting in deacetylation of the core histones. Repression of AR-mediated transcription by TGIF indicates a potentially significant mechanism to control androgen-induced cell growth, which may play a critical role in the growth and development of normal prostate tissue, and in the development and progression of PRAD [216]. RNA splicing factors, splicing factor proline- and glutamine-rich (SFPQ) and non-POU domain-containing octamer-binding (NONO) interact and attract SIN3A synergistically to form protein complexes with the AR in a ligand-independent manner and inhibit its transcriptional activity [256].
The CCCTC-binding factor (CTCF) comprises autonomous silencing domains that mediate transcriptional repression when tethered to a promoter sequence [257,258]. At least one of these domains, the zinc-finger region of CTCF, binds to SIN3A without binding to NCOR1/NCOR2 and recruits HDAC [257]. Two SIN3A domains interact independently with the CTCF zinc-finger cluster. The ability of CTCF regions to retain deacetylase activity is correlated with the ability to bind to SIN3A and to repress transcription. Taking these results together, the synergy in repression mediated by the NR, TR and CTCF is achieved by the binding of multiple molecules of SIN3A to the TR/CTCF–DNA complex, thus providing a large platform for the recruitment of histone deacetylases [259].
SIN3A also acts as a regulator of gene expression, survival, and growth in Erα-positive BRCA cells [260]. For example, reduction of SIN3A expression by RNA interference specifically inhibits estrogen-induced repression of ERα [261]. Furthermore, the SIN3A corepressor complex interacts with a long non-coding RNA NEAT1 to physically associate with FOXN3 in ER+ BRCA cells. The FOXN3–NEAT1–SIN3A complex represses genes that are critically involved in EMT, including GATA3. The SIN3A–FOXN3–NEAT1 complex transrepresses ERα itself, forming a negative-feedback loop in transcription regulation [214]. The SIN3A/ZNF704 corepressor complex also regulates the circadian clock and, in this manner, is a further potential driver for breast carcinogenesis. ZNF704 is a transcriptional repressor that mediates its actions by associating with the SIN3A complex. The ZNF704/SIN3A complex represses a panel of genes, including PER2, that are critically involved in the function of the circadian clock. ZNF704 promotes the proliferation and invasion of BRCA cells in vitro and accelerates the growth and metastasis of BRCA in vivo [212].
SIN3A is able to interact with other factors outside its PAH domains. For example COPS2/ALIEN-mediates silencing through the recruitment of SIN3 to its N-terminus and interacting with DNA-bound transcriptional silencers at the C-terminus. This functional interaction has been demonstrated with VDR on the human CYP24 promoter containing VDRE [262]. COPS2 also interacts in a hormone-sensitive manner with TR to function as a corepressor [263].

3.4. Ligand-Dependent Nuclear Receptor Corepressor (LCOR)

The LCOR is a transcriptional corepressor, first cloned from a fetal brain, with high expression in the liver, testis, ovary, kidney and brain. The LCOR is recruited to agonist bound-NRs through a single NR box. The LCOR comprises a NR motif, a nuclear localization signal, tandem PXDLS motifs and a HTH domain [264].
The LCOR binds to the LBDs on the ERα/β, AR, GR, RAR α/β/γ, RXRα, PPARγ, VDR and PGR in a ligand-dependent manner and acts to attenuate signaling by agonist-bound receptors [220,265,266,267,268]. LCOR-mediated repression of ERα and GR is inhibited by use of the HDAC inhibitor trichostatin A (TSA), but repression of VDR and PR does not happen in the presence of TSA, which indicates that the LCOR functions as a corepressor in an HDAC-dependent and independent manner. LCOR N-terminal PXDLS motifs recruit CTBP 1 and 2 corepressors, which are responsible for their corepressor functioning [221]. LCOR and CTBP1 colocalize in nuclear bodies which also contain the CTBP-interacting protein (CTIP) and polycomb group repressor complex marker BMI1. LCOR function is strongly dependent on the HTH domain, and its deletion completely abolishes the LCOR corepressor function. The LCOR, CTBP and CTIP are recruited endogenous PR- and Erα-stimulated genes in a hormone-dependent manner. SiRNA-mediated knockdown of the LCOR/CTBP1 helped in the expression of PR- and Erα-stimulated reporter genes as well as in endogenous PR-stimulated target genes, whereas complete deletion resulted in gene-specific effects on Erα-regulated transcription, which mostly was reduced gene expression. Thus, the LCOR and CTBP1 usually act as repressors of PR-related transcription, but in some cases can enhance the transcription of some genes [266].
In BRCA, the LCOR has two mRNA variants called LCOR and LCOR2. Both LCOR and LCOR2 inhibit estrogen-mediated signaling and BRCA cell proliferation via the HTH domain. Both the LCORs bind with NRIP1 via the LCOR HTH domain and C-terminal domain (CTD) of NRIP1 and this interaction is crucial for inhibition of gene expression and cell proliferation [269].
The LCOR acts as a corepressor for the AR and inhibits PRAD in vivo. The LCOR interacts with the AR and is recruited to chromatin in the presence of AR. However, the NR box of LCOR is dispensable for interaction with the AR and LCOR and instead interacts with the DBD of the AR on chromatin. In androgen-independent PRAD, the interaction between the LCOR and AR is inhibited by signaling pathways involving SRC [220]. The CTD of the LCOR interacts with KLF6 in multiple cancers and they bind together to the promoters of genes, including CDKN1A. The LCOR contributes to repression mediated by KLF6 in a promoter- and cell type-dependent manner. SiRNA-mediated knockdown of LCOR, CTBP1 and KLF6 in PRAD cell line PC3 induced expression of CDKN1A, CHD1 and other KLF6 target genes [270].

4. Cointegrators

To complement these roles for coactivators and corepressors, it has become clear that these protein complexes link to several co-called cointegrator complexes. For example, the CREB-binding protein (CBP) was initially identified as a coactivator of multiple transcriptional activators, including p53, NF-𝜅B and NRs [271]. The CBP forms a ternary complex with NCOA family members and NRs; for example, the CBP synergizes with NCOA1 in the transactivation of ERα and PR [272]. Direct interactions between the CBP and RXRs and TR occur via the N-terminal of CBP, which comprises the NR box required for receptor interaction [273,274]. The interaction between NCOA1 and unliganded ER and PR are relatively weak in comparison to the recruitment of other NCOA1 complex proteins to ER and PR [275,276]. Overall, it indicates that the CBP acts as a common limiting cofactor for diverse TF/coactivator complexes, acting as a cellular cointegrator to collate multiple different signals into an integrated response at promoters containing multiple cis-acting elements [274]. Similarly, activating signal cointegrator-2 (ASC-2) has an indirect binding site for the AR, in addition to two LXXLL motifs which interact with other NRs. ASC2 stimulates AR transactivation, whereas the retinoblastoma protein hinders transcriptional activation of the AR. ASC2 thereby has three distinct nuclear receptor interaction domains [277].

5. Disrupted Coregulators Function in Therapy-Resistant Hormone-Dependent Cancers

A number of different selective antiestrogens (e.g., tamoxifen) and aromatase inhibitors (e.g., anastrozole) have been developed for hormone-sensitive ER+ BRCA. Drugs with a combined activity of aromatase inhibition and ER degradation activity, such as fulvestrant, have also been approved against ER+ and HER2- advanced BRCA. Resistance to these therapies in BRCA is attributed to the retention of ER signaling, which is mediated by the interactions between activated ERα and coregulator proteins [278,279,280,281].
A role for EZH2 has emerged, which mediates epigenetic silencing of the ERα cofactor GREB1 and which causes ERα coregulators to reallocate on chromatin to drive the conversion of antiestrogen into an agonist [282]. Furthermore, adaptive mutations in the ligand-binding domain of ERα in response to endocrine therapy render them with high transcriptional activity even in the absence of estrogen [283,284]. This happens mostly due to the ability of the mutant ER proteins to interact with coregulator proteins to further contribute to endocrine therapy resistance [285]. Mutations in ERα are also associated with estrogen insensitivity as they affect the interaction between ligands and the recruitment of coactivators, such as NCOA1, NCOA2 and NCOA3, to drive metastatic BRCA [286]. A mutant ERα recruits coactivators in the absence of estrogen to confer antiestrogen resistance by giving rise to an altered antagonist state that resists inhibition within the ERα helix 11 and helix 12 [287].
More widely, other coactivators involved in tamoxifen resistance include TIMELESS, which binds to ERα to enhance its transcription and when overexpressed it increases PARP1 expression and ER-induced gene regulation through its proximal LXXLL motifs [288].
Strong parallels exist in the emergence of ADT-R-PCa in response to drugs, including enzalutamide, fulvestrant, flutamide and darolutamide. In half of the PRAD patients, AR gene amplification has been detected among the most frequent genomic changes for recurrent cancer [203,289]. NCOA2/SRC-2 mRNA levels are elevated in 20% of primary and 63% of metastatic PCA, which are associated with worse recurrent-free survival, whereas sensitivity to bicalutamide can be restored by the genetic KO of SRC-2 [290]. Various coregulators are implicated in changing cells’ response to therapy. The AR coregulator, the serum response factor (SNF) inhibitor, can modulate the cell response to enzalutamide in hormone-resistant cells [291].
By contrast, 5–10% of recurrent PRADs have loss-of-function mutations in corepressors, such as NCOR1 or NCOR2, that enhance the enzalutamide resistance. Altered expression of NCOR1 and NCOR2 has been reported across several different hormone-dependent cancers including BRCA, PRAD and TRC [32,33,292,293]. It is unclear what drives NR-selectivity, with NCOR1 regulating PPARγ and elevated NCOR1 expression, and activity-distorted PPARα/γ gene targets involved in cell cycle control [33]. NCOR2 has some selectivity for the VDR, as NCOR2 mutants defective in the VDR interaction domain are unable to repress endogenous VDR-target genes [294]. Elevated NCOR2 levels in PRAD cells also disrupt VDR signaling, resulting in suppression of target genes associated with antiproliferative action and 1α,25(OH)2D3 sensitivity [295]. By contrast, elevated NCOR1 mRNA in BRCA cells correlates with the suppressed regulation of VDR-target genes and the ability of cells to undergo arrest in the G(1) phase of the cell cycle. A similar increased ratio of corepressor mRNA to VDR occurred in matched primary tumors and normal cells, noticeably in ERα negative tumors. Combinations of vitamin D3 compounds with TSA restored VDR antiproliferative signaling to levels which were indistinguishable from MCF-12A cells [32].
A reduction in NCOR1 levels causes bicalutamide resistance in LNCaP cells and murine PRAD in vivo [38]. PRAD cell lines treated with HDAC inhibitor sodium butyrate (NaB) form the AR-NCOR2/NCOR1 complex binding on the PSA promoter, indicating that AR and NCOR1/2 corepressors may form a stable complex in vitro and NaB may facilitate the interaction between the AR and NCOR2 [296]. In PRAD cell models, loss of NCOR2 leads to methylation at sites distal to transcription start sites and CpG islands, which phenotypically contributes to a neuroendocrine phenotype [63]. More recently, it has been reported that the reduced expression of NCOR2 increases PRAD progression and shorter disease-free survival and accelerates disease progression following androgen deprivation, suggesting a more direct role to regulate the AR [63].

6. Technology and Computational Approaches to Discover Coregulator and Nuclear Interactions

The central role of coregulators and TF interactions is to fine-tune and integrate endogenous and exogenous signals to impact cell-fate decisions. The regulatory interactions are highly specific and depend on the structural and physico-chemical characteristics of protein domains. There are a myriad of coregulator-dependent mechanisms of altered gene expression patterns in hormone-dependent cancers, and it is challenging to define cancer-driver mechanisms with these alterations. Germline and somatic structural variations alter TFs and coregulators, and the regulatory regions they bind [297,298]. In parallel, alterations to the epigenome [30] and 3-D genome [299,300,301] change the accessibility and responsiveness of gene promoter/enhancer regions. However, altered TF-enhancer relationships have potential to be targeted through selective epigenetic therapies [302,303,304], or the downstream gene networks may be uniquely drug-sensitive [305,306]. Furthermore, disease-specific enhancers provide the rationale for targeted tumor deep sequencing to dissect the interactions of germline and structural variation [307,308]. Therefore, there is considerable interest in identifying and dissecting TF-coregulator functions in hormone-dependent cancers.

6.1. Two-Hybrid Assay

One of the most utilized techniques that facilitated the discovery of protein–protein interactions is a two-hybrid assay that resulted in the discovery of NCOR1, NCOR2, NCOA1/SRC1 and many more coregulators. This technique requires using a reporter gene for expression control and fusing both the bait and the prey proteins with the DNA-binding domain (DBD, also called bait) and the transcriptional activation domain (AD, also known as prey), respectively. The two-hybrid assay is one of the basic low-throughput experiments developed in yeast, but now it can be utilized in mammalian cells as well [309,310], leading to the discovery of many VDR, RARα and AR coregulators [294,311]. Recently, Yachie N et al. have reported a two-hybrid assay capable of screening a full matrix of proteins in a single multiplexed strain pool. The advanced high-throughput screens have recently been developed to study the entire genome. Two of the categories are matrix-based and library-based techniques.
In the matrix-based, different prey protein-expressing clones will be plated in different wells, and afterwards they will be mated with bait strains in order to form chimeric diploids. Depending on the reporter gene expression, the chimeric colonies will be selected. For example, an RARα interaction with NCOR1 is one of the many interactions resulting from a high-throughput matrix-based approach [312,313]. In the library-based approach, a comparison between each bait and a pool of random prey (cDNA fragments) is made and the resulting interactions will be evaluated with DNA sequencing.

6.2. Affinity Purification Followed by Mass Spectrometry (MS)

The tandem affinity purification (TAP) method uses the principle of tagging a target protein with a TAP tag, and then this fused protein gets expressed in yeast to form the typical complexes in the cells. After two steps of IgG matrix-protein bond cleaving and calcium calmodulin-coated beads incubation, the fragments can be analyzed using MS. This technique has the advantage of a higher-order interaction direction and is not limited to binary information. Some of the insight into NCORs involved in AR regulation is due to this technique [314]. Recently the proximity-labeling MS (APEX-MS) has developed as a complementary approach for the affinity purification assay. APEX-MS is an in-cell labeling method that, upon treatment with hydrogen peroxide biotin-phenol, gets converted into radicals that can form covalent labeling of the proteins in a 20 nm radius [315].

6.3. Rapid Immunoprecipitation Mass Spectrometry of Endogenous Proteins (RIME)

Given that coregulators modulate the NR’s transcriptional activity, scientists have developed a total protein interactome method to understand the role of NRs’ dysregulation in different diseases. RIME allows the study of complexes in their native interactome by formaldehyde fixation and immunoprecipitation of the target protein. In the last step, on-bead digestion reduces the contamination, and the resulting peptides are tested with MS to identify direct protein interactions to the target protein. The advantage of this technique is that there is no need for an extra tag to purify or overexpress the target protein, because RIME utilizes the endogenous proteins and mimics the biologically and clinically relevant interactomes. Complex interactions can be further investigated using the RIME data to model protein networks and the interaction of complexes with each other [316].
Other CRISPR-based approaches, such as CAPTURE [317] or C-BURST [318], have also been developed to measure the protein complex and chromatin interactions in a specific genomic loci. To date, these have not been used to study NRs as they may lack sensitivity and have been most optimally applied to regions such as telomeres or globulin locus where the protein heterogeneity is less and the protein concentration is higher. However, it is likely that these approaches will be refined to identify the specific NR complex components at specific loci.

6.4. Protein–Protein Monitoring Methods

These approaches usually result in low-throughput data collection but can result in more in-depth understanding of two specific protein interactions. Some of these techniques are: fluorescence resonance energy transfer (FRET), X-ray crystallography and NMR spectrometry [319]. FRET involves detecting two fluorophore-tagged proteins in close proximity to each other [320]. However, to skip the spectroscopic labeling of the target protein, surface plasmon resonance (SPR) has been developed to quantitatively identify ligand–receptor interactions while the receptor is immobilized [321].
All aforementioned techniques are examples of protein–protein interaction detection methods and scientists are still developing more advanced techniques. Each method has its advantages and disadvantages, but all of them can open entirely new avenues of research and novel ideas to better understand the protein functions.

6.5. Bioinformatics Approaches to Identifying Coregulators

The first goal of analyses of coregulators from a genome-wide perspective is to establish comprehensive lists of TFs and co-regulator genes. Previously, we applied a text-mining approach to capture gene sets in Gene Ontology (GO) terms that contained phrases including “positive control of transcription”, “negative control of transcription”, “co-activator” and “co-repressor”, as well as other curated collections; for example, Panther, which stands for Protein ANalysis THrough Evolutionary Relationships, and is a classification system that was designed to classify proteins (and their genes) in order to facilitate high-throughput analysis [322], and FANTOM, which was established to undertake functional annotations of the full-length cDNAs that were collected during the Mouse Encyclopedia Project at RIKEN [2]. From these GO terms, the HGNC gene name and ENSEMBL transcript IDs can be retrieved using bioinformatics approaches [323] and combined to gene lists cross-referenced for uniqueness of, for example, coactivators.
Bootstrapping permutation approaches can then be used to test if the proportions of each gene group are altered more than predicted by chance; for example, in TCGA data. For mRNA or protein expression, copy number variation or mutation status, Z-scores of normalized data can be calculated and tested whether the observed alteration frequencies within gene classes (e.g., coactivators) are altered significantly, or a vector of changes for all genes was calculated and the observed values (e.g., mean Z-score of expression) of a given class tested in comparison to all genes detected in each cancer cohort. Random sampling methods such as bootstrapping can be used to simulate the distribution of changes across the genome for statistical comparison [324]. Empirical p-values are then calculated based on the group position relative to the sampling distribution of the genome. Finally, within each group the most altered genes within a class can be focused on for multivariate modeling [1]. Similarly, text-mining approaches to PubMed searches would be able to test how the distribution of published coregulator studies reflects the incidence of their alterations in a given cancer, for example. Our experience to date suggests that the experimental investigation is not evenly distributed across the coregulators that are actually most impactful [1].
Such lists of coregulators can be used to annotate other high-dimensional datasets to investigate enrichment. For example, recently we undertook RIME analyses of NCOR2 [63] to identify interacting proteins, and then by cross-referencing we were able to define which of these proteins were a part of the so-called “long-tail” of oncogenic drivers in PRAD [325]. More broadly, it is an intriguing challenge to identify how cancer phenotypes could emerge from the interactions of multiple relatively rare mutants of coregulators either with strong oncogenic drivers, or by converging on shared complexes, as seen with SWI/SNF complexes [326]. This would reflect the concept of “BRCAness” in which multiple genomic alterations phenotypically converge [327]. For example, we focused on the mutation status in tumors in advanced PRAD [289] and were able to identify coregulators that bind to NCOR2, exert known epigenomic functions, and are mutated in advanced PRAD.
Similarly, the annotation of ChIP-Seq data for either a given NR or a coregulator with a ChromHMM algorithm [328] that defined underlying epigenetic states allows testing of the unique and overlapping genomic distributions in different epigenetic states (e.g., in active enhancers or at poised enhancers), in cells/treatments or in the presence of ROSE algorithm-defined super-enhancers [329,330]. Indeed, testing the overlap of target ChIP-Seq data with comprehensive datasets as contained in Cistrome DB [331] allows co-enrichment testing of hundreds of TF and histone modification ChIP-Seq datasets to reveal the extent of the enrichment with other nuclear receptors or coregulators. RNA-Seq undertaken in parallel treatments can be matched with cistromic data to define cistrome-transcriptome relationships and test their phenotypic associations, for example, by using Kolmogorov–Smirnov tests to examine differences in cumulative distribution plots for cistrome binding sites with respect to the nearest gene, and again using bootstrapping approaches to measure how the specific cistrome relationships associate with gene expression patterns [332].
Finally, to test how coregulators impact correlations between, for example, AR and AR-target genes, partial correlation analyses can be used [33] to test how the expression of coregulators impacts the correlation of AR to DHT-regulated genes and identify how coregulators selectively promote the expression of genes associated with advanced PRAD emergence.

Author Contributions

Conceptualization and supervision, M.J.C. writing—review and editing, H.J., S.H., M.J.C.; funding acquisition, M.J.C. All authors have read and agreed to the published version of the manuscript.

Funding

M.J.C. acknowledges support in part from the Prostate program of the Department of Defense Congressionally Directed Medical Research Programs [W81XWH-20-1-0373; W81XWH-21-1-0850]; Breast program of the Department of Defense Congressionally Directed Medical Research Programs [W81XWH-21-1-0555]; Prostate Cancer UK [RIA18-ST2-022]; and the National Institute of Health Cancer Center Support Grant (P30CA016058) to the OSUCCC The James. S.H. acknowledges support from Pelotonia of a Post-Doctoral Fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siddappa, M.; Wani, S.A.; Long, M.D.; Leach, D.A.; Mathé, E.A.; Bevan, C.L.; Campbell, M.J. Identification of transcription factor co-regulators that drive prostate cancer progression. Sci. Rep. 2020, 10, 20332. [Google Scholar] [CrossRef] [PubMed]
  2. Abugessaisa, I.; Ramilowski, J.A.; Lizio, M.; Severin, J.; Hasegawa, A.; Harshbarger, J.; Kondo, A.; Noguchi, S.; Yip, C.W.; Ooi, J.L.C.; et al. FANTOM enters 20th year: Expansion of transcriptomic atlases and functional annotation of non-coding RNAs. Nucleic Acids Res. 2021, 49, D892–D898. [Google Scholar] [CrossRef] [PubMed]
  3. Wingender, E.; Schoeps, T.; Haubrock, M.; Krull, M.; Donitz, J. TFClass: Expanding the classification of human transcription factors to their mammalian orthologs. Nucleic Acids Res. 2018, 46, D343–D347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ellis, H. Daniel Whistler: English physician who published the first book on rickets in 1645. Br. J. Hosp. Med. 2019, 80, 51. [Google Scholar] [CrossRef] [PubMed]
  5. Carlberg, C.; Campbell, M.J. Vitamin D receptor signaling mechanisms: Integrated actions of a well-defined transcription factor. Steroids 2013, 78, 127–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Mangelsdorf, D.J.; Thummel, C.; Beato, M.; Herrlich, P.; Schutz, G.; Umesono, K.; Blumberg, B.; Kastner, P.; Mark, M.; Chambon, P.; et al. The nuclear receptor superfamily: The second decade. Cell 1995, 83, 835–839. [Google Scholar] [CrossRef] [Green Version]
  7. Paul, J. Sir George Beatson and the Royal Beatson Memorial Hospital. Med. Hist 1981, 25, 200–201. [Google Scholar]
  8. Huggins, C.; Stevens, R.; Hodges, C.V. Studies on prostatic cancer: II. The effects of castration on advanced carcinoma of the prostate gland. Arch. Surg. 1941, 43, 209–223. [Google Scholar] [CrossRef]
  9. Kumar, S.; Freelander, A.; Lim, E. Type 1 Nuclear Receptor Activity in Breast Cancer: Translating Preclinical Insights to the Clinic. Cancers 2021, 13, 4972. [Google Scholar] [CrossRef]
  10. Porter, B.A.; Ortiz, M.A.; Bratslavsky, G.; Kotula, L. Structure and Function of the Nuclear Receptor Superfamily and Current Targeted Therapies of Prostate Cancer. Cancers 2019, 11, 1852. [Google Scholar] [CrossRef] [Green Version]
  11. Santos, R.; Ursu, O.; Gaulton, A.; Bento, A.P.; Donadi, R.S.; Bologa, C.G.; Karlsson, A.; Al-Lazikani, B.; Hersey, A.; Oprea, T.I.; et al. A comprehensive map of molecular drug targets. Nat. Rev. Drug Discov. 2017, 16, 19–34. [Google Scholar] [CrossRef] [PubMed]
  12. Prakash, C.; Zuniga, B.; Song, C.S.; Jiang, S.; Cropper, J.; Park, S.; Chatterjee, B. Nuclear Receptors in Drug Metabolism, Drug Response and Drug Interactions. Nucl. Recept. Res. 2015, 2, 101178. [Google Scholar] [CrossRef] [PubMed]
  13. Kerr, J.F.; Wyllie, A.H.; Currie, A.R. Apoptosis: A basic biological phenomenon with wide-ranging implications in tissue kinetics. Br. J. Cancer 1972, 26, 239–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Reid, G.; Metivier, R.; Lin, C.-Y.; Denger, S.; Ibberson, D.; Ivacevic, T.; Brand, H.; Benes, V.; Liu, E.T.; Gannon, F. Multiple mechanisms induce transcriptional silencing of a subset of genes, including oestrogen receptor α, in response to deacetylase inhibition by valproic acid and trichostatin A. Oncogene 2005, 24, 4894–4907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Waghray, A.; Schober, M.; Feroze, F.; Yao, F.; Virgin, J.; Chen, Y. Identification of differentially expressed genes by serial analysis of gene expression in human prostate cancer. Cancer Res. 2001, 61, 4283–4286. [Google Scholar]
  16. 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] [Green Version]
  17. Pomerantz, M.M.; Qiu, X.; Zhu, Y.; Takeda, D.Y.; Pan, W.; Baca, S.C.; Gusev, A.; Korthauer, K.D.; 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]
  18. Toropainen, S.; Niskanen, E.A.; Malinen, M.; Sutinen, P.; Kaikkonen, M.U.; Palvimo, J.J. Global analysis of transcription in castration-resistant prostate cancer cells uncovers active enhancers and direct androgen receptor targets. Sci. Rep. 2016, 6, 33510. [Google Scholar] [CrossRef]
  19. Massie, C.E.; Adryan, B.; Barbosa-Morais, N.L.; Lynch, A.G.; Tran, M.G.; Neal, D.E.; Mills, I.G. sNew androgen receptor genomic targets show an interaction with the ETS1 transcription factor. EMBO Rep. 2007, 8, 871–878. [Google Scholar] [CrossRef] [Green Version]
  20. Copeland, B.T.; Pal, S.K.; Bolton, E.C.; Jones, J.O. The androgen receptor malignancy shift in prostate cancer. Prostate 2018, 78, 521–531. [Google Scholar] [CrossRef]
  21. Chattopadhyay, I.; Wang, J.; Qin, M.; Gao, L.; Holtz, R.; Vessella, R.L.; Leach, R.W.; Gelman, I.H. Src promotes castration-recurrent prostate cancer through androgen receptor-dependent canonical and non-canonical transcriptional signatures. Oncotarget 2017, 8, 10324–10347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Malinen, M.; Niskanen, E.A.; Kaikkonen, M.U.; Palvimo, J.J. Crosstalk between androgen and pro-inflammatory signaling remodels androgen receptor and NF-kappaB cistrome to reprogram the prostate cancer cell transcriptome. Nucleic Acids Res. 2017, 45, 619–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Volante, M.; Tota, D.; Giorcelli, J.; Bollito, E.; Napoli, F.; Vatrano, S.; Buttigliero, C.; Molinaro, L.; Gontero, P.; Porpiglia, F.; et al. Androgen deprivation modulates gene expression profile along prostate cancer progression. Hum. Pathol 2016, 56, 81–88. [Google Scholar] [CrossRef] [PubMed]
  24. Pomerantz, M.M.; Li, F.; Takeda, D.Y.; Lenci, R.; Chonkar, A.; Chabot, M.; 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]
  25. Lu, J.; Lonergan, P.E.; Nacusi, L.P.; Wang, L.; Schmidt, L.J.; Sun, Z.; Van der Steen, T.; Boorjian, S.A.; Kosari, F.; Vasmatzis, G.; et al. The cistrome and gene signature of androgen receptor splice variants in castration resistant prostate cancer cells. J. Urol. 2015, 193, 690–698. [Google Scholar] [CrossRef] [Green Version]
  26. Chen, Y.; Chi, P.; Rockowitz, S.; Iaquinta, P.J.; Shamu, T.; Shukla, S.; Gao, D.; Sirota, I.; Carver, B.S.; Wongvipat, J.; et al. ETS factors reprogram the androgen receptor cistrome and prime prostate tumorigenesis in response to PTEN loss. Nat. Med. 2013, 19, 1023–1029. [Google Scholar] [CrossRef] [Green Version]
  27. Beltran, H.; Hruszkewycz, A.; Scher, H.I.; Hildesheim, J.; Isaacs, J.; Yu, E.Y.; Kelly, K.; Lin, D.; Dicker, A.P.; Arnold, J.T.; et al. The role of lineage plasticity in prostate cancer therapy resistance. Clin. Cancer Res. 2019, 25, 6916–6924. [Google Scholar] [CrossRef] [Green Version]
  28. Berger, A.; Brady, N.J.; Bareja, R.; Robinson, B.; Conteduca, V.; Augello, M.A.; Puca, L.; Ahmed, A.; Dardenne, E.; Lu, X.; et al. N-Myc-mediated epigenetic reprogramming drives lineage plasticity in advanced prostate cancer. J. Clin. Investig. 2019, 130, 3924–3940. [Google Scholar] [CrossRef]
  29. Sheahan, A.V.; Ellis, L. Epigenetic reprogramming: A key mechanism driving therapeutic resistance. Urol. Oncol. 2018, 36, 375–379. [Google Scholar] [CrossRef]
  30. Ku, S.Y.; Rosario, S.; Wang, Y.; Mu, P.; Seshadri, M.; Goodrich, Z.W.; Goodrich, M.M.; Labbe, D.P.; Gomez, E.C.; Wang, J.; et al. Rb1 and Trp53 cooperate to suppress prostate cancer lineage plasticity, metastasis, and antiandrogen resistance. Science 2017, 355, 78–83. [Google Scholar] [CrossRef] [Green Version]
  31. Bohrer, L.R.; Chen, S.; Hallstrom, T.C.; Huang, H. Androgens suppress EZH2 expression via retinoblastoma (RB) and p130-dependent pathways: A potential mechanism of androgen-refractory progression of prostate cancer. Endocrinology 2010, 151, 5136–5145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Banwell, C.M.; MacCartney, D.P.; Guy, M.; Miles, A.E.; Uskokovic, M.R.; Mansi, J.; Stewart, P.M.; O’Neill, L.P.; Turner, B.M.; Colston, K.W.; et al. Altered nuclear receptor corepressor expression attenuates vitamin D receptor signaling in breast cancer cells. Clin. Cancer Res. 2006, 12, 2004–2013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Battaglia, S.; Maguire, O.; Thorne, J.L.; Hornung, L.B.; Doig, C.L.; Liu, S.; Sucheston, L.E.; Bianchi, A.; Khanim, F.L.; Gommersall, L.M.; et al. Elevated NCOR1 disrupts PPARalpha/gamma signaling in prostate cancer and forms a targetable epigenetic lesion. Carcinogenesis 2010, 31, 1650–1660. [Google Scholar] [CrossRef] [PubMed]
  34. Cheng, S.; Brzostek, S.; Lee, S.R.; Hollenberg, A.N.; Balk, S.P. Inhibition of the dihydrotestosterone-activated androgen receptor by nuclear receptor corepressor. Mol. Endocrinol. 2002, 16, 1492–1501. [Google Scholar] [CrossRef]
  35. Doig, C.L.; Singh, P.K.; Dhiman, V.K.; Thorne, J.L.; Battaglia, S.; Sobolewski, M.; Maguire, O.; O’Neill, L.P.; Turner, B.M.; McCabe, C.J. Recruitment of NCOR1 to VDR target genes is enhanced in prostate cancer cells and associates with altered DNA methylation patterns. Carcinogenesis 2013, 34, 248–256. [Google Scholar] [CrossRef] [PubMed]
  36. Fereshteh, M.P.; Tilli, M.T.; Kim, S.E.; Xu, J.; O’Malley, B.W.; Wellstein, A.; Furth, P.A.; Riegel, A.T. The nuclear receptor coactivator amplified in breast cancer-1 is required for Neu (ErbB2/HER2) activation, signaling, and mammary tumorigenesis in mice. Cancer Res. 2008, 68, 3697–3706. [Google Scholar] [CrossRef] [Green Version]
  37. Lavinsky, R.M.; Jepsen, K.; Heinzel, T.; Torchia, J.; Mullen, T.M.; Schiff, R.; Del-Rio, A.L.; Ricote, M.; Ngo, S.; Gemsch, J.; et al. Diverse signaling pathways modulate nuclear receptor recruitment of N-CoR and SMRT complexes. Proc. Natl. Acad. Sci. USA 1998, 95, 2920–2925. [Google Scholar] [CrossRef] [Green Version]
  38. Lopez, S.M.; Agoulnik, A.I.; Zhang, M.; Peterson, L.E.; Suarez, E.; Gandarillas, G.A.; Frolov, A.; Li, R.; Rajapakshe, K.; Coarfa, C. Nuclear receptor corepressor 1 expression and output declines with prostate cancer progression. Clin. Cancer Res. 2016, 22, 3937–3949. [Google Scholar] [CrossRef] [Green Version]
  39. Hodgson, M.C.; Shen, H.C.; Hollenberg, A.N.; Balk, S.P. Structural basis for nuclear receptor corepressor recruitment by antagonist-liganded androgen receptor. Mol. Cancer Ther. 2008, 7, 3187–3194. [Google Scholar] [CrossRef] [Green Version]
  40. McKenna, N.J.; Lanz, R.B.; O’Malley, B.W. Nuclear receptor coregulators: Cellular and molecular biology. Endocr. Rev. 1999, 20, 321–344. [Google Scholar]
  41. Lazar, M.A. Maturing of the nuclear receptor family. J. Clin. Investig. 2017, 127, 1123–1125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Aranda, A.; Pascual, A. Nuclear hormone receptors and gene expression. Physiol. Rev. 2001, 81, 1269–1304. [Google Scholar] [CrossRef] [PubMed]
  43. Gronemeyer, H.; Gustafsson, J.A.; Laudet, V. Principles for modulation of the nuclear receptor superfamily. Nat. Rev. Drug Discov. 2004, 3, 950–964. [Google Scholar] [CrossRef] [PubMed]
  44. Nettles, K.W.; Greene, G.L. Ligand control of coregulator recruitment to nuclear receptors. Annu. Rev. Physiol. 2005, 67, 309–333. [Google Scholar] [CrossRef]
  45. Weston, A.D.; Blumberg, B.; Underhill, T.M. Active repression by unliganded retinoid receptors in development: Less is sometimes more. J. Cell Biol. 2003, 161, 223–228. [Google Scholar] [CrossRef] [PubMed]
  46. Statello, L.; Guo, C.J.; Chen, L.L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef]
  47. Foulds, C.E.; Panigrahi, A.K.; Coarfa, C.; Lanz, R.B.; O’Malley, B.W. Long Noncoding RNAs as Targets and Regulators of Nuclear Receptors. Curr. Top. Microbiol. Immunol. 2016, 394, 143–176. [Google Scholar]
  48. Lanz, R.B.; Razani, B.; Goldberg, A.D.; O’Malley, B.W. Distinct RNA motifs are important for coactivation of steroid hormone receptors by steroid receptor RNA activator (SRA). Proc. Natl. Acad. Sci. USA 2002, 99, 16081–16086. [Google Scholar] [CrossRef] [Green Version]
  49. Lambert, S.A.; Jolma, A.; Campitelli, L.F.; Das, P.K.; Yin, Y.; Albu, M.; Chen, X.; Taipale, J.; Hughes, T.R.; Weirauch, M.T. The Human Transcription Factors. Cell 2018, 172, 650–665. [Google Scholar] [CrossRef] [Green Version]
  50. Umesono, K.; Evans, R.M. Determinants of target gene specificity for steroid/thyroid hormone receptors. Cell 1989, 57, 1139–1146. [Google Scholar] [CrossRef]
  51. Bourguet, W.; Ruff, M.; Chambon, P.; Gronemeyer, H.; Moras, D. Crystal structure of the ligand-binding domain of the human nuclear receptor RXR-alpha. Nature 1995, 375, 377–382. [Google Scholar] [CrossRef] [PubMed]
  52. Brzozowski, A.M.; Pike, A.C.; Dauter, Z.; Hubbard, R.E.; Bonn, T.; Engström, O.; Ohman, L.; Greene, G.L.; Gustafsson, J.A.; Carlquist, M. Molecular basis of agonism and antagonism in the oestrogen receptor. Nature 1997, 389, 753–758. [Google Scholar] [CrossRef] [PubMed]
  53. Darimont, B.D.; Wagner, R.L.; Apriletti, J.W.; Stallcup, M.R.; Kushner, P.J.; Baxter, J.D.; Fletterick, R.J.; Yamamoto, K.R. Structure and specificity of nuclear receptor-coactivator interactions. Genes Dev. 1998, 12, 3343–3356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Moras, D.; Gronemeyer, H. The nuclear receptor ligand-binding domain: Structure and function. Curr. Opin. Cell Biol. 1998, 10, 384–391. [Google Scholar] [CrossRef]
  55. Nolte, R.T.; Wisely, G.B.; Westin, S.; Cobb, J.E.; Lambert, M.H.; Kurokawa, R.; Rosenfeld, M.G.; Willson, T.M.; Glass, C.K.; Milburn, M.V. Ligand binding and co-activator assembly of the peroxisome proliferator-activated receptor-gamma. Nature 1998, 395, 137–143. [Google Scholar] [CrossRef]
  56. Renaud, J.P.; Rochel, N.; Ruff, M.; Vivat, V.; Chambon, P.; Gronemeyer, H.; Moras, D. Crystal structure of the RAR-gamma ligand-binding domain bound to all-trans retinoic acid. Nature 1995, 378, 681–689. [Google Scholar] [CrossRef]
  57. Shiau, A.K.; Barstad, D.; Loria, P.M.; Cheng, L.; Kushner, P.J.; Agard, D.A.; Greene, G.L. The structural basis of estrogen receptor/coactivator recognition and the antagonism of this interaction by tamoxifen. Cell 1998, 95, 927–937. [Google Scholar] [CrossRef] [Green Version]
  58. Wagner, R.L.; Apriletti, J.W.; McGrath, M.E.; West, B.L.; Baxter, J.D.; Fletterick, R.J. A structural role for hormone in the thyroid hormone receptor. Nature 1995, 378, 690–697. [Google Scholar] [CrossRef]
  59. Arnal, J.F.; Lenfant, F.; Metivier, R.; Flouriot, G.; Henrion, D.; Adlanmerini, M.; Fontaine, C.; Gourdy, P.; Chambon, P.; Katzenellenbogen, B.; et al. Membrane and Nuclear Estrogen Receptor Alpha Actions: From Tissue Specificity to Medical Implications. Physiol. Rev. 2017, 97, 1045–1087. [Google Scholar] [CrossRef]
  60. Perissi, V.; Rosenfeld, M.G. Controlling nuclear receptors: The circular logic of cofactor cycles. Nat. Rev. Mol. Cell Biol. 2005, 6, 542–554. [Google Scholar] [CrossRef]
  61. Mangelsdorf, D.J.; Evans, R.M. The RXR heterodimers and orphan receptors. Cell 1995, 83, 841–850. [Google Scholar] [CrossRef] [Green Version]
  62. Ptashne, M.; Gann, A.A. Activators and targets. Nature 1990, 346, 329–331. [Google Scholar] [CrossRef] [PubMed]
  63. Long, M.D.; Jacobi, J.J.; Singh, P.K.; Llimos, G.; Wani, S.A.; Rowsam, A.M.; Rosario, S.R.; Hoogstraat, M.; Linder, S.; Kirk, J.; et al. Reduced NCOR2 expression accelerates androgen deprivation therapy failure in prostate cancer. Cell Rep. 2021, 37, 110109. [Google Scholar] [CrossRef] [PubMed]
  64. Hu, X.; Lazar, M.A. The CoRNR motif controls the recruitment of corepressors by nuclear hormone receptors. Nature 1999, 402, 93–96. [Google Scholar] [CrossRef] [PubMed]
  65. Nagy, L.; Kao, H.Y.; Chakravarti, D.; Lin, R.J.; Hassig, C.A.; Ayer, D.E.; Schreiber, S.L.; Evans, R.M. Nuclear receptor repression mediated by a complex containing SMRT, mSin3A, and histone deacetylase. Cell 1997, 89, 373–380. [Google Scholar] [CrossRef] [Green Version]
  66. Perissi, V.; Aggarwal, A.; Glass, C.K.; Rose, D.W.; Rosenfeld, M.G. A corepressor/coactivator exchange complex required for transcriptional activation by nuclear receptors and other regulated transcription factors. Cell 2004, 116, 511–526. [Google Scholar] [CrossRef] [Green Version]
  67. Webb, P.; Anderson, C.M.; Valentine, C.; Nguyen, P.; Marimuthu, A.; West, B.L.; Baxter, J.D.; Kushner, P.J. The nuclear receptor corepressor (N-CoR) contains three isoleucine motifs (I/LXXII) that serve as receptor interaction domains (IDs). Mol. Endocrinol. 2000, 14, 1976–1985. [Google Scholar] [CrossRef]
  68. Soutourina, J. Transcription regulation by the Mediator complex. Nat. Rev. Mol. Cell Biol. 2018, 19, 262–274. [Google Scholar] [CrossRef]
  69. Su, Q.; Hu, S.; Gao, H.; Ma, R.; Yang, Q.; Pan, Z.; Wang, T.; Li, F. Role of AIB1 for tamoxifen resistance in estrogen receptor-positive breast cancer cells. Oncology 2008, 75, 159–168. [Google Scholar] [CrossRef]
  70. Ahmadian, M.; Suh, J.M.; Hah, N.; Liddle, C.; Atkins, A.R.; Downes, M.; Evans, R.M. PPARγ signaling and metabolism: The good, the bad and the future. Nat. Med. 2013, 19, 557–566. [Google Scholar] [CrossRef] [Green Version]
  71. Huang, W.; Glass, C.K. Nuclear receptors and inflammation control: Molecular mechanisms and pathophysiological relevance. Arterioscler. Thromb. Vasc. Biol. 2010, 30, 1542–1549. [Google Scholar] [CrossRef] [PubMed]
  72. Melekhova, A.; Baniahmad, A. ING Tumour Suppressors and ING Splice Variants as Coregulators of the Androgen Receptor Signalling in Prostate Cancer. Cells 2021, 10, 2599. [Google Scholar] [CrossRef] [PubMed]
  73. Mendoza-Parra, M.A.; Gronemeyer, H. Integrative genomics to dissect retinoid functions. Subcell Biochem. 2014, 70, 181–202. [Google Scholar] [PubMed]
  74. Moore, N.L.; Hickey, T.E.; Butler, L.M.; Tilley, W.D. Multiple nuclear receptor signaling pathways mediate the actions of synthetic progestins in target cells. Mol. Cell Endocrinol. 2012, 357, 60–70. [Google Scholar] [CrossRef]
  75. Panigrahi, A.; O’Malley, B.W. Mechanisms of enhancer action: The known and the unknown. Genome Biol. 2021, 22, 108. [Google Scholar] [CrossRef]
  76. Yang, P.B.; Hou, P.P.; Liu, F.Y.; Hong, W.B.; Chen, H.Z.; Sun, X.Y.; Li, P.; Zhang, Y.; Ju, C.Y.; Luo, L.J.; et al. Blocking PPARgamma interaction facilitates Nur77 interdiction of fatty acid uptake and suppresses breast cancer progression. Proc. Natl. Acad. Sci. USA 2020, 117, 27412–27422. [Google Scholar] [CrossRef]
  77. Jeong, K.W.; Lee, Y.H.; Stallcup, M.R. Recruitment of the SWI/SNF chromatin remodeling complex to steroid hormone-regulated promoters by nuclear receptor coactivator flightless-I. J. Biol. Chem. 2009, 284, 29298–29309. [Google Scholar] [CrossRef] [Green Version]
  78. Perez-Calero, C.; Bayona-Feliu, A.; Xue, X.; Barroso, S.I.; Munoz, S.; Gonzalez-Basallote, V.M.; Sung, P.; Aguilera, A. Corrigendum: UAP56/DDX39B is a major cotranscri.iptional RNA-DNA helicase that unwinds harmful R loops genome-wide. Genes Dev. 2021, 35, 573. [Google Scholar] [CrossRef]
  79. Zhou, X.; Khan, S.G.; Tamura, D.; Ueda, T.; Boyle, J.; Compe, E.; Egly, J.M.; DiGiovanna, J.J.; Kraemer, K.H. Abnormal XPD-induced nuclear receptor transactivation in DNA repair disorders: Trichothiodystrophy and xeroderma pigmentosum. Eur. J. Hum. Genet. 2013, 21, 831–837. [Google Scholar] [CrossRef] [Green Version]
  80. Lonard, D.M.; O’Malley, B.W. Nuclear receptor coregulators: Modulators of pathology and therapeutic targets. Nat. Rev. Endocrinol. 2012, 8, 598–604. [Google Scholar] [CrossRef]
  81. D’Santos, C.; Taylor, C.; Carroll, J.S.; Mohammed, H. RIME proteomics of estrogen and progesterone receptors in breast cancer. Data Brief 2015, 5, 276–280. [Google Scholar] [CrossRef] [PubMed]
  82. Foulds, C.E.; Feng, Q.; Ding, C.; Bailey, S.; Hunsaker, T.L.; Malovannaya, A.; Hamilton, R.A.; Gates, L.A.; Zhang, Z.; Li, C.; et al. Proteomic analysis of coregulators bound to ERα on DNA and nucleosomes reveals coregulator dynamics. Mol. Cell 2013, 51, 185–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Siddappa, M.H.; Wani, S.A.; Tang, H.; Gray, J.S.; Jafari, H.; Wu, H.; Long, M.D.; Elhussin, I.; Karanam, B.; Wang, H.; et al. Vitamin D Receptor Cistrome-Transcriptome Analyses Establishes Quantitatively Distinct Receptor Genomic Interactions in African American Prostate Cancer Regulated by BAZ1A. 2022. Available online: https://www.biorxiv.org/content/10.1101/2022.01.31.478573v1 (accessed on 28 April 2022).
  84. Papachristou, E.K.; Kishore, K.; Holding, A.N.; Harvey, K.; Roumeliotis, T.I.; Chilamakuri, C.S.R.; Omarjee, S.; Chia, K.M.; Swarbrick, A.; Lim, E.; et al. A quantitative mass spectrometry-based approach to monitor the dynamics of endogenous chromatin-associated protein complexes. Nat. Commun. 2018, 9, 2311. [Google Scholar] [CrossRef]
  85. Browne, A.L.; Charmsaz, S.; Vareslija, D.; Fagan, A.; Cosgrove, N.; Cocchiglia, S.; Purcell, S.; Ward, E.; Bane, F.; Hudson, L.; et al. Network analysis of SRC-1 reveals a novel transcription factor hub which regulates endocrine resistant breast cancer. Oncogene 2018, 37, 2008–2021. [Google Scholar] [CrossRef] [PubMed]
  86. Stashi, E.; York, B.; O’Malley, B.W. Steroid receptor coactivators: Servants and masters for control of systems metabolism. Trends Endocrinol. Metab. 2014, 25, 337–347. [Google Scholar] [CrossRef] [Green Version]
  87. Bagheri-Yarmand, R.; Talukder, A.H.; Wang, R.A.; Vadlamudi, R.K.; Kumar, R. Metastasis-associated protein 1 deregulation causes inappropriate mammary gland development and tumorigenesis. Development 2004, 131, 3469–3479. [Google Scholar] [CrossRef] [Green Version]
  88. O’Malley, B.W.; Kumar, R. Nuclear receptor coregulators in cancer biology. Cancer Res. 2009, 69, 8217–8222. [Google Scholar] [CrossRef] [Green Version]
  89. Dou, Y.; Milne, T.A.; Tackett, A.J.; Smith, E.R.; Fukuda, A.; Wysocka, J.; Allis, C.D.; Chait, B.T.; Hess, J.L.; Roeder, R.G. Physical association and coordinate function of the H3 K4 methyltransferase MLL1 and the H4 K16 acetyltransferase MOF. Cell 2005, 121, 873–885. [Google Scholar] [CrossRef] [Green Version]
  90. Lindroth, A.M.; Shultis, D.; Jasencakova, Z.; Fuchs, J.; Johnson, L.; Schubert, D.; Patnaik, D.; Pradhan, S.; Goodrich, J.; Schubert, I.; et al. Dual histone H3 methylation marks at lysines 9 and 27 required for interaction with CHROMOMETHYLASE3. EMBO J. 2004, 23, 4286–4296. [Google Scholar] [CrossRef] [Green Version]
  91. Yu, J.; Li, Y.; Ishizuka, T.; Guenther, M.G.; Lazar, M.A. A SANT motif in the SMRT corepressor interprets the histone code and promotes histone deacetylation. EMBO J. 2003, 22, 3403–3410. [Google Scholar] [CrossRef] [Green Version]
  92. Khan, O.Y.; Fu, G.; Ismail, A.; Srinivasan, S.; Cao, X.; Tu, Y.; Lu, S.; Nawaz, Z. Multifunction steroid receptor coactivator, E6-associated protein, is involved in development of the prostate gland. Mol. Endocrinol. 2006, 20, 544–559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Chen, D.; Ma, H.; Hong, H.; Koh, S.S.; Huang, S.M.; Schurter, B.T.; Aswad, D.W.; Stallcup, M.R. Regulation of transcription by a protein methyltransferase. Science 1999, 284, 2174–2177. [Google Scholar] [CrossRef] [PubMed]
  94. Poukka, H.; Karvonen, U.; Janne, O.A.; Palvimo, J.J. Covalent modification of the androgen receptor by small ubiquitin-like modifier 1 (SUMO-1). Proc. Natl. Acad. Sci. USA 2000, 97, 14145–14150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zheng, Z.; Cai, C.; Omwancha, J.; Chen, S.Y.; Baslan, T.; Shemshedini, L. SUMO-3 enhances androgen receptor transcriptional activity through a sumoylation-independent mechanism in prostate cancer cells. J. Biol. Chem. 2006, 281, 4002–4012. [Google Scholar] [CrossRef] [Green Version]
  96. Ishitani, K.; Yoshida, T.; Kitagawa, H.; Ohta, H.; Nozawa, S.; Kato, S. p54nrb acts as a transcriptional coactivator for activation function 1 of the human androgen receptor. Biochem. Biophys. Res. Commun. 2003, 306, 660–665. [Google Scholar] [CrossRef]
  97. Galbraith, L.C.A.; Mui, E.; Nixon, C.; Hedley, A.; Strachan, D.; MacKay, G.; Sumpton, D.; Sansom, O.J.; Leung, H.Y.; Ahmad, I. PPAR-gamma induced AKT3 expression increases levels of mitochondrial biogenesis driving prostate cancer. Oncogene 2021, 40, 2355–2366. [Google Scholar] [CrossRef]
  98. Liu, S.; Lin, S.J.; Li, G.; Kim, E.; Chen, Y.T.; Yang, D.R.; Tan, M.H.; Yong, E.L.; Chang, C. Differential roles of PPARgamma vs TR4 in prostate cancer and metabolic diseases. Endocr. Relat. Cancer 2014, 21, R279–R300. [Google Scholar] [CrossRef] [Green Version]
  99. Kaikkonen, S.; Paakinaho, V.; Sutinen, P.; Levonen, A.L.; Palvimo, J.J. Prostaglandin 15d-PGJ(2) inhibits androgen receptor signaling in prostate cancer cells. Mol. Endocrinol. 2013, 27, 212–223. [Google Scholar] [CrossRef] [Green Version]
  100. Govindarajan, R.; Ratnasinghe, L.; Simmons, D.L.; Siegel, E.R.; Midathada, M.V.; Kim, L.; Kim, P.J.; Owens, R.J.; Lang, N.P. Thiazolidinediones and the risk of lung, prostate, and colon cancer in patients with diabetes. J. Clin. Oncol 2007, 25, 1476–1481. [Google Scholar] [CrossRef] [Green Version]
  101. Butler, R.; Mitchell, S.H.; Tindall, D.J.; Young, C.Y. Nonapoptotic cell death associated with S-phase arrest of prostate cancer cells via the peroxisome proliferator-activated receptor gamma ligand, 15-deoxy-delta12,14-prostaglandin J2. Cell Growth Differ. 2000, 11, 49–61. [Google Scholar]
  102. Zheng, K.; Chen, S.; Hu, X. Peroxisome Proliferator Activated Receptor Gamma Coactivator-1 Alpha: A Double-Edged Sword in Prostate Cancer. Curr. Cancer Drug Targets 2022. [Google Scholar] [CrossRef] [PubMed]
  103. Torrano, V.; Valcarcel-Jimenez, L.; Cortazar, A.R.; Liu, X.; Urosevic, J.; Castillo-Martin, M.; Fernandez-Ruiz, S.; Morciano, G.; Caro-Maldonado, A.; Guiu, M.; et al. The metabolic co-regulator PGC1alpha suppresses prostate cancer metastasis. Nat. Cell Biol. 2016, 18, 645–656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Mayeur, G.L.; Kung, W.J.; Martinez, A.; Izumiya, C.; Chen, D.J.; Kung, H.J. Ku is a novel transcriptional recycling coactivator of the androgen receptor in prostate cancer cells. J. Biol. Chem. 2005, 280, 10827–10833. [Google Scholar] [CrossRef] [Green Version]
  105. Fan, C.Y.; Ren, H.Y.; Lee, P.; Caplan, A.J.; Cyr, D.M. The type I Hsp40 zinc finger-like region is required for Hsp70 to capture non-native polypeptides from Ydj1. J. Biol. Chem. 2005, 280, 695–702. [Google Scholar] [CrossRef] [Green Version]
  106. Takeshita, Y.; Fujinaga, R.; Zhao, C.; Yanai, A.; Shinoda, K. Huntingtin-associated protein 1 (HAP1) interacts with androgen receptor (AR) and suppresses SBMA-mutant-AR-induced apoptosis. Hum. Mol. Genet. 2006, 15, 2298–2312. [Google Scholar] [CrossRef] [PubMed]
  107. Lu, M.L.; Schneider, M.C.; Zheng, Y.; Zhang, X.; Richie, J.P. Caveolin-1 interacts with androgen receptor. A positive modulator of androgen receptor mediated transactivation. J. Biol. Chem. 2001, 276, 13442–13451. [Google Scholar] [CrossRef] [Green Version]
  108. Fujimoto, N.; Yeh, S.; Kang, H.Y.; Inui, S.; Chang, H.C.; Mizokami, A.; Chang, C. Cloning and characterization of androgen receptor coactivator, ARA55, in human prostate. J. Biol. Chem. 1999, 274, 8316–8321. [Google Scholar] [CrossRef] [Green Version]
  109. Wang, X.; Yang, Y.; Guo, X.; Sampson, E.R.; Hsu, C.L.; Tsai, M.Y.; Yeh, S.; Wu, G.; Guo, Y.; Chang, C. Suppression of androgen receptor transactivation by Pyk2 via interaction and phosphorylation of the ARA55 coregulator. J. Biol. Chem. 2002, 277, 15426–15431. [Google Scholar] [CrossRef] [Green Version]
  110. Kasai, M.; Guerrero-Santoro, J.; Friedman, R.; Leman, E.S.; Getzenberg, R.H.; DeFranco, D.B. The Group 3 LIM domain protein paxillin potentiates androgen receptor transactivation in prostate cancer cell lines. Cancer Res. 2003, 63, 4927–4935. [Google Scholar]
  111. Gangisetty, O.; Lauffart, B.; Sondarva, G.V.; Chelsea, D.M.; Still, I.H. The transforming acidic coiled coil proteins interact with nuclear histone acetyltransferases. Oncogene 2004, 23, 2559–2563. [Google Scholar] [CrossRef] [Green Version]
  112. Wilde, J.J.; Siegenthaler, J.A.; Dent, S.Y.; Niswander, L.A. Diencephalic Size Is Restricted by a Novel Interplay Between GCN5 Acetyltransferase Activity and Retinoic Acid Signaling. J. Neurosci 2017, 37, 2565–2579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Guyot, R.; Vincent, S.; Bertin, J.; Samarut, J.; Ravel-Chapuis, P. The transforming acidic coiled coil (TACC1) protein modulates the transcriptional activity of the nuclear receptors TR and RAR. BMC Mol. Biol 2010, 11, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Xie, Z.; Moy, L.Y.; Sanada, K.; Zhou, Y.; Buchman, J.J.; Tsai, L.H. Cep120 and TACCs control interkinetic nuclear migration and the neural progenitor pool. Neuron 2007, 56, 79–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Cully, M.; Shiu, J.; Piekorz, R.P.; Muller, W.J.; Done, S.J.; Mak, T.W. Transforming acidic coiled coil 1 promotes transformation and mammary tumorigenesis. Cancer Res. 2005, 65, 10363–10370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Still, I.H.; Hamilton, M.; Vince, P.; Wolfman, A.; Cowell, J.K. Cloning of TACC1, an embryonically expressed, potentially transforming coiled coil containing gene, from the 8p11 breast cancer amplicon. Oncogene 1999, 18, 4032–4038. [Google Scholar] [CrossRef] [Green Version]
  117. Fish, L.; Pencheva, N.; Goodarzi, H.; Tran, H.; Yoshida, M.; Tavazoie, S.F. Muscleblind-like 1 suppresses breast cancer metastatic colonization and stabilizes metastasis suppressor transcripts. Genes Dev. 2016, 30, 386–398. [Google Scholar] [CrossRef] [Green Version]
  118. Long, M.D.; Singh, P.K.; Russell, J.R.; Llimos, G.; Rosario, S.; Rizvi, A.; van den Berg, P.R.; Kirk, J.; Sucheston-Campbell, L.E.; Smiraglia, D.J.; et al. The miR-96 and RARgamma signaling axis governs androgen signaling and prostate cancer progression. Oncogene 2019, 38, 421–444. [Google Scholar] [CrossRef] [Green Version]
  119. Hodges, C.; Kirkland, J.G.; Crabtree, G.R. The Many Roles of BAF (mSWI/SNF) and PBAF Complexes in Cancer. Cold Spring Harb. Perspect. Med. 2016, 6, a026930. [Google Scholar] [CrossRef] [Green Version]
  120. Jelinic, P.; Mueller, J.J.; Olvera, N.; Dao, F.; Scott, S.N.; Shah, R.; Gao, J.; Schultz, N.; Gonen, M.; Soslow, R.A.; et al. Recurrent SMARCA4 mutations in small cell carcinoma of the ovary. Nat. Genet. 2014, 46, 424–426. [Google Scholar] [CrossRef]
  121. Ramos, P.; Karnezis, A.N.; Craig, D.W.; Sekulic, A.; Russell, M.L.; Hendricks, W.P.; Corneveaux, J.J.; Barrett, M.T.; Shumansky, K.; Yang, Y.; et al. Small cell carcinoma of the ovary, hypercalcemic type, displays frequent inactivating germline and somatic mutations in SMARCA4. Nat. Genet. 2014, 46, 427–429. [Google Scholar] [CrossRef] [Green Version]
  122. Cheng, A.S.; Jin, V.X.; Fan, M.; Smith, L.T.; Liyanarachchi, S.; Yan, P.S.; Leu, Y.W.; Chan, M.W.; Plass, C.; Nephew, K.P.; et al. Combinatorial analysis of transcription factor partners reveals recruitment of c-MYC to estrogen receptor-alpha responsive promoters. Mol. Cell 2006, 21, 393–404. [Google Scholar] [CrossRef] [PubMed]
  123. Li, H.; Liu, J. The novel function of HINFP as a co-activator in sterol-regulated transcription of PCSK9 in HepG2 cells. Biochem. J. 2012, 443, 757–768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Unno, A.; Takada, I.; Takezawa, S.; Oishi, H.; Baba, A.; Shimizu, T.; Tokita, A.; Yanagisawa, J.; Kato, S. TRRAP as a hepatic coactivator of LXR and FXR function. Biochem. Biophys. Res. Commun. 2005, 327, 933–938. [Google Scholar] [CrossRef] [PubMed]
  125. Majumder, S.; Liu, Y.; Ford, O.H., 3rd; Mohler, J.L.; Whang, Y.E. Involvement of arginine methyltransferase CARM1 in androgen receptor function and prostate cancer cell viability. Prostate 2006, 66, 1292–1301. [Google Scholar] [CrossRef] [PubMed]
  126. Peng, B.L.; Li, W.J.; Ding, J.C.; He, Y.H.; Ran, T.; Xie, B.L.; Wang, Z.R.; Shen, H.F.; Xiao, R.Q.; Gao, W.W.; et al. A hypermethylation strategy utilized by enhancer-bound CARM1 to promote estrogen receptor α-dependent transcriptional activation and breast carcinogenesis. Theranostics 2020, 10, 3451–3473. [Google Scholar] [CrossRef]
  127. Purcell, D.J.; Chauhan, S.; Jimenez-Stinson, D.; Elliott, K.R.; Tsewang, T.D.; Lee, Y.H.; Marples, B.; Lee, D.Y. Novel CARM1-Interacting Protein, DZIP3, Is a Transcriptional Coactivator of Estrogen Receptor-α. Mol. Endocrinol. 2015, 29, 1708–1719. [Google Scholar] [CrossRef] [Green Version]
  128. Yadav, N.; Cheng, D.; Richard, S.; Morel, M.; Iyer, V.R.; Aldaz, C.M.; Bedford, M.T. CARM1 promotes adipocyte differentiation by coactivating PPARgamma. EMBO Rep. 2008, 9, 193–198. [Google Scholar] [CrossRef]
  129. Giraud, S.; Bienvenu, F.; Avril, S.; Gascan, H.; Heery, D.M.; Coqueret, O. Functional interaction of STAT3 transcription factor with the coactivator NcoA/SRC1a. J. Biol. Chem. 2002, 277, 8004–8011. [Google Scholar] [CrossRef] [Green Version]
  130. Katoh, M. Functional proteomics of the epigenetic regulators ASXL1, ASXL2 and ASXL3: A convergence of proteomics and epigenetics for translational medicine. Expert Rev. Proteom. 2015, 12, 317–328. [Google Scholar] [CrossRef]
  131. Wang, N.; Zou, Q.; Xu, J.; Zhang, J.; Liu, J. Ligand binding and heterodimerization with retinoid X receptor α (RXRα) induce farnesoid X receptor (FXR) conformational changes affecting coactivator binding. J. Biol. Chem. 2018, 293, 18180–18191. [Google Scholar] [CrossRef] [Green Version]
  132. Huang, D.; Yang, C.; Wang, Y.; Liao, Y.; Huang, K. PARP-1 suppresses adiponectin expression through poly(ADP-ribosyl)ation of PPAR gamma in cardiac fibroblasts. Cardiovasc. Res. 2009, 81, 98–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Jeyakumar, M.; Liu, X.F.; Erdjument-Bromage, H.; Tempst, P.; Bagchi, M.K. Phosphorylation of thyroid hormone receptor-associated nuclear receptor corepressor holocomplex by the DNA-dependent protein kinase enhances its histone deacetylase activity. J. Biol. Chem. 2007, 282, 9312–9322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Pavri, R.; Lewis, B.; Kim, T.K.; Dilworth, F.J.; Erdjument-Bromage, H.; Tempst, P.; de Murcia, G.; Evans, R.; Chambon, P.; Reinberg, D. PARP-1 determines specificity in a retinoid signaling pathway via direct modulation of mediator. Mol. Cell 2005, 18, 83–96. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, F.; Wang, Y.; Wang, L.; Luo, X.; Huang, K.; Wang, C.; Du, M.; Liu, F.; Luo, T.; Huang, D.; et al. Poly(ADP-ribose) polymerase 1 is a key regulator of estrogen receptor α-dependent gene transcription. J. Biol. Chem. 2013, 288, 11348–11357. [Google Scholar] [CrossRef] [Green Version]
  136. Scarpulla, R.C. Metabolic control of mitochondrial biogenesis through the PGC-1 family regulatory network. Biochim. Biophys. Acta 2011, 1813, 1269–1278. [Google Scholar] [CrossRef] [Green Version]
  137. Tan, H.W.S.; Anjum, B.; Shen, H.M.; Ghosh, S.; Yen, P.M.; Sinha, R.A. Lysosomal inhibition attenuates peroxisomal gene transcription via suppression of PPARA and PPARGC1A levels. Autophagy 2019, 15, 1455–1459. [Google Scholar] [CrossRef]
  138. Clapier, C.R.; Cairns, B.R. The biology of chromatin remodeling complexes. Annu. Rev. Biochem. 2009, 78, 273–304. [Google Scholar] [CrossRef]
  139. He, S.; Wu, Z.; Tian, Y.; Yu, Z.; Yu, J.; Wang, X.; Li, J.; Liu, B.; Xu, Y. Structure of nucleosome-bound human BAF complex. Science 2020, 367, 875–881. [Google Scholar] [CrossRef]
  140. Seth-Vollenweider, T.; Joshi, S.; Dhawan, P.; Sif, S.; Christakos, S. Novel mechanism of negative regulation of 1,25-dihydroxyvitamin D3-induced 25-hydroxyvitamin D3 24-hydroxylase (Cyp24a1) Transcription: Epigenetic modification involving cross-talk between protein-arginine methyltransferase 5 and the SWI/SNF complex. J. Biol. Chem. 2014, 289, 33958–33970. [Google Scholar] [CrossRef] [Green Version]
  141. Roberts, C.W.; Orkin, S.H. The SWI/SNF complex—Chromatin and cancer. Nat. Rev. Cancer 2004, 4, 133–142. [Google Scholar] [CrossRef]
  142. Wu, Q.; Lian, J.B.; Stein, J.L.; Stein, G.S.; Nickerson, J.A.; Imbalzano, A.N. The BRG1 ATPase of human SWI/SNF chromatin remodeling enzymes as a driver of cancer. Epigenomics 2017, 9, 919–931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Kim, M.; Lu, F.; Zhang, Y. Loss of HDAC-Mediated Repression and Gain of NF-kappaB Activation Underlie Cytokine Induction in ARID1A- and PIK3CA-Mutation-Driven Ovarian Cancer. Cell Rep. 2016, 17, 275–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Allen, M.D.; Bycroft, M.; Zinzalla, G. Structure of the BRK domain of the SWI/SNF chromatin remodeling complex subunit BRG1 reveals a potential role in protein-protein interactions. Protein Sci. 2020, 29, 1047–1053. [Google Scholar] [CrossRef] [PubMed]
  145. Lu, C.; Allis, C.D. SWI/SNF complex in cancer. Nat. Genet. 2017, 49, 178–179. [Google Scholar] [CrossRef] [PubMed]
  146. Hohmann, A.F.; Vakoc, C.R. A rationale to target the SWI/SNF complex for cancer therapy. Trends Genet. 2014, 30, 356–363. [Google Scholar] [CrossRef] [Green Version]
  147. Medina, P.P.; Sanchez-Cespedes, M. Involvement of the chromatin-remodeling factor BRG1/SMARCA4 in human cancer. Epigenetics 2008, 3, 64–68. [Google Scholar] [CrossRef] [Green Version]
  148. Trotter, K.W.; Archer, T.K. The BRG1 transcriptional coregulator. Nucl. Recept. Signal. 2008, 6, e004. [Google Scholar] [CrossRef] [Green Version]
  149. Huang, C.Y.; Beliakoff, J.; Li, X.; Lee, J.; Li, X.; Sharma, M.; Lim, B.; Sun, Z. hZimp7, a novel PIAS-like protein, enhances androgen receptor-mediated transcription and interacts with SWI/SNF-like BAF complexes. Mol. Endocrinol. 2005, 19, 2915–2929. [Google Scholar] [CrossRef] [Green Version]
  150. Jin, M.L.; Kim, Y.W.; Jeong, K.W. BAF53A regulates androgen receptor-mediated gene expression and proliferation in LNCaP cells. Biochem. Biophys. Res. Commun. 2018, 505, 618–623. [Google Scholar] [CrossRef]
  151. Tropee, R.; de la Pena Avalos, B.; Gough, M.; Snell, C.; Duijf, P.H.G.; Dray, E. The SWI/SNF subunit SMARCD3 regulates cell cycle progression and predicts survival outcome in ER+ breast cancer. Breast Cancer Res. Treat. 2021, 185, 601–614. [Google Scholar] [CrossRef]
  152. Yates, L.R.; Knappskog, S.; Wedge, D.; Farmery, J.H.R.; Gonzalez, S.; Martincorena, I.; Alexandrov, L.B.; Van Loo, P.; Haugland, H.K.; Lilleng, P.K.; et al. Genomic Evolution of Breast Cancer Metastasis and Relapse. Cancer Cell 2017, 32, 169–184.e167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Trotter, K.W.; King, H.A.; Archer, T.K. Glucocorticoid Receptor Transcriptional Activation via the BRG1-Dependent Recruitment of TOP2beta and Ku70/86. Mol. Cell Biol. 2015, 35, 2799–2817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Hsia, E.Y.; Goodson, M.L.; Zou, J.X.; Privalsky, M.L.; Chen, H.W. Nuclear receptor coregulators as a new paradigm for therapeutic targeting. Adv. Drug Deliv. Rev. 2010, 62, 1227–1237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Onate, 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]
  156. Carapeti, M.; Aguiar, R.C.; Chase, A.; Goldman, J.M.; Cross, N.C. Assignment of the steroid receptor coactivator-1 (SRC-1) gene to human chromosome band 2p23. Genomics 1998, 52, 242–244. [Google Scholar] [CrossRef] [PubMed]
  157. 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]
  158. Guan, X.Y.; Xu, J.; Anzick, S.L.; Zhang, H.; Trent, J.M.; Meltzer, P.S. Hybrid selection of transcribed sequences from microdissected DNA: Isolation of genes within amplified region at 20q11-q13.2 in breast cancer. Cancer Res. 1996, 56, 3446–3450. [Google Scholar]
  159. Dasgupta, S.; O’Malley, B.W. Transcriptional coregulators: Emerging roles of SRC family of coactivators in disease pathology. J. Mol. Endocrinol. 2014, 53, R47–R59. [Google Scholar] [CrossRef] [Green Version]
  160. York, B.; Sagen, J.V.; Tsimelzon, A.; Louet, J.F.; Chopra, A.R.; Reineke, E.L.; Zhou, S.; Stevens, R.D.; Wenner, B.R.; Ilkayeva, O.; et al. Research resource: Tissue- and pathway-specific metabolomic profiles of the steroid receptor coactivator (SRC) family. Mol. Endocrinol. 2013, 27, 366–380. [Google Scholar] [CrossRef] [Green Version]
  161. Hsia, E.Y.; Zou, J.X.; Chen, H.W. The roles and action mechanisms of p160/SRC coactivators and the ANCCA coregulator in cancer. Prog Mol. Biol. Transl. Sci. 2009, 87, 261–298. [Google Scholar]
  162. Sharif, G.M.; Campbell, M.J.; Nasir, A.; Sengupta, S.; Graham, G.T.; Kushner, M.H.; Kietzman, W.B.; Schmidt, M.O.; Pearson, G.W.; Loudig, O.; et al. An AIB1 Isoform Alters Enhancer Access and Enables Progression of Early-Stage Triple-Negative Breast Cancer. Cancer Res. 2021, 81, 4230–4241. [Google Scholar] [CrossRef] [PubMed]
  163. Agoulnik, I.U.; Vaid, A.; Nakka, M.; Alvarado, M.; Bingman, W.E., 3rd; Erdem, H.; Frolov, A.; Smith, C.L.; Ayala, G.E.; Ittmann, M.M.; et al. Androgens modulate expression of transcription intermediary factor 2, an androgen receptor coactivator whose expression level correlates with early biochemical recurrence in prostate cancer. Cancer Res. 2006, 66, 10594–10602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Al-azawi, D.; Ilroy, M.M.; Kelly, G.; Redmond, A.M.; Bane, F.T.; Cocchiglia, S.; Hill, A.D.; Young, L.S. Ets-2 and p160 proteins collaborate to regulate c-Myc in endocrine resistant breast cancer. Oncogene 2008, 27, 3021–3031. [Google Scholar] [CrossRef] [Green Version]
  165. Louie, M.C.; Zou, J.X.; Rabinovich, A.; Chen, H.W. ACTR/AIB1 functions as an E2F1 coactivator to promote breast cancer cell proliferation and antiestrogen resistance. Mol. Cell Biol. 2004, 24, 5157–5171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Zhou, H.J.; Yan, J.; Luo, W.; Ayala, G.; Lin, S.H.; Erdem, H.; Ittmann, M.; Tsai, S.Y.; Tsai, M.J. SRC-3 is required for prostate cancer cell proliferation and survival. Cancer Res. 2005, 65, 7976–7983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Zou, J.X.; Zhong, Z.; Shi, X.B.; Tepper, C.G.; de Vere White, R.W.; Kung, H.J.; Chen, H. ACTR/AIB1/SRC-3 and androgen receptor control prostate cancer cell proliferation and tumor growth through direct control of cell cycle genes. Prostate 2006, 66, 1474–1486. [Google Scholar] [CrossRef]
  168. Lonard, D.M.; O’Malley, B.W. Nuclear receptor coregulators: Judges, juries, and executioners of cellular regulation. Mol. Cell 2007, 27, 691–700. [Google Scholar] [CrossRef]
  169. Fleming, F.J.; Myers, E.; Kelly, G.; Crotty, T.B.; McDermott, E.W.; O’Higgins, N.J.; Hill, A.D.; Young, L.S. Expression of SRC-1, AIB1, and PEA3 in HER2 mediated endocrine resistant breast cancer; a predictive role for SRC-1. J. Clin. Pathol. 2004, 57, 1069–1074. [Google Scholar] [CrossRef] [Green Version]
  170. Gnanapragasam, V.J.; Leung, H.Y.; Pulimood, A.S.; Neal, D.E.; Robson, C.N. Expression of RAC 3, a steroid hormone receptor co-activator in prostate cancer. Br. J. Cancer 2001, 85, 1928–1936. [Google Scholar] [CrossRef]
  171. Gregory, C.W.; He, B.; Johnson, R.T.; Ford, O.H.; Mohler, J.L.; French, F.S.; Wilson, E.M. A mechanism for androgen receptor-mediated prostate cancer recurrence after androgen deprivation therapy. Cancer Res. 2001, 61, 4315–4319. [Google Scholar]
  172. Harigopal, M.; Heymann, J.; Ghosh, S.; Anagnostou, V.; Camp, R.L.; Rimm, D.L. Estrogen receptor co-activator (AIB1) protein expression by automated quantitative analysis (AQUA) in a breast cancer tissue microarray and association with patient outcome. Breast Cancer Res. Treat. 2009, 115, 77–85. [Google Scholar] [CrossRef] [PubMed]
  173. Osborne, C.K.; Bardou, V.; Hopp, T.A.; Chamness, G.C.; Hilsenbeck, S.G.; Fuqua, S.A.; Wong, J.; Allred, D.C.; Clark, G.M.; Schiff, R. Role of the estrogen receptor coactivator AIB1 (SRC-3) and HER-2/neu in tamoxifen resistance in breast cancer. J. Natl. Cancer Inst. 2003, 95, 353–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Yan, J.; Erdem, H.; Li, R.; Cai, Y.; Ayala, G.; Ittmann, M.; Yu-Lee, L.Y.; Tsai, S.Y.; Tsai, M.J. Steroid receptor coactivator-3/AIB1 promotes cell migration and invasiveness through focal adhesion turnover and matrix metalloproteinase expression. Cancer Res. 2008, 68, 5460–5468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Alkner, S.; Bendahl, P.O.; Grabau, D.; Lövgren, K.; Stål, O.; Rydén, L.; Fernö, M. AIB1 is a predictive factor for tamoxifen response in premenopausal women. Ann. Oncol. 2010, 21, 238–244. [Google Scholar] [CrossRef]
  176. Iwase, H.; Omoto, Y.; Toyama, T.; Yamashita, H.; Hara, Y.; Sugiura, H.; Zhang, Z. Clinical significance of AIB1 expression in human breast cancer. Breast Cancer Res. Treat. 2003, 80, 339–345. [Google Scholar] [CrossRef]
  177. Jiang, L.; Ren, L.; Chen, H.; Pan, J.; Zhang, Z.; Kuang, X.; Chen, X.; Bao, W.; Lin, C.; Zhou, Z.; et al. NCAPG confers trastuzumab resistance via activating SRC/STAT3 signaling pathway in HER2-positive breast cancer. Cell Death Dis. 2020, 11, 547. [Google Scholar] [CrossRef]
  178. Shou, J.; Massarweh, S.; Osborne, C.K.; Wakeling, A.E.; Ali, S.; Weiss, H.; Schiff, R. Mechanisms of tamoxifen resistance: Increased estrogen receptor-HER2/neu cross-talk in ER/HER2-positive breast cancer. J. Natl. Cancer Inst. 2004, 96, 926–935. [Google Scholar] [CrossRef] [Green Version]
  179. Zhao, W.; Zhang, Q.; Kang, X.; Jin, S.; Lou, C. AIB1 is required for the acquisition of epithelial growth factor receptor-mediated tamoxifen resistance in breast cancer cells. Biochem. Biophys. Res. Commun. 2009, 380, 699–704. [Google Scholar] [CrossRef]
  180. Agoulnik, I.U.; Vaid, A.; Bingman, W.E., 3rd; Erdeme, H.; Frolov, A.; Smith, C.L.; Ayala, G.; Ittmann, M.M.; Weigel, N.L. Role of SRC-1 in the promotion of prostate cancer cell growth and tumor progression. Cancer Res. 2005, 65, 7959–7967. [Google Scholar] [CrossRef] [Green Version]
  181. Qin, L.; Chen, J.; Lu, D.; Jain, P.; Yu, Y.; Cardenas, D.; Peng, X.; Yu, X.; Xu, J.; Wang, J.; et al. Development of improved SRC-3 inhibitors as breast cancer therapeutic agents. Endocr. Relat. Cancer 2021, 28, 657–670. [Google Scholar] [CrossRef]
  182. Cheng, S.; Douglas-Jones, A.; Yang, X.; Mansel, R.E.; Jiang, W.G. Transforming acidic coiled-coil-containing protein 2 (TACC2) in human breast cancer, expression pattern and clinical/prognostic relevance. Cancer Genom. Proteom. 2010, 7, 67–73. [Google Scholar]
  183. Piekorz, R.P.; Hoffmeyer, A.; Duntsch, C.D.; McKay, C.; Nakajima, H.; Sexl, V.; Snyder, L.; Rehg, J.; Ihle, J.N. The centrosomal protein TACC3 is essential for hematopoietic stem cell function and genetically interfaces with p53-regulated apoptosis. EMBO J. 2002, 21, 653–664. [Google Scholar] [CrossRef] [Green Version]
  184. Conte, N.; Charafe-Jauffret, E.; Delaval, B.; Adélaïde, J.; Ginestier, C.; Geneix, J.; Isnardon, D.; Jacquemier, J.; Birnbaum, D. Carcinogenesis and translational controls: TACC1 is down-regulated in human cancers and associates with mRNA regulators. Oncogene 2002, 21, 5619–5630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Conte, N.; Delaval, B.; Ginestier, C.; Ferrand, A.; Isnardon, D.; Larroque, C.; Prigent, C.; Séraphin, B.; Jacquemier, J.; Birnbaum, D. TACC1-chTOG-Aurora A protein complex in breast cancer. Oncogene 2003, 22, 8102–8116. [Google Scholar] [CrossRef] [Green Version]
  186. Delaval, B.; Ferrand, A.; Conte, N.; Larroque, C.; Hernandez-Verdun, D.; Prigent, C.; Birnbaum, D. Aurora B -TACC1 protein complex in cytokinesis. Oncogene 2004, 23, 4516–4522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Gabillard, J.C.; Ulisse, S.; Baldini, E.; Sorrenti, S.; Cremet, J.Y.; Coccaro, C.; Prigent, C.; D’Armiento, M.; Arlot-Bonnemains, Y. Aurora-C interacts with and phosphorylates the transforming acidic coiled-coil 1 protein. Biochem. Biophys. Res. Commun. 2011, 408, 647–653. [Google Scholar] [CrossRef] [PubMed]
  188. Ali, J.; Sabiha, B.; Jan, H.U.; Haider, S.A.; Khan, A.A.; Ali, S.S. Genetic etiology of oral cancer. Oral Oncol 2017, 70, 23–28. [Google Scholar] [CrossRef]
  189. Ding, A.; Zhao, W.; Shi, X.; Yao, R.; Zhou, F.; Yue, L.; Liu, S.; Qiu, W. Impact of NPM, TFF3 and TACC1 on the prognosis of patients with primary gastric cancer. PLoS ONE 2013, 8, e82136. [Google Scholar] [CrossRef]
  190. Xiang, L.; Li, J.; Jiang, W.; Shen, X.; Yang, W.; Wu, X.; Yang, H. Comprehensive analysis of targetable oncogenic mutations in chinese cervical cancers. Oncotarget 2015, 6, 4968–4975. [Google Scholar] [CrossRef] [Green Version]
  191. Petrie, K.; Urban-Wojciuk, Z.; Sbirkov, Y.; Graham, A.; Hamann, A.; Brown, G. Retinoic acid receptor gamma is a therapeutically targetable driver of growth and survival in prostate cancer. Cancer Rep. 2020, 3, e1284. [Google Scholar]
  192. Cai, K.; Gudas, L.J. Retinoic acid receptors and GATA transcription factors activate the transcription of the human lecithin:retinol acyltransferase gene. Int. J. Biochem. Cell Biol. 2009, 41, 546–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Puigserver, P.; Wu, Z.; Park, C.W.; Graves, R.; Wright, M.; Spiegelman, B.M. A cold-inducible coactivator of nuclear receptors linked to adaptive thermogenesis. Cell 1998, 92, 829–839. [Google Scholar] [CrossRef] [Green Version]
  194. Knutti, D.; Kaul, A.; Kralli, A. A tissue-specific coactivator of steroid receptors, identified in a functional genetic screen. Mol. Cell Biol. 2000, 20, 2411–2422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Lerin, C.; Rodgers, J.T.; Kalume, D.E.; Kim, S.H.; Pandey, A.; Puigserver, P. GCN5 acetyltransferase complex controls glucose metabolism through transcriptional repression of PGC-1alpha. Cell Metab. 2006, 3, 429–438. [Google Scholar] [CrossRef] [Green Version]
  196. Rodgers, J.T.; Lerin, C.; Haas, W.; Gygi, S.P.; Spiegelman, B.M.; Puigserver, P. Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature 2005, 434, 113–118. [Google Scholar] [CrossRef]
  197. Lawson, D.E.; Charman, M.; Wilson, P.W.; Edelstein, S. Some characteristics of new tissue-binding proteins for metabolites of vitamin D other than 1,25-dihydroxyvitamin D. Biochim. Biophys. Acta 1976, 437, 403–415. [Google Scholar] [CrossRef]
  198. Shiota, M.; Yokomizo, A.; Tada, Y.; Inokuchi, J.; Tatsugami, K.; Kuroiwa, K.; Uchiumi, T.; Fujimoto, N.; Seki, N.; Naito, S. Peroxisome proliferator-activated receptor gamma coactivator-1alpha interacts with the androgen receptor (AR) and promotes prostate cancer cell growth by activating the AR. Mol. Endocrinol. 2010, 24, 114–127. [Google Scholar] [CrossRef]
  199. LaGory, E.L.; Wu, C.; Taniguchi, C.M.; Ding, C.C.; Chi, J.T.; von Eyben, R.; Scott, D.A.; Richardson, A.D.; Giaccia, A.J. Suppression of PGC-1alpha Is Critical for Reprogramming Oxidative Metabolism in Renal Cell Carcinoma. Cell Rep. 2015, 12, 116–127. [Google Scholar] [CrossRef] [Green Version]
  200. Zu, Y.; Chen, X.F.; Li, Q.; Zhang, S.T.; Si, L.N. PGC-1α activates SIRT3 to modulate cell proliferation and glycolytic metabolism in breast cancer. Neoplasma 2021, 68, 352–361. [Google Scholar] [CrossRef]
  201. 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] [Green Version]
  202. Grasso, C.S.; Wu, Y.M.; Robinson, D.R.; Cao, X.; Dhanasekaran, S.M.; Khan, A.P.; Quist, M.J.; Jing, X.; Lonigro, R.J.; Brenner, J.C.; et al. The mutational landscape of lethal castration-resistant prostate cancer. Nature 2012, 487, 239–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. 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, 161, 1215–1228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Tiwari, R.; Manzar, N.; Bhatia, V.; Yadav, A.; Nengroo, M.A.; Datta, D.; Carskadon, S.; Gupta, N.; Sigouros, M.; Khani, F.; et al. Androgen deprivation upregulates SPINK1 expression and potentiates cellular plasticity in prostate cancer. Nat. Commun. 2020, 11, 384. [Google Scholar] [CrossRef] [Green Version]
  205. Lee, N.S.; Evgrafov, O.V.; Souaiaia, T.; Bonyad, A.; Herstein, J.; Lee, J.Y.; Kim, J.; Ning, Y.; Sixto, M.; Weitz, A.C.; et al. Non-coding RNAs derived from an alternatively spliced REST transcript (REST-003) regulate breast cancer invasiveness. Sci. Rep. 2015, 5, 11207. [Google Scholar] [CrossRef] [Green Version]
  206. Svensson, C.; Ceder, J.; Iglesias-Gato, D.; Chuan, Y.C.; Pang, S.T.; Bjartell, A.; Martinez, R.M.; Bott, L.; Helczynski, L.; Ulmert, D.; et al. REST mediates androgen receptor actions on gene repression and predicts early recurrence of prostate cancer. Nucleic Acids Res. 2014, 42, 999–1015. [Google Scholar] [CrossRef] [PubMed]
  207. Mulligan, P.; Westbrook, T.F.; Ottinger, M.; Pavlova, N.; Chang, B.; Macia, E.; Shi, Y.J.; Barretina, J.; Liu, J.; Howley, P.M.; et al. CDYL bridges REST and histone methyltransferases for gene repression and suppression of cellular transformation. Mol. Cell 2008, 32, 718–726. [Google Scholar] [CrossRef]
  208. Lempiainen, J.K.; Niskanen, E.A.; Vuoti, K.M.; Lampinen, R.E.; Goos, H.; Varjosalo, M.; Palvimo, J.J. Agonist-specific Protein Interactomes of Glucocorticoid and Androgen Receptor as Revealed by Proximity Mapping. Mol. Cell Proteom. 2017, 16, 1462–1474. [Google Scholar] [CrossRef] [Green Version]
  209. Tong, D.; Zhang, J.; Wang, X.; Li, Q.; Liu, L.Y.; Yang, J.; Guo, B.; Ni, L.; Zhao, L.; Huang, C. MeCP2 facilitates breast cancer growth via promoting ubiquitination-mediated P53 degradation by inhibiting RPL5/RPL11 transcription. Oncogenesis 2020, 9, 56. [Google Scholar] [CrossRef]
  210. Luo, D.; Ge, W. MeCP2 Promotes Colorectal Cancer Metastasis by Modulating ZEB1 Transcription. Cancers 2020, 12, 758. [Google Scholar] [CrossRef] [Green Version]
  211. Bian, E.; Chen, X.; Xu, Y.; Ji, X.; Cheng, M.; Wang, H.; Fang, Z.; Zhao, B. A central role for MeCP2 in the epigenetic repression of miR-200c during epithelial-to-mesenchymal transition of glioma. J. Exp. Clin. Cancer Res. 2019, 38, 366. [Google Scholar] [CrossRef]
  212. Yang, C.; Wu, J.; Liu, X.; Wang, Y.; Liu, B.; Chen, X.; Wu, X.; Yan, D.; Han, L.; Liu, S.; et al. Circadian Rhythm Is Disrupted by ZNF704 in Breast Carcinogenesis. Cancer Res. 2020, 80, 4114–4128. [Google Scholar] [CrossRef] [PubMed]
  213. Yang, Y.; Huang, W.; Qiu, R.; Liu, R.; Zeng, Y.; Gao, J.; Zheng, Y.; Hou, Y.; Wang, S.; Yu, W.; et al. LSD1 coordinates with the SIN3A/HDAC complex and maintains sensitivity to chemotherapy in breast cancer. J. Mol. Cell Biol. 2018, 10, 285–301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Li, W.; Zhang, Z.; Liu, X.; Cheng, X.; Zhang, Y.; Han, X.; Zhang, Y.; Liu, S.; Yang, J.; Xu, B.; et al. The FOXN3-NEAT1-SIN3A repressor complex promotes progression of hormonally responsive breast cancer. J. Clin. Investig. 2017, 127, 3421–3440. [Google Scholar] [CrossRef] [PubMed]
  215. Neupane, M.; Clark, A.P.; Landini, S.; Birkbak, N.J.; Eklund, A.C.; Lim, E.; Culhane, A.C.; Barry, W.T.; Schumacher, S.E.; Beroukhim, R.; et al. MECP2 Is a Frequently Amplified Oncogene with a Novel Epigenetic Mechanism That Mimics the Role of Activated RAS in Malignancy. Cancer Discov. 2016, 6, 45–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Sharma, M.; Sun, Z. 5’TG3’ interacting factor interacts with Sin3A and represses AR-mediated transcription. Mol. Endocrinol. 2001, 15, 1918–1928. [Google Scholar] [PubMed]
  217. Kemper, J.K.; Kim, H.; Miao, J.; Bhalla, S.; Bae, Y. Role of an mSin3A-Swi/Snf chromatin remodeling complex in the feedback repression of bile acid biosynthesis by SHP. Mol. Cell Biol. 2004, 24, 7707–7719. [Google Scholar] [CrossRef] [Green Version]
  218. Sharma, D.; Saxena, N.K.; Davidson, N.E.; Vertino, P.M. Restoration of tamoxifen sensitivity in estrogen receptor-negative breast cancer cells: Tamoxifen-bound reactivated ER recruits distinctive corepressor complexes. Cancer Res. 2006, 66, 6370–6378. [Google Scholar] [CrossRef] [Green Version]
  219. Lee, M.O.; Kang, H.J. Role of coactivators and corepressors in the induction of the RARbeta gene in human colon cancer cells. Biol. Pharm. Bull. 2002, 25, 1298–1302. [Google Scholar] [CrossRef] [Green Version]
  220. Asim, M.; Hafeez, B.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]
  221. White, J.H.; Fernandes, I.; Mader, S.; Yang, X.J. Corepressor recruitment by agonist-bound nuclear receptors. Vitam Horm. 2004, 68, 123–143. [Google Scholar]
  222. Klotz-Noack, K.; Klinger, B.; Rivera, M.; Bublitz, N.; Uhlitz, F.; Riemer, P.; Luthen, M.; Sell, T.; Kasack, K.; Gastl, B.; et al. SFPQ Depletion Is Synthetically Lethal with BRAF(V600E) in Colorectal Cancer Cells. Cell Rep. 2020, 32, 108184. [Google Scholar] [CrossRef] [PubMed]
  223. Mitobe, Y.; Iino, K.; Takayama, K.I.; Ikeda, K.; Suzuki, T.; Aogi, K.; Kawabata, H.; Suzuki, Y.; Horie-Inoue, K.; Inoue, S. PSF Promotes ER-Positive Breast Cancer Progression via Posttranscriptional Regulation of ESR1 and SCFD2. Cancer Res. 2020, 80, 2230–2242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. De Silva, H.C.; Lin, M.Z.; Phillips, L.; Martin, J.L.; Baxter, R.C. IGFBP-3 interacts with NONO and SFPQ in PARP-dependent DNA damage repair in triple-negative breast cancer. Cell Mol. Life Sci. 2019, 76, 2015–2030. [Google Scholar] [CrossRef] [PubMed]
  225. Mathur, M.; Tucker, P.W.; Samuels, H.H. PSF is a novel corepressor that mediates its effect through Sin3A and the DNA binding domain of nuclear hormone receptors. Mol. Cell Biol. 2001, 21, 2298–2311. [Google Scholar] [CrossRef] [Green Version]
  226. Andres, M.E.; Burger, C.; Peral-Rubio, M.J.; Battaglioli, E.; Anderson, M.E.; Grimes, J.; Dallman, J.; Ballas, N.; Mandel, G. CoREST: A functional corepressor required for regulation of neural-specific gene expression. Proc. Natl. Acad. Sci. USA 1999, 96, 9873–9878. [Google Scholar] [CrossRef] [Green Version]
  227. Metzger, E.; Wissmann, M.; Yin, N.; Muller, J.M.; Schneider, R.; Peters, A.H.; Gunther, T.; Buettner, R.; Schule, R. LSD1 demethylates repressive histone marks to promote androgen-receptor-dependent transcription. Nature 2005, 437, 436–439. [Google Scholar] [CrossRef]
  228. Lee, M.G.; Wynder, C.; Cooch, N.; Shiekhattar, R. An essential role for CoREST in nucleosomal histone 3 lysine 4 demethylation. Nature 2005, 437, 432–435. [Google Scholar] [CrossRef]
  229. Shi, Y.; Lan, F.; Matson, C.; Mulligan, P.; Whetstine, J.R.; Cole, P.A.; Casero, R.A.; Shi, Y. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 2004, 119, 941–953. [Google Scholar] [CrossRef] [Green Version]
  230. Yao, H.; Goldman, D.C.; Nechiporuk, T.; Kawane, S.; McWeeney, S.K.; Tyner, J.W.; Fan, G.; Kerenyi, M.A.; Orkin, S.H.; Fleming, W.H.; et al. Corepressor Rcor1 is essential for murine erythropoiesis. Blood 2014, 123, 3175–3184. [Google Scholar] [CrossRef] [Green Version]
  231. Horlein, A.J.; Naar, A.M.; Heinzel, T.; Torchia, J.; Gloss, B.; Kurokawa, R.; Ryan, A.; Kamei, Y.; Soderstrom, M.; Glass, C.K.; et al. Ligand-independent repression by the thyroid hormone receptor mediated by a nuclear receptor co-repressor. Nature 1995, 377, 397–404. [Google Scholar] [CrossRef]
  232. Chen, J.D.; Evans, R.M. A transcriptional co-repressor that interacts with nuclear hormone receptors. Nature 1995, 377, 454–457. [Google Scholar] [CrossRef] [PubMed]
  233. Jepsen, K.; Hermanson, O.; Onami, T.M.; Gleiberman, A.S.; Lunyak, V.; McEvilly, R.J.; Kurokawa, R.; Kumar, V.; Liu, F.; Seto, E.; et al. Combinatorial roles of the nuclear receptor corepressor in transcription and development. Cell 2000, 102, 753–763. [Google Scholar] [CrossRef] [Green Version]
  234. Yamamoto, H.; Williams, E.G.; Mouchiroud, L.; Canto, C.; Fan, W.; Downes, M.; Heligon, C.; Barish, G.D.; Desvergne, B.; Evans, R.M.; et al. NCoR1 is a conserved physiological modulator of muscle mass and oxidative function. Cell 2011, 147, 827–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Jepsen, K.; Solum, D.; Zhou, T.; McEvilly, R.J.; Kim, H.J.; Glass, C.K.; Hermanson, O.; Rosenfeld, M.G. SMRT-mediated repression of an H3K27 demethylase in progression from neural stem cell to neuron. Nature 2007, 450, 415–419. [Google Scholar] [CrossRef]
  236. Mottis, A.; Mouchiroud, L.; Auwerx, J. Emerging roles of the corepressors NCoR1 and SMRT in homeostasis. Genes Dev. 2013, 27, 819–835. [Google Scholar] [CrossRef] [Green Version]
  237. Perissi, V.; Jepsen, K.; Glass, C.K.; Rosenfeld, M.G. Deconstructing repression: Evolving models of co-repressor action. Nat. Rev. Genet. 2010, 11, 109–123. [Google Scholar] [CrossRef]
  238. Stanya, K.J.; Kao, H.Y. New insights into the functions and regulation of the transcriptional corepressors SMRT and N-CoR. Cell Div. 2009, 4, 7. [Google Scholar] [CrossRef] [Green Version]
  239. Ordentlich, P.; Downes, M.; Evans, R.M. Corepressors and nuclear hormone receptor function. Curr. Top. Microbiol. Immunol. 2001, 254, 101–116. [Google Scholar]
  240. Yang, W.M.; Tsai, S.C.; Wen, Y.D.; Fejer, G.; Seto, E. Functional domains of histone deacetylase-3. J. Biol. Chem. 2002, 277, 9447–9454. [Google Scholar] [CrossRef] [Green Version]
  241. Tiefenbach, J.; Novac, N.; Ducasse, M.; Eck, M.; Melchior, F.; Heinzel, T. SUMOylation of the corepressor N-CoR modulates its capacity to repress transcription. Mol. Biol. Cell 2006, 17, 1643–1651. [Google Scholar] [CrossRef] [Green Version]
  242. Privalsky, M.L.; Goodson, M.L. Evolution of NCoR-1 and NCoR-2 corepressor alternative mRNA splicing in placental mammals. BMC Res. Notes 2019, 12, 343. [Google Scholar] [CrossRef] [PubMed]
  243. Short, S.; Peterkin, T.; Guille, M.; Patient, R.; Sharpe, C. Short linear motif acquisition, exon formation and alternative splicing determine a pathway to diversity for NCoR-family co-repressors. Open Biol. 2015, 5, 150063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Muscat, G.E.; Burke, L.J.; Downes, M. The corepressor N-CoR and its variants RIP13a and RIP13Delta1 directly interact with the basal transcription factors TFIIB, TAFII32 and TAFII70. Nucleic Acids Res. 1998, 26, 2899–2907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Goodson, M.L.; Jonas, B.A.; Privalsky, M.L. Alternative mRNA splicing of SMRT creates functional diversity by generating corepressor isoforms with different affinities for different nuclear receptors. J. Biol. Chem. 2005, 280, 7493–7503. [Google Scholar] [CrossRef] [Green Version]
  246. Jonas, B.A.; Varlakhanova, N.; Hayakawa, F.; Goodson, M.; Privalsky, M.L. Response of SMRT (silencing mediator of retinoic acid and thyroid hormone receptor) and N-CoR (nuclear receptor corepressor) corepressors to mitogen-activated protein kinase kinase kinase cascades is determined by alternative mRNA splicing. Mol. Endocrinol. 2007, 21, 1924–1939. [Google Scholar] [CrossRef] [PubMed]
  247. Jonas, B.A.; Privalsky, M.L. SMRT and N-CoR corepressors are regulated by distinct kinase signaling pathways. J. Biol. Chem. 2004, 279, 54676–54686. [Google Scholar] [CrossRef] [Green Version]
  248. Hong, S.H.; Privalsky, M.L. The SMRT corepressor is regulated by a MEK-1 kinase pathway: Inhibition of corepressor function is associated with SMRT phosphorylation and nuclear export. Mol. Cell Biol. 2000, 20, 6612–6625. [Google Scholar] [CrossRef] [PubMed]
  249. Hong, S.H.; Wong, C.W.; Privalsky, M.L. Signaling by tyrosine kinases negatively regulates the interaction between transcription factors and SMRT (silencing mediator of retinoic acid and thyroid hormone receptor) corepressor. Mol. Endocrinol. 1998, 12, 1161–1171. [Google Scholar] [CrossRef]
  250. Sternberg, P.W.; Stern, M.J.; Clark, I.; Herskowitz, I. Activation of the yeast HO gene by release from multiple negative controls. Cell 1987, 48, 567–577. [Google Scholar] [CrossRef]
  251. Nasmyth, K.; Stillman, D.; Kipling, D. Both positive and negative regulators of HO transcription are required for mother-cell-specific mating-type switching in yeast. Cell 1987, 48, 579–587. [Google Scholar] [CrossRef]
  252. Wang, H.; Stillman, D.J. In vitro regulation of a SIN3-dependent DNA-binding activity by stimulatory and inhibitory factors. Proc. Natl. Acad. Sci. USA 1990, 87, 9761–9765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. He, Y.; Radhakrishnan, I. Solution NMR studies of apo-mSin3A and -mSin3B reveal that the PAH1 and PAH2 domains are structurally independent. Protein Sci. 2008, 17, 171–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Dannenberg, J.H.; David, G.; Zhong, S.; van der Torre, J.; Wong, W.H.; Depinho, R.A. mSin3A corepressor regulates diverse transcriptional networks governing normal and neoplastic growth and survival. Genes Dev. 2005, 19, 1581–1595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Le Guezennec, X.; Vermeulen, M.; Stunnenberg, H.G. Molecular characterization of Sin3 PAH-domain interactor specificity and identification of PAH partners. Nucleic Acids Res. 2006, 34, 3929–3937. [Google Scholar] [CrossRef] [Green Version]
  256. Dong, X.; Sweet, J.; Challis, J.R.; Brown, T.; Lye, S.J. Transcriptional activity of androgen receptor is modulated by two RNA splicing factors, PSF and p54nrb. Mol. Cell Biol. 2007, 27, 4863–4875. [Google Scholar] [CrossRef] [Green Version]
  257. Awad, T.A.; Bigler, J.; Ulmer, J.E.; Hu, Y.J.; Moore, J.M.; Lutz, M.; Neiman, P.E.; Collins, S.J.; Renkawitz, R.; Lobanenkov, V.V.; et al. Negative transcriptional regulation mediated by thyroid hormone response element 144 requires binding of the multivalent factor CTCF to a novel target DNA sequence. J. Biol. Chem. 1999, 274, 27092–27098. [Google Scholar] [CrossRef] [Green Version]
  258. Bigler, J.; Eisenman, R.N. Novel location and function of a thyroid hormone response element. EMBO J. 1995, 14, 5710–5723. [Google Scholar] [CrossRef]
  259. Lutz, M.; Baniahmad, A.; Renkawitz, R. Modulation of thyroid hormone receptor silencing function by co-repressors and a synergizing transcription factor. Biochem. Soc. Trans. 2000, 28, 386–389. [Google Scholar] [CrossRef]
  260. Ellison-Zelski, S.J.; Alarid, E.T. Maximum growth and survival of estrogen receptor-alpha positive breast cancer cells requires the Sin3A transcriptional repressor. Mol. Cancer 2010, 9, 263. [Google Scholar] [CrossRef] [Green Version]
  261. Ellison-Zelski, S.J.; Solodin, N.M.; Alarid, E.T. Repression of ESR1 through actions of estrogen receptor alpha and Sin3A at the proximal promoter. Mol. Cell Biol. 2009, 29, 4949–4958. [Google Scholar] [CrossRef] [Green Version]
  262. Moehren, U.; Dressel, U.; Reeb, C.A.; Vaisanen, S.; Dunlop, T.W.; Carlberg, C.; Baniahmad, A. The highly conserved region of the co-repressor Sin3A functionally interacts with the co-repressor Alien. Nucleic Acids Res. 2004, 32, 2995–3004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Polly, P.; Herdick, M.; Moehren, U.; Baniahmad, A.; Heinzel, T.; Carlberg, C. VDR-Alien: A novel, DNA-selective vitamin D(3) receptor-corepressor partnership. FASEB J. 2000, 14, 1455–1463. [Google Scholar] [PubMed]
  264. Nagase, T.; Nakayama, M.; Nakajima, D.; Kikuno, R.; Ohara, O. Prediction of the coding sequences of unidentified human genes. XX. The complete sequences of 100 new cDNA clones from brain which code for large proteins in vitro. DNA Res. 2001, 8, 85–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Shalom-Barak, T.; Liersemann, J.; Memari, B.; Flechner, L.; Devor, C.E.; Bernardo, T.M.; Kim, S.; Matsumoto, N.; Friedman, S.L.; Evans, R.M.; et al. Ligand-Dependent Corepressor (LCoR) Is a Rexinoid-Inhibited Peroxisome Proliferator-Activated Receptor gamma-Retinoid X Receptor alpha Coactivator. Mol. Cell Biol. 2018, 38, e00107-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Palijan, A.; Fernandes, I.; Verway, M.; Kourelis, M.; Bastien, Y.; Tavera-Mendoza, L.E.; Sacheli, A.; Bourdeau, V.; Mader, S.; White, J.H. Ligand-dependent corepressor LCoR is an attenuator of progesterone-regulated gene expression. J. Biol. Chem. 2009, 284, 30275–30287. [Google Scholar] [CrossRef] [Green Version]
  267. Palijan, A.; Fernandes, I.; Bastien, Y.; Tang, L.; Verway, M.; Kourelis, M.; Tavera-Mendoza, L.E.; Li, Z.; Bourdeau, V.; Mader, S.; et al. Function of histone deacetylase 6 as a cofactor of nuclear receptor coregulator LCoR. J. Biol. Chem. 2009, 284, 30264–30274. [Google Scholar] [CrossRef] [Green Version]
  268. 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]
  269. Jalaguier, S.; Teyssier, C.; Nait Achour, T.; Lucas, A.; Bonnet, S.; Rodriguez, C.; Elarouci, N.; Lapierre, M.; Cavailles, V. Complex regulation of LCoR signaling in breast cancer cells. Oncogene 2017, 36, 4790–4801. [Google Scholar] [CrossRef]
  270. Calderon, M.R.; Verway, M.; An, B.S.; DiFeo, A.; Bismar, T.A.; Ann, D.K.; Martignetti, J.A.; Shalom-Barak, T.; White, J.H. Ligand-dependent corepressor (LCoR) recruitment by Kruppel-like factor 6 (KLF6) regulates expression of the cyclin-dependent kinase inhibitor CDKN1A gene. J. Biol. Chem. 2012, 287, 8662–8674. [Google Scholar] [CrossRef] [Green Version]
  271. Kwok, R.P.; Lundblad, J.R.; Chrivia, J.C.; Richards, J.P.; Bachinger, H.P.; Brennan, R.G.; Roberts, S.G.; Green, M.R.; Goodman, R.H. Nuclear protein CBP is a coactivator for the transcription factor CREB. Nature 1994, 370, 223–226. [Google Scholar] [CrossRef]
  272. Smith, C.L.; Onate, S.A.; Tsai, M.J.; O’Malley, B.W. CREB binding protein acts synergistically with steroid receptor coactivator-1 to enhance steroid receptor-dependent transcription. Proc. Natl. Acad. Sci. USA 1996, 93, 8884–8888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Heery, D.M.; Kalkhoven, E.; Hoare, S.; Parker, M.G. A signature motif in transcriptional co-activators mediates binding to nuclear receptors. Nature 1997, 387, 733–736. [Google Scholar] [CrossRef] [PubMed]
  274. Kamei, Y.; Xu, L.; Heinzel, T.; Torchia, J.; Kurokawa, R.; Gloss, B.; Lin, S.C.; Heyman, R.A.; Rose, D.W.; Glass, C.K.; et al. A CBP integrator complex mediates transcriptional activation and AP-1 inhibition by nuclear receptors. Cell 1996, 85, 403–414. [Google Scholar] [CrossRef] [Green Version]
  275. Zhou, G.; Cummings, R.; Li, Y.; Mitra, S.; Wilkinson, H.A.; Elbrecht, A.; Hermes, J.D.; Schaeffer, J.M.; Smith, R.G.; Moller, D.E. Nuclear receptors have distinct affinities for coactivators: Characterization by fluorescence resonance energy transfer. Mol. Endocrinol. 1998, 12, 1594–1604. [Google Scholar] [CrossRef]
  276. McKenna, N.J.; Nawaz, Z.; Tsai, S.Y.; Tsai, M.J.; O’Malley, B.W. Distinct steady-state nuclear receptor coregulator complexes exist in vivo. Proc. Natl. Acad. Sci. USA 1998, 95, 11697–11702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Goo, Y.H.; Na, S.Y.; Zhang, H.; Xu, J.; Hong, S.; Cheong, J.; Lee, S.K.; Lee, J.W. Interactions between activating signal cointegrator-2 and the tumor suppressor retinoblastoma in androgen receptor transactivation. J. Biol. Chem. 2004, 279, 7131–7135. [Google Scholar] [CrossRef] [Green Version]
  278. Altwegg, K.A.; Vadlamudi, R.K. Role of estrogen receptor coregulators in endocrine resistant breast cancer. Explor. Target. Antitumor Ther. 2021, 2, 385–400. [Google Scholar] [CrossRef]
  279. Turner, N.C.; Neven, P.; Loibl, S.; Andre, F. Advances in the treatment of advanced oestrogen-receptor-positive breast cancer. Lancet 2017, 389, 2403–2414. [Google Scholar] [CrossRef]
  280. Musgrove, E.A.; Sutherland, R.L. Biological determinants of endocrine resistance in breast cancer. Nat. Rev. Cancer 2009, 9, 631–643. [Google Scholar] [CrossRef]
  281. Liu, J.; Liu, Z.; Li, M.; Tang, W.; Pratap, U.P.; Luo, Y.; Altwegg, K.A.; Li, X.; Zou, Y.; Zhu, H.; et al. Interaction of transcription factor AP-2 gamma with proto-oncogene PELP1 promotes tumorigenesis by enhancing RET signaling. Mol. Oncol. 2021, 15, 1146–1161. [Google Scholar] [CrossRef]
  282. Wu, Y.; Zhang, Z.; Cenciarini, M.E.; Proietti, C.J.; Amasino, M.; Hong, T.; Yang, M.; Liao, Y.; Chiang, H.C.; Kaklamani, V.G.; et al. Tamoxifen Resistance in Breast Cancer Is Regulated by the EZH2-ERalpha-GREB1 Transcriptional Axis. Cancer Res. 2018, 78, 671–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Yang, J.; Wei, X.; Tufan, T.; Kuscu, C.; Unlu, H.; Farooq, S.; Demirtas, E.; Paschal, B.M.; Adli, M. Recurrent mutations at estrogen receptor binding sites alter chromatin topology and distal gene expression in breast cancer. Genome Biol. 2018, 19, 190. [Google Scholar] [CrossRef] [PubMed]
  284. Toy, W.; Weir, H.; Razavi, P.; Lawson, M.; Goeppert, A.U.; Mazzola, A.M.; Smith, A.; Wilson, J.; Morrow, C.; Wong, W.L.; et al. Activating ESR1 Mutations Differentially Affect the Efficacy of ER Antagonists. Cancer Discov. 2017, 7, 277–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Jeselsohn, R.; Buchwalter, G.; De Angelis, C.; Brown, M.; Schiff, R. ESR1 mutations-a mechanism for acquired endocrine resistance in breast cancer. Nat. Rev. Clin. Oncol. 2015, 12, 573–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Li, Y.; Coons, L.A.; Houtman, R.; Carlson, K.E.; Martin, T.A.; Mayne, C.G.; Melchers, D.; Jefferson, T.B.; Ramsey, J.T.; Katzenellenbogen, J.A.; et al. A mutant form of ERalpha associated with estrogen insensitivity affects the coupling between ligand binding and coactivator recruitment. Sci. Signal. 2020, 13, eaaw4653. [Google Scholar] [CrossRef] [PubMed]
  287. Fanning, S.W.; Mayne, C.G.; Dharmarajan, V.; Carlson, K.E.; Martin, T.A.; Novick, S.J.; Toy, W.; Green, B.; Panchamukhi, S.; Katzenellenbogen, B.S.; et al. Estrogen receptor alpha somatic mutations Y537S and D538G confer breast cancer endocrine resistance by stabilizing the activating function-2 binding conformation. Elife 2016, 5, e12792. [Google Scholar] [CrossRef]
  288. Magne Nde, C.B.; Casas Gimeno, G.; Docanto, M.; Knower, K.C.; Young, M.J.; Buehn, J.; Sayed, E.; Clyne, C.D. Timeless Is a Novel Estrogen Receptor Co-activator Involved in Multiple Signaling Pathways in MCF-7 Cells. J. Mol. Biol. 2018, 430, 1531–1543. [Google Scholar] [CrossRef]
  289. Abida, W.; Cyrta, J.; Heller, G.; Prandi, D.; Armenia, J.; Coleman, I.; Cieslik, M.; Benelli, M.; Robinson, D.; Van Allen, E.M.; et al. Genomic correlates of clinical outcome in advanced prostate cancer. Proc. Natl. Acad. Sci. USA 2019, 116, 11428–11436. [Google Scholar] [CrossRef] [Green Version]
  290. Feng, S.; Tang, Q.; Sun, M.; Chun, J.Y.; Evans, C.P.; Gao, A.C. Interleukin-6 increases prostate cancer cells resistance to bicalutamide via TIF2. Mol. Cancer Ther. 2009, 8, 665–671. [Google Scholar] [CrossRef] [Green Version]
  291. Watson, R.W.; Azam, H.; Aura, C.; Russell, N.; McCormack, J.; Corey, E.; Morrissey, C.; Crown, J.; Gallagher, W.M.; Prencipe, M. Inhibition of Serum Response Factor Improves Response to Enzalutamide in Prostate Cancer. Cancers 2020, 12, 3540. [Google Scholar] [CrossRef]
  292. Zhang, Z.; Yamashita, H.; Toyama, T.; Sugiura, H.; Ando, Y.; Mita, K.; Hamaguchi, M.; Hara, Y.; Kobayashi, S.; Iwase, H. NCOR1 mRNA is an independent prognostic factor for breast cancer. Cancer Lett. 2006, 237, 123–129. [Google Scholar] [CrossRef] [PubMed]
  293. Girault, I.; Lerebours, F.; Amarir, S.; Tozlu, S.; Tubiana-Hulin, M.; Lidereau, R.; Bieche, I. Expression analysis of estrogen receptor alpha coregulators in breast carcinoma: Evidence that NCOR1 expression is predictive of the response to tamoxifen. Clin. Cancer Res. 2003, 9, 1259–1266. [Google Scholar] [PubMed]
  294. Kim, J.Y.; Son, Y.L.; Lee, Y.C. Involvement of SMRT corepressor in transcriptional repression by the vitamin D receptor. Mol. Endocrinol. 2009, 23, 251–264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Khanim, F.L.; Gommersall, L.M.; Wood, V.H.; Smith, K.L.; Montalvo, L.; O’Neill, L.P.; Xu, Y.; Peehl, D.M.; Stewart, P.M.; Turner, B.M.; et al. Altered SMRT levels disrupt vitamin D3 receptor signalling in prostate cancer cells. Oncogene 2004, 23, 6712–6725. [Google Scholar] [CrossRef] [Green Version]
  296. Trtkova, K.; Paskova, L.; Matijescukova, N.; Kolar, Z. Formation of AR-SMRT binding in prostate cancer cells treated with natural histone deacetylase inhibitor. Cancer Biomark. 2010, 7, 79–90. [Google Scholar] [CrossRef]
  297. Quigley, D.A.; Dang, H.X.; Zhao, S.G.; Lloyd, P.; Aggarwal, R.; Alumkal, J.J.; Foye, A.; Kothari, V.; Perry, M.D.; Bailey, A.M.; et al. Genomic Hallmarks and Structural Variation in Metastatic Prostate Cancer. Cell 2018, 174, 758–769.e759. [Google Scholar] [CrossRef] [Green Version]
  298. Viswanathan, S.R.; Ha, G.; Hoff, A.M.; Wala, J.A.; Carrot-Zhang, J.; Whelan, C.W.; Haradhvala, N.J.; Freeman, S.S.; Reed, S.C.; Rhoades, J.; et al. Structural Alterations Driving Castration-Resistant Prostate Cancer Revealed by Linked-Read Genome Sequencing. Cell 2018, 174, 433–447.e419. [Google Scholar] [CrossRef] [Green Version]
  299. Guo, Y.; Perez, A.A.; Hazelett, D.J.; Coetzee, G.A.; Rhie, S.K.; Farnham, P.J. CRISPR-mediated deletion of prostate cancer risk-associated CTCF loop anchors identifies repressive chromatin loops. Genome Biol. 2018, 19, 160. [Google Scholar] [CrossRef] [PubMed]
  300. Rhie, S.K.; Perez, A.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]
  301. Taberlay, P.C.; Achinger-Kawecka, J.; Lun, A.T.; Buske, F.A.; Sabir, K.; Gould, C.M.; Zotenko, E.; Bert, S.A.; Giles, K.A.; Bauer, D.C.; 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]
  302. Flavahan, W.A.; Drier, Y.; Johnstone, S.E.; Hemming, M.L.; Tarjan, D.R.; Hegazi, E.; Shareef, S.J.; Javed, N.M.; Raut, C.P.; Eschle, B.K.; et al. Altered chromosomal topology drives oncogenic programs in SDH-deficient GISTs. Nature 2019, 575, 229–233. [Google Scholar] [CrossRef] [PubMed]
  303. Bell, C.C.; Fennell, K.A.; Chan, Y.C.; Rambow, F.; Yeung, M.M.; Vassiliadis, D.; Lara, L.; Yeh, P.; Martelotto, L.G.; Rogiers, A.; et al. Targeting enhancer switching overcomes non-genetic drug resistance in acute myeloid leukaemia. Nat. Commun. 2019, 10, 2723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Labbe, D.P.; Sweeney, C.J.; Brown, M.; Galbo, P.; Rosario, S.; Wadosky, K.M.; Ku, S.Y.; Sjostrom, M.; Alshalalfa, M.; Erho, N.; et al. TOP2A and EZH2 Provide Early Detection of an Aggressive Prostate Cancer Subgroup. Clin. Cancer Res. 2017, 23, 7072–7083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Rotinen, M.; You, S.; Yang, J.; Coetzee, S.G.; Reis-Sobreiro, M.; Huang, W.C.; Huang, F.; Pan, X.; Yanez, A.; Hazelett, D.J.; et al. ONECUT2 is a targetable master regulator of lethal prostate cancer that suppresses the androgen axis. Nat. Med. 2018, 24, 1887–1898. [Google Scholar] [CrossRef]
  306. Nerlakanti, N.; Yao, J.; Nguyen, D.T.; Patel, A.K.; Eroshkin, A.M.; Lawrence, H.R.; Ayaz, M.; Kuenzi, B.M.; Agarwal, N.; Chen, Y.; et al. Targeting the BRD4-HOXB13 Coregulated Transcriptional Networks with Bromodomain-Kinase Inhibitors to Suppress Metastatic Castration-Resistant Prostate Cancer. Mol. Cancer Ther. 2018, 17, 2796–2810. [Google Scholar] [CrossRef] [Green Version]
  307. Stelloo, S.; Nevedomskaya, E.; Kim, Y.; Schuurman, K.; Valle-Encinas, E.; Lobo, J.; Krijgsman, O.; Peeper, D.S.; Chang, S.L.; Feng, F.Y.; et al. Integrative epigenetic taxonomy of primary prostate cancer. Nat. Commun. 2018, 9, 4900. [Google Scholar] [CrossRef]
  308. Patten, D.K.; Corleone, G.; Gyorffy, B.; Perone, Y.; Slaven, N.; Barozzi, I.; Erdos, E.; Saiakhova, A.; Goddard, K.; Vingiani, A.; et al. Enhancer mapping uncovers phenotypic heterogeneity and evolution in patients with luminal breast cancer. Nat. Med. 2018, 24, 1469–1480. [Google Scholar] [CrossRef]
  309. Hodgson, M.C.; Astapova, I.; Cheng, S.; Lee, L.J.; Verhoeven, M.C.; Choi, E.; Balk, S.P.; Hollenberg, A.N. The androgen receptor recruits nuclear receptor CoRepressor (N-CoR) in the presence of mifepristone via its N and C termini revealing a novel molecular mechanism for androgen receptor antagonists. J. Biol. Chem. 2005, 280, 6511–6519. [Google Scholar] [CrossRef] [Green Version]
  310. Song, L.N.; Coghlan, M.; Gelmann, E.P. Antiandrogen effects of mifepristone on coactivator and corepressor interactions with the androgen receptor. Mol. Endocrinol. 2004, 18, 70–85. [Google Scholar] [CrossRef] [Green Version]
  311. Masiello, D.; Chen, S.Y.; Xu, Y.; Verhoeven, M.C.; Choi, E.; Hollenberg, A.N.; Balk, S.P. Recruitment of beta-catenin by wild-type or mutant androgen receptors correlates with ligand-stimulated growth of prostate cancer cells. Mol. Endocrinol. 2004, 18, 2388–2401. [Google Scholar] [CrossRef] [Green Version]
  312. Yachie, N.; Petsalaki, E.; Mellor, J.C.; Weile, J.; Jacob, Y.; Verby, M.; Ozturk, S.B.; Li, S.; Cote, A.G.; Mosca, R.; et al. Pooled-matrix protein interaction screens using Barcode Fusion Genetics. Mol. Syst. Biol. 2016, 12, 863. [Google Scholar] [CrossRef] [PubMed]
  313. Wang, J.; Huo, K.; Ma, L.; Tang, L.; Li, D.; Huang, X.; Yuan, Y.; Li, C.; Wang, W.; Guan, W.; et al. Toward an understanding of the protein interaction network of the human liver. Mol. Syst. Biol. 2011, 7, 536. [Google Scholar] [CrossRef] [PubMed]
  314. Liao, G.; Chen, L.Y.; Zhang, A.; Godavarthy, A.; Xia, F.; Ghosh, J.C.; Li, H.; Chen, J.D. Regulation of androgen receptor activity by the nuclear receptor corepressor SMRT. J. Biol. Chem. 2003, 278, 5052–5061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Liu, X.; Salokas, K.; Weldatsadik, R.G.; Gawriyski, L.; Varjosalo, M. Combined proximity labeling and affinity purification-mass spectrometry workflow for mapping and visualizing protein interaction networks. Nat. Protoc. 2020, 15, 3182–3211. [Google Scholar] [CrossRef]
  316. Mohammed, H.; Taylor, C.; Brown, G.D.; Papachristou, E.K.; Carroll, J.S.; D’Santos, C.S. Rapid immunoprecipitation mass spectrometry of endogenous proteins (RIME) for analysis of chromatin complexes. Nat. Protoc. 2016, 11, 316–326. [Google Scholar] [CrossRef]
  317. Liu, X.; Zhang, Y.; Chen, Y.; Li, M.; Zhou, F.; Li, K.; Cao, H.; Ni, M.; Liu, Y.; Gu, Z.; et al. In Situ Capture of Chromatin Interactions by Biotinylated dCas9. Cell 2017, 170, 1028–1043.e1019. [Google Scholar] [CrossRef] [Green Version]
  318. Gao, X.D.; Tu, L.C.; Mir, A.; Rodriguez, T.; Ding, Y.; Leszyk, J.; Dekker, J.; Shaffer, S.A.; Zhu, L.J.; Wolfe, S.A.; et al. C-BERST: Defining subnuclear proteomic landscapes at genomic elements with dCas9-APEX2. Nat. Methods 2018, 15, 433–436. [Google Scholar] [CrossRef]
  319. Purslow, J.A.; Khatiwada, B.; Bayro, M.J.; Venditti, V. NMR Methods for Structural Characterization of Protein-Protein Complexes. Front. Mol. Biosci. 2020, 7, 9. [Google Scholar] [CrossRef]
  320. Rainey, K.H.; Patterson, G.H. Photoswitching FRET to monitor protein-protein interactions. Proc. Natl. Acad. Sci. USA 2019, 116, 864–873. [Google Scholar] [CrossRef] [Green Version]
  321. Chen, F.; Liu, J.; Huang, M.; Hu, M.; Su, Y.; Zhang, X.K. Identification of a New RXRalpha Antagonist Targeting the Coregulator-Binding Site. ACS Med. Chem. Lett. 2014, 5, 736–741. [Google Scholar] [CrossRef]
  322. Mi, H.; Muruganujan, A.; Ebert, D.; Huang, X.; Thomas, P.D. PANTHER version 14: More genomes, a new PANTHER GO-slim and improvements in enrichment analysis tools. Nucleic Acids Res. 2019, 47, D419–D426. [Google Scholar] [CrossRef] [PubMed]
  323. Durinck, S.; Spellman, P.T.; Birney, E.; Huber, W. Mapping identifiers for the integration of genomic datasets with the R/Bioconductor package biomaRt. Nat. Protoc. 2009, 4, 1184–1191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Franke, J.; Neumann, M.H. Bootstrapping neural networks. Neural Comput. 2000, 12, 1929–1949. [Google Scholar] [CrossRef] [PubMed]
  325. Armenia, J.; Wankowicz, S.A.M.; Liu, D.; Gao, J.; Kundra, R.; Reznik, E.; Chatila, W.K.; Chakravarty, D.; Han, G.C.; Coleman, I.; et al. The long tail of oncogenic drivers in prostate cancer. Nat. Genet. 2018, 50, 645–651. [Google Scholar] [CrossRef] [PubMed]
  326. Wanior, M.; Kramer, A.; Knapp, S.; Joerger, A.C. Exploiting vulnerabilities of SWI/SNF chromatin remodelling complexes for cancer therapy. Oncogene 2021, 40, 3637–3654. [Google Scholar] [CrossRef]
  327. Lord, C.J.; Ashworth, A. BRCAness revisited. Nat. Rev. Cancer 2016, 16, 110–120. [Google Scholar] [CrossRef]
  328. Ernst, J.; Kellis, M. Chromatin-state discovery and genome annotation with ChromHMM. Nat. Protoc. 2017, 12, 2478–2492. [Google Scholar] [CrossRef]
  329. Kim, D.H.; Sun, D.; Storck, W.K.; Welker Leng, K.; Jenkins, C.; Coleman, D.J.; Sampson, D.; Guan, X.; Kumaraswamy, A.; Rodansky, E.S.; et al. BET Bromodomain Inhibition Blocks an AR-Repressed, E2F1-Activated Treatment-Emergent Neuroendocrine Prostate Cancer Lineage Plasticity Program. Clin. Cancer Res. 2021, 27, 4923–4936. [Google Scholar] [CrossRef]
  330. Whyte, W.A.; Orlando, D.A.; Hnisz, D.; Abraham, B.J.; Lin, C.Y.; Kagey, M.H.; Rahl, P.B.; Lee, T.I.; Young, R.A. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 2013, 153, 307–319. [Google Scholar] [CrossRef] [Green Version]
  331. Layer, R.M.; Pedersen, B.S.; DiSera, T.; Marth, G.T.; Gertz, J.; Quinlan, A.R. GIGGLE: A search engine for large-scale integrated genome analysis. Nat. Methods 2018, 15, 123–126. [Google Scholar] [CrossRef] [Green Version]
  332. Long, M.D.; van Den Berg, P.; Singh, P.K.; .Battaglia, S.; Campbell, M.J. Integrative genomic analysis in K562 chronic myelogenous leukemia cells reveals that proximal NCOR1 binding positively regulates genes that govern erythroid differentiation and Imatinib sensitivity. Nucleic Acids Res. 2015, 43, 7330–7348. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic model of NR-target gene transcriptional activation (left panel) and repression (right panel).
Figure 1. Schematic model of NR-target gene transcriptional activation (left panel) and repression (right panel).
Cancers 14 02402 g001
Table 1. Examples of coactivators that interact with nuclear receptors and impact their functions.
Table 1. Examples of coactivators that interact with nuclear receptors and impact their functions.
GeneSuperfamily *Functional Complex *Example of NR Interaction **Gene Defects ***References
SMARCA4PBAF complexNucleosome positioningVDR, AR, PPARγBRCA—2.6% point mutations, 0.6% CNVs.
CC—4.9% point mutations, 1.67% CNVs.
OVCA—5.8% point mutations, 3.3% CNVs.
PRAD—11.66% point mutations.
[119,120,121]
TRRAPSAGA coactivator complexHistone modificationsLXR, FXR, ERα,BRCA—5% point mutations, 0.3% CNVs.
CC—8.2% point mutations, 0.33% CNVs.
OVCA—4.9% point mutations, 0.7% CNVs.
PRAD—4.9% point mutations, 0.5% CNVs.
[122,123,124]
CARM17BS protein arginine methyltransferasesHistone modificationsERα, AR, FXR, PPARγ,BRCA—1.6% point mutations, 0.7% CNVs.
OVCA—1.3% point mutations, 3.3% CNVs.
[125,126,127,128]
NCOA1Lysine acetyl transferaseHistone modificationsRARβ, RXRα, PPARγ, ERα, TRα, GRBRCA—6.1% point mutations. CAN—0.1% point mutations.
PRAD—6.3% point mutations. OC—5.7% point mutations, 0.4% CNVs.
[129,130,131]
PARP1Poly (ADP-ribose) polymerasesDNA repair, link to MediatorPPARγ, ERα, NR4A4, RARβ, TRBRCA—1.9% point mutations, 7.3% CNVs.
CC—1.8% point mutations, 1.6% CNVs.
OVCA—1.2% point mutations, 1.1% CNVs.
PRAD—1.2% point mutations, 0.1% CNVs.
[132,133,134,135]
TACC1TACC familySignal integrators and transducers, scaffolds, and adaptorsTHRβ, THRα, PPARγ, RARα, RARγ and RXRαBRCA—3.7% point mutations, 8.8% CNVs.
CC—1.2% point mutations, 1% CNVs.
OVCA—2.8% point mutations, 1.7% CNVs.
PRAD—2.8% point mutations, 0.7% CNVs.
[113,116]
PPARGC1ARNA-binding motif containingCoactivator complexes for metabolic programmingInteracts with PPARγ, PPARα, TRBRCA—3.2% point mutations, 0.1% CNVs.
OVCA—4.7% point mutations, 0.3% CNVs.
PRAD—4.5% point mutations.
[136,137]
* All the information obtained from HGNC and UniprotKB websites, and literature review. ** Nuclear receptor-related interactions among top 25 interactions on STRING tool. *** Mutation percentages are obtained from COSMIC repository.
Table 2. Examples of corepressors that interact with nuclear receptors and impact their functions.
Table 2. Examples of corepressors that interact with nuclear receptors and impact their functions.
GeneSuperfamily *Functional Complex *Example of NR Interaction **Genetic Defects ***References
RCOR1Myb/SANT domain-containingHistone modificationsGR, ARBRCA—2.9% point mutations, 0.1% CNVs. PRAD—3.0% point mutations. THCA—0.6% point mutations. OVCA—3.3% point mutations, 0.5% CNVs.[204,205,206,207,208]
MECP2Methyl-CpG binding domain-containingHistone modificationsNCOR2, NCOR1BRCA—2.3% point mutations, 0.5% CNVs. PRAD—1.4% point mutations, 0.2% point mutations. THCA—0.8% point mutations, 0.4 CNVs.[209,210,211]
SIN3ASIN3 histone deacetylase complex subunitsNucleosome positioningAR, ERα, RARβ, NR0B1/SHPBRCA—2.7% point mutations, 0.1 CNVs. PRAD—2.4% point mutations.
THCA—0.8% point mutations.
[212,213,214,215,216,217,218,219]
NCOR1NCoR/SMRT transcriptional repression complex subunitsHistone modificationsAR, TR, RAR, VDR, PPARα/γ BRCA—5.6% point mutations, 0.8% CNVs. PRAD—4.9% point mutations, 0.1% CNVs. THCA—1.8% point mutations. OVCA—3.4% point mutations, 0.1% CNVs.[33,35,38]
LCOR Histone modificationsERα, ERβ, AR, PGR, RARα, RARβ, RARγ, RXRα VDRBRCA—2.7% point mutations. PRAD—2.9% point mutations, 0.1% CNVs. THCA—0.3% point mutations. OVCA—2.7% point mutations, 0.1% CNVs.[220,221]
SFPQRNA-binding motif containingRNA splicing and metabolismSIN3A, THRα, RXRαBRCA—0.8% point mutations, 0.1% CNVs. PRAD—0.7% point mutations. THCA—2.1% point mutations.[222,223,224,225]
* All the information obtained from HGNC, UniprotKB website and the literature review. ** Nuclear receptor-related among top 25 interactions on STRING tool. *** Mutation percentages are obtained from COSMIC repository.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jafari, H.; Hussain, S.; Campbell, M.J. Nuclear Receptor Coregulators in Hormone-Dependent Cancers. Cancers 2022, 14, 2402. https://doi.org/10.3390/cancers14102402

AMA Style

Jafari H, Hussain S, Campbell MJ. Nuclear Receptor Coregulators in Hormone-Dependent Cancers. Cancers. 2022; 14(10):2402. https://doi.org/10.3390/cancers14102402

Chicago/Turabian Style

Jafari, Hedieh, Shahid Hussain, and Moray J. Campbell. 2022. "Nuclear Receptor Coregulators in Hormone-Dependent Cancers" Cancers 14, no. 10: 2402. https://doi.org/10.3390/cancers14102402

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