Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Jan 1.
Published in final edited form as: Curr Probl Cancer. 2014 Nov 25;39(1):17–28. doi: 10.1016/j.currproblcancer.2014.11.004

The Biology of Castrate Resistant Prostate Cancer

Fei Lian 1, Nitya Sharma 2, Josue D Moran 3, Carlos S Moreno 1,2,3,4,5,6,*
PMCID: PMC4314372  NIHMSID: NIHMS652713  PMID: 25547388

Background

Prostate cancer is one of the most commonly diagnosed cancers and the second leading cause of cancer death for males in the United States1. Although there are recently approved therapeutic options for men with advanced and metastatic disease, the unfortunate reality is that advanced prostate cancer is inevitably fatal. Hormonal androgen deprivation therapy (ADT)2 is the standard of care once a patient has recurrent disease following primary surgical or radiation therapy; however, the benefits from ADT are typically short-lived. Recurrent disease that follows ADT treatment is termed castrate-resistant prostate cancer (CRPC)2, which is the most aggressive and lethal form of prostate cancer. The current treatment regime for CRPC consists primarily of chemotherapy with docetaxel, but other agents such as the immunotherapy agent sipuleucel-T, the taxane cabazitaxel, the CYP17 inhibitor abiraterone acetate, and the androgen receptor (AR) antagonist enzalutamide are also FDA approved for the treatment of CRPC3-8. While there are clinical trials testing the potential therapeutic benefits of many other compounds, CRPC remains an incurable disease4,6,9. Recent advancements in our understanding of the biochemical and genetic pathways critical to CRPC progression have identified many novel potential druggable targets as well as biomarkers of disease progression10,11. This review will focus on the biological aspects of aggressive prostate cancer, biomarkers, therapeutic targets, and future directions for the treatment of CRPC.

The Androgen Receptor

It is well-documented that prostate cancer cells depend on the androgen receptor (AR) and steroid hormones for continued oncogenic growth. Though ADT targets these pathways, this therapy can also select for tumor cells with several alternative survival mechanisms that allow AR to continue to function in the absence of circulating androgens. One such mechanism that generates androgen insensitivity seen in CRPC involves either the amplification of the AR gene or increased AR mRNA expression, along with a concomitant increase in the abundance of AR protein expression in CRPC tumor cells12. Another survival mechanism relies on the specific mutations in AR that enable binding and activation of this receptor by other steroidal hormones such as progesterone, hydrocortisone, estradiol13, or even AR antagonists such as flutamide14. One critical pathway for CRPC is the intratumoral synthesis of steroidal hormones via upregulation of the cytochrome P450 gene CYP17A1, steroid-5-alpha-reductase, alpha polypeptide 1 (SRD5A1), aldo-keto reductase family 1, member C3 (AKR1C3), and hydroxy-delta-5-steroid dehydrogenase, 3 beta- and steroid delta-isomerase 2 (HSD3B2)13. These enzymes contribute to de novo synthesis of dihydrotestosterone (DHT) from endogenous cholesterol, as well as from uptake of exogenous dehydroepiandrosterone sulfate (DHEA-S), providing the rationale for treatment with the CYP17A1 inhibitor abiraterone. Furthermore, post-translational modifications such as aberrant phosphorylation of AR, by the Src tyrosine kinase on Y534 leads to increased sensitivity of the AR to low levels of circulating androgens15. Wang et al. showed through gene expression profiling that AR in CRPC selectively unregulates cell cycle M-phase genes promoting CRPC growth2, suggesting that AR in CRPC may regulate a distinct set of genes independent of the typical AR-dependent transcriptional program.

Genome wide localization analysis via chromatin immunoprecipitation and sequencing (ChIP-seq) of AR in CRPC tissues has identified a set of genes regulated by AR that are distinct from those observed from cultured prostate cancer cell lines16. Analysis of DNA binding motifs adjacent to AR binding sites determined that unlike cultured cells in which AR interacts with FOXA1 and NF-1, in vivo AR interacts with E2F, MYC, and STAT516.

PI3K-AKT-mTOR pathway

Activation of the PI3K-AKT-mTOR pathway is extremely common, if not universal, in CRPC. It is critical for the regulation of cell survival, apoptosis, cell proliferation, autophagy, metabolism, and protein synthesis, and has been extensively studied in prostate cancer17. PI3K is activated by a wide range of growth factor receptors and signaling pathways, including epidermal growth factor receptor (EGFR), insulin-like growth factor 1 receptor (IGF-1R), fibroblast growth factor receptor (FGFR), and platelet-derived growth factor receptor (PDGFR). Activated PI3K activates PDK1, which in turn activates AKT, while PI3K can be inactivated by the PTEN tumor suppressor18-20. AKT separately phosphorylates and activates mTOR, which promotes cell cycle progression and growth and provides positive feedback via phosphorylation of AKT. Aberrant constitutive AKT activation is one of the most frequent pathway alterations observed in a number of different cancers. Recently, Chin et al. demonstrated that the AKT2 isoform was specifically required for cellular maintenance in prostate cancer tumors lacking PTEN expression21. Genome sequencing analysis of CRPC has shown the PTEN gene is mutated in 48% of CRPC samples, and that 33% of samples have mutations in genes that interact with PTEN, resulting in 81% of samples harboring some mutation in the PTEN interaction network22. A number of PI3K and mTOR inhibitors are currently under investigation in clinical trials for CRPC including the dual inhibitor NVP-BEZ23523, and the mTOR inhibitor RAD001 or everolimus24,25. Of these compounds, NVP-BEZ235 likely has the most promise, although everolimus may be effective in subsets of patients.

Receptor tyrosine kinase (RTK) growth factor pathways

Multiple RTK growth factor signaling pathways including EGFR26,27, IGF-1R28,29, FGFR30, PDGFR31,32, and HGFR/c-MET33-35 pathways have been investigated extensively in prostate cancer progression due to their activation of the PI3K-AKT and RAS-MAPK pathways. Combined loss of PTEN with activation of the RAS-MAPK pathway can cooperate to induce epithelial to mesenchymal transition (EMT) and metastases36. Several clinical trials have been conducted targeting EGFR, including two using lapatinib, one of which was negative37, and another that was more encouraging38 with single agent activity in a subset of patients. A third trial using cetuximab showed significant progression free survival in approximately half of the treated patients39. Antibodies against IGF-1R have shown some efficacy in xenografts29 and in human trials40.

Recent epidemiological studies have determined that the generic anti-diabetes drug metformin results in lower prostate cancer specific mortality41, suggesting that not only insulin and IGF, but also glucose regulation may impact patient survival. RTK growth factor pathways can allow cells to meet bioenergetic demands of rapid proliferation by influencing cell signaling via phosphorylation. The Warburg effect has been recognized for nearly a century as perturbation of cancer cell metabolism, but it has only recently become the focus of intense investigation as an avenue for potential therapy42,43. In the Warburg effect, cancer cells undergo aerobic glycolysis rather than mitochondrial respiration and oxidative phosphorylation in order to generate the necessary precursors for protein, nucleotide, and lipid synthesis that are essential to rapid cell growth. In the reverse Warburg effect, perturbations of metabolic and catabolic metabolism can lead to a shuffling of nutrients between prostate cancer epithelial cells and surrounding stromal cells, in which pyruvate and lactate are transported to prostate tumor cells by the stroma44. Almost all glycolysis genes are overexpressed in prostate cancer, especially in advanced disease45. These data have led to calls for clinical trials employing combination therapies of RTK inhibitors or PI3K-AKT inhibitors with metformin46.

Analysis of CRPC Genomes

Although primary prostate cancers are characterized by genomic translocations that place androgen-responsive promoters such as TMPRSS2 adjacent to ETS-family transcription factors such as ERG47, ETV148, ETV549, and ETV450, metastatic prostate cancers can harbor rare gene rearrangements that generate fusions of ubiquitin conjugating enzymes with KRAS51. Integrated gene expression and copy number analysis of prostate cancers also indicates that more aggressive prostate cancers have increased copy number alterations relative to less aggressive cancers52. Moreover, PI3K-AKT was altered in 100% of metastases, while RAS-RAF signaling was altered in 90% of metastases in this study52. Exome sequencing of CRPC has identified several recurrent mutations in transcription factors and epigenetic factors that interact with AR such as FOXA1, MLL2, UTX, and ASXL122.

Epigenetic Pathways

The hypermethylation of CpG islands in gene promoters can inhibit expression of genes such as GSTP153 and has been implicated in tumorigenesis and cancer progression. Recent data have suggested that methylation of microRNA genes that target AR such as miR-34a can lead to increased AR expression, and that forced expression of miR-34a can reduce AR levels54. Since epigenetic modification of genetic DNA is a potentially reversible process, it is an enticing target in the treatment of CRPC. One such pathway target involves the inhibition of DNA methyltransferases (DNMT) to decrease the effects of methylation-related gene silencing. 5-Aza-2’-deoxycytidine (5-Aza-CdR) has been shown to increase expression of miR-146a in LNCaP cells and delay progression of tumor xenografts55, but more data is needed on drug efficacy in the face of docetaxel-treated CRPC.

The enhancer of zeste 2 (EZH2) protein is a histone methyltransferase and the catalytic subunit of the Polycomb repressive complex 2 (PRC2) that tri-methylates histone 3 lysine 27 (H3K27me3) to repress many developmental genes, especially in metastatic prostate cancer56. EZH2 is elevated in metastatic prostate cancer versus clinically localized disease or benign prostate57,58, and high expression of EZH2 in patients correlates with an increased probability of developing skeletal metastases, along with seminal vesicle and lymph node infiltration59.

Recent data from exome sequencing of CRPC have shown that there are multiple mutational alterations in the chromatin and histone modifying genes. Grasso et al. found that in pre-treated lethal metastatic CRPC tumors, there are interactions between the MLL complex family and AR, further demonstrating that there is dysregulation in epigenetic activation, commonly seen in CRPC22. Berger et al. showed that chromatin modifying genes CHD1, CHD5, and HDAC9 were mutated in 43% of sequenced Gleason 7 or higher prostate cancer tumors60. Specifically, CHD1 sequencing exhibited splice site mutations as well as intragenic breakpoints—all leading to truncated protein expression60.

Histone and chromatin-remodeling complexes are potential targets in regards to CRPC therapy, with targeting of histone deacetylases (HDACs) that facilitate AR-mediated transcription activation and repression. However, current HDAC inhibitors, such as vorinostat61,62 and panobinostat63, have shown high rates of side effects and disappointing efficacy in the treatment of docetaxel-refractory CRPC.

Nevertheless, there have been exciting developments in the development of inhibitors of bromodomain-containing proteins that recognize and bind to acetylated lysines of histone proteins. Recently, a new group of small molecules has emerged as novel inhibitors of Bromodomain Containing Protein 4 (BRD4). BRD4, along with the Mediator complex, binds at super-enhancer sites to facilitate initiation of transcription of target genes64. BRD4 inhibitors, such as JQ165, bind to bromodomains of proteins such as BRD4, and inhibit BRD4 from binding to super-enhancers of known proto-oncogenes, including MYC66,67. A mouse model of aggressive prostate cancer with simultaneous loss of PTEN and p53 tumor suppressor genes, termed RapidCaP, demonstrated highly penetrant metastases and activation of MYC68. Moreover, these castrate resistant tumors were sensitive to BRD4 inhibition using JQ1 the inhibitor68. Additional studies69 using JQ1 and the orally bioavailable BRD4 inhibitor IBET76270, found that BRD4 inhibitors disrupt AR signaling, and recruitment to and activation of downstream target genes such as the TMPRSS2-ERG gene fusion69.

EMT and SOX family genes

Typically, prostate cancer mortality is often related to metastasis to the bone, adrenal gland, liver and lung71. The epithelial to mesenchymal transition (EMT) is a major step in the metastatic process. To metastasize cancer cells need to acquire migratory and invasive capabilities, a process that involves EMT72. EMT encompasses vast molecular changes including gain of mesenchymal markers such as vimentin and N-cadherin, and loss of epithelial markers such as E-cadherin, mediated by aberrant developmental signaling pathway activation that allows epithelial cells to discard differentiated characteristics and acquire migratory and invasive capabilities typical of mesenchymal cells72. These changes include the loss of cell-cell adhesion, planar and apical-basal polarity, increased motility, and resistance to apoptosis and anoikis (cell death due to the detachment from the extracellular matrix)72,73. Among the developmental signaling pathways that are aberrantly activated during EMT is the TGF-β signaling pathway, a highly studied major inducer of EMT74. The canonical TGF-β pathway is stimulated via TGF-β induced receptor complex activation, leading to phosphorylation of SMAD 2/3. Subsequently, these SMADs form a trimer with SMAD4 and translocate to the nucleus and associate with other transcription factors to transcribe EMT-inducing genes75.

Recently, it was found that SOX4 is a master regulator of TGF-ß induced EMT via induction of EZH276. Tiwari et al demonstrated that SOX4 directly activates EZH2 expression upon TGF-β treatment and that forced expression of EZH2 can overcome SOX4 knockdown and restore TGF-β induced EMT76. Moreover, Wang et al. found that, in prostate cancer cells, SOX4 knockdown inhibited TGF-β induced EMT, while SOX4 over expression promoted adoption of the mesenchymal phenotype77. They also demonstrated that TMPRSS2-ERG is critical for TGF-β induction of SOX4 expression77. Tiwari et al. and Zhang et al. both demonstrated that ectopic expression of SOX4 could induce EMT by increasing the expression of mesenchymal markers and decreasing the expression of epithelial markers76,78. In addition, SOX4 knockdown was sufficient to cause a reversion from a mesenchymal to epithelial phenotype after a 15-day TGF-β treatment76.

Another SOX family factor, SOX9, has also been implicated in prostate cancer progression. Deletion of SOX9 in two different mouse models (TRAM and Hi-Myc) inhibited prostate cancer initiation79. ERG redirects AR to a cryptic enhancer of SOX9 to activate SOX9 expression, and knockdown of SOX9 inhibits invasion and growth of VCaP cells in vitro and in vivo80. SOX9 cooperates with PTEN deletion to drive prostate tumorigenesis81, and it activates expression of Wnt pathway components such as LRP6 and TCF482. Like SOX9, SOX4 also plays an important role in Wnt signaling via direct interaction with β-catenin83,84. SOX4 can act as an oncogene in prostate cells85, and activates expression of additional Wnt pathway components such as FZD3, FZD5, and FZD883,86.

WNT signaling

There have been many lines of evidence that suggest that Wnt signaling may be important in CRPC87. Wnt pathway genes are frequently mutated in metastatic CRPC22. WNT7B ligand is a direct transcriptional target of AR, and can induce osteoblastic bone lesions88. TMPRSS2-ERG directly activates LEF1 transcription and expression of ligands such as WNT1 and WNT3A89. Moreover, AR interacts directly with β-catenin in CRPC xenografts, whose gene expression profiles exhibit enhanced Wnt signaling90. Interestingly, a novel small compound (iCRT3) that inhibits β-catenin transcriptional activity blocks AR binding to target genes, inhibits growth of tumor xenografts, and interferes with self-renewal of bicalutamide resistant cells91.

Notch and Hedgehog

The Notch and Hedgehog pathways are also developmental pathways that likely play roles in prostate cancer progression. Notch signaling is important for prostate differentiation and maintenance of prostate stem cells92. Hedgehog signaling has also been associated with aggressive prostate cancer, and ADT alone or combined with chemotherapy induces hedgehog pathway activation93. Moreover, combined inhibition of AR and hedgehog signaling synergizes to inhibit growth of CRPC xenografts94. Docetaxel resistant CRPC cells survive via activation of Notch and Hedgehog pathways that inhibit apoptosis, and depletion of these cells results in resensitization to chemotherapy95. Thus, combination therapies that target developmental pathways along with targeting of AR and/or conventional chemotherapy may prove effective in CRPC.

lncRNA-mediated pathways

There is much data suggesting that long non-coding RNAs (lncRNA) greater than 200 base pairs in length play a critical role in tumor biology96-100. lncRNAs are typically transcribed by RNA polymerase II and are associated with epigenetic modification of histones. These lncRNAs partner with PRC1 and PRC2, leading to downstream ubiquitination and methylation activity that inhibits gene expression101. Furthermore, many lncRNAs have been linked to prostate cancer96,97,102-107. In one study, two lncRNAs, PRNCR1 and PCGEM1, were shown to bind to AR and increase ligand-dependent and ligand-independent AR-mediated gene expression, which causes downstream cellular growth signaling and cancer proliferation107. However, subsequent studies have contradicted these findings, and suggest that while PCGEM1 is associated with prostate cancer, neither of these lncRNAs interacts with AR signaling nor are prognostic in prostate cancer108. Nevertheless, other lncRNAs play important roles in prostate cancer. Pickard et al. demonstrated that transiently increased GAS5 lncRNA levels upregulated basal apoptosis in PC3 prostate cancer cells, and a similar effect was seen in 22Rv1 cells102. Prensner et al. found that SChLAP1 lncRNA overexpression, which is involved in SWI-SNF activity knockdown, was observed in 25% of prostate cancers and was an independent predictor of aggressive disease103. Other lncRNAs that have been linked to CRPC include MALAT-1104, CTBP1-AS105, and Linc00963106. Linc00963, specifically, was shown to enhance the evolution from androgen-dependence to androgen-independence through the EGFR pathway106. From these studies, it is clear that lncRNAs are intimately involved in prostate cancer biology and that dysregulation of lncRNA expression may lead to tumor suppressor antagonism as well as to castrate resistance. Future goals in lncRNA research include defining specific lncRNAs in the human genome, understanding which lncRNAs are altered in the course of tumorigenesis, and delineating the structure and binding mechanisms of lncRNAs to their targets96,98-100.

Angiogenic Pathways

Angiogenesis refers to the growth of new blood vessels from existing vasculature109. This process is an important pathway for CRPC progression, and is correlated with both increased rate of disease progression and decreased survival110. The most common target of angiogenesis is vascular endothelial growth factor (VEGF) through the use of an anti-VEGF antibody, such as bevacizumab109. VEGF can also drive EMT111, and this pathway is currently a target of several therapeutic clinical trials testing new drugs including itraconazole11, bevacizumab112, tasquinimod113, and ramucirumab114. Alternative strategies include inhibiting the binding of VEGF using VEGF-like receptors, such that VEGF is unable to interact with its normal target VEGFR receptor. VEGF ligand directed drugs have seen success in improving patient survival in other cancers, such as colorectal, breast, and non-small-cell lung cancer. However, in two prospective randomized control trials, VENICE and CALGB 90401, neither docetaxel/aflibercept nor docetaxel/bevacizumab combinations showed any significant improvement in median overall survival in patients with CRPC115,116. Docetaxel plus bevacizumab did show significant improvement in PSA decrease and progression-free-survival when compared to docetaxel alone, but this was not observed in the VENICE trial. A newer small molecule inhibitor of VEGFR2 and MET, cabozantinib, early on displayed a decrease in growth of breast, lung, and glioma tumor models while increasing apoptosis117. Later, in a prospective randomized control trial, patients with CRPC who took cabozantinib significantly increased progression-free survival when compared to placebo118. Furthermore, other trials demonstrated that cabozantinib decreased narcotic use, significantly improved sleep quality119, and inhibited prostate cancer bone growth when tested on androgen-sensitive and castration-resistant cell lines120. Other angiogenic strategies are to target the downstream activation pathways of VEGF, or to bypass the VEGF pathway completely and focus on established CRPC vessels and nutrient delivery109,110.

Biomarkers

Currently, there is a critical need for biomarkers of prostate cancer that can distinguish between locally indolent and metastatically aggressive disease121. The only FDA approved biomarker, prostate specific antigen (PSA), has historically been used for screening and detection of prostate cancer, but not for prognosis nor detection of CRPC. A major issue with PSA is that it detects prostate cells, not prostate cancer cells, and thus can result in over diagnosis and over treatment. While results of the Prostate, Lung, Colorectal, and Ovarian (PLCO) Cancer Screening Trial showed no reduction in prostate cancer specific mortality when comparing systematic annual PSA screening to opportunistic screening122, the European Randomized Study of Screening for Prostate Cancer (ERPSC) found a small reduction in mortality (1 death per 1000 men screened) in a PSA-screening naïve population123. However, several potential flaws in the ERPSC study124 and the associated harms of over diagnosis and over treatment led the US Preventive Services Task Force (USPSTF) to recommend against PSA screening125,126. Given the recent controversy in its screening effectiveness, PSA is now recommended primarily for the determination of prostate cancer progression and recurrence.

However, there are potential replacements for PSA for detection screening under development, including PCA3 and the ETS fusion gene TMPRSS2-ERG127-133. TMPRSS2-ERG has been shown to be a predictive indicator of prostate cancer development in patients who present with high-grade prostatic intraepithelial neoplasia130. FISH analyses of CRPC and metastatic prostate cancers found that TMPRSS2-ERG was detected more often in metastatic cancer, and that TMPRSS2-ERG positivity was strongly correlated to both AR and ERG expression131. TMPRSS2-ERG expression has also been seen in otherwise histologically benign radical prostatectomy as well as in cystoprostatectomy specimens. TMPRSS2-ERG and PCA3 can also be detected in patient urine post-digital rectal exam134. However, there are still conflicting data regarding the prognostic value of PCA3 or TMPRSS2-ERG. A 2011 study by Danila et al. found that the TMPRSS2-ERG fusion gene could be accurately assayed in circulating prostate tumor cells present in the blood of CRPC patients, but did not show that the presence of the fusion was a significant factor in abiraterone acetate treatment response128. Another prospective study of 322 patients illustrated that urinary PCA3 or TMPRSS2-ERG scores were not reliable in staging advanced prostate cancer.

Recent integrated bioinformatics network analyses that have included data mining and profiling of both animal models and human specimens have generated some interesting candidate biomarker gene sets for distinction of aggressive from indolent disease135 and drivers of aggressive prostate cancer136. Irshad et al. identified a panel of three genes (FGFR1, PMP22, and CDKN1A) that could predict indolence outcome for low Gleason score (Gleason 6) patients via meta-analysis of multiple studies followed by independent validation on a cohort of 95 mRNA samples and 44 fixed biopsy samples at the protein level135. Aytes et al. performed cross-species analysis of mouse and human prostate cancer gene expression patterns to identify FOXM1 and CENPF as synergistic regulators of aggressive disease136. They also found that FOXM1 and CENPF could function as prognostic biomarkers of metastasis in two independent prostate datasets137,138. Another distinct 16-gene signature of AR target genes derived from ChIP-seq of AR in CRCP tissues16 was able to accurately predict CRPC and prostate cancer recurrence in two distinct patient cohorts52,137.

Our research group at Emory University recently completed RNAseq analysis of 100 formalin-fixed paraffin embedded prostatectomy specimens, and identified a 24-gene biomarker panel that robustly predicts biochemical recurrence following surgery139. Our biomarker panel accurately predicted recurrence in an independent patient cohort52, and outperformed previously developed RNA biomarkers developed by Myriad Genetics140. Whether this set of 24 biomarker genes will also detect aggressive CRPC or discriminate aggressive from indolent disease requires further research on additional patient cohorts.

Conclusion

Recent advancements in the molecular understanding of CRPC have given us a number of potential biological targets for the treatment of CRPC. Many of the pathways and targets covered in this review currently have agents that are undergoing clinical trials, and some are FDA approved for the treatment of CRPC. Unfortunately, most of these pharmaceutical agents only moderately increase survival and CRPC still remains incurable. However, recent discoveries and avenues of research may enable more effective molecularly targeted therapies as well as a better understanding of the mechanisms of CRPC.

Figure 1. Signaling Pathways in Castration Resistant Prostate Cancer.

Figure 1

(A) TGF-β pathway is stimulated via TGF-β induced receptor complex activation, leading to phosphorylation of SMAD2/3 and subsequently form a trimer with SMAD4 and translocate to the nucleus and associate with other transcription factors to transcribe SOX4 which activates EZH2 expression leading to EMT. (B) Prostate cancer cells depend on the androgen receptor (AR) pathway and steroid hormones for continued oncogenic growth. This can occur by way of AR over expression or by AR gene amplification. Mutations in AR allow binding and activation of AR by other steroidal hormones. Wnt pathway genes are frequently mutated in CRPC and are transcriptional targets of AR. (C) Activation of the PI3K-AKT-mTOR pathway is extremely common in CRPC. Activated growth factor receptor tyrosine kinases (e.g. EGFR and IGF-1R) leads to activated PDK1, which in turn activates AKT. AKT separately phosphorylates and activates mTOR which promotes cell cycle progression, protein synthesis and decreased apoptosis. AKT can interact with AR in an androgen independent manner. (D) Activation of Receptor Tyrosine Kinase (RTK) pathway (PDGFR, HGFR/c-MET, etc) leads to proliferation through RAS-MAPK. Combined with loss of PTEN, overactive RAS-MAPK can induce EMT. (E) Intratumoral synthesis of steroidal hormones from cholesterol via upregulation of the cytochrome P450 gene CYP17A1. (F) Loss of the PTEN Tumor Suppressor promotes aberrant PI3K-AKT-mTOR signaling. (G) Aberrant Y534 phosphorylation by Src increases AR sensitivity to androgens. (H) Epigenetic Pathways: Hypermethylation of CpG islands in gene promoters inhibits expression of tumor suppressor genes or miRNAs targeting AR.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Conflict of Interest: The authors have no conflicts of interest to disclose.

References

  • 1.Siegel R, Ma J, Zou Z, Jemal A. Cancer statistics, 2014. CA: a cancer journal for clinicians. 2014;64:9–29. doi: 10.3322/caac.21208. [DOI] [PubMed] [Google Scholar]
  • 2.Wang Q, Li W, Zhang Y, et al. Androgen receptor regulates a distinct transcription program in androgen-independent prostate cancer. Cell. 2009;138:245–56. doi: 10.1016/j.cell.2009.04.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Zarour L, Alumkal J. Emerging therapies in castrate-resistant prostate cancer. Current urology reports. 2010;11:152–8. doi: 10.1007/s11934-010-0104-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Karantanos T, Corn PG, Thompson TC. Prostate cancer progression after androgen deprivation therapy: mechanisms of castrate resistance and novel therapeutic approaches. Oncogene. 2013;32:5501–11. doi: 10.1038/onc.2013.206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Enzalutamide shows strong activity in prostate cancer. Cancer discovery. 2014;4:OF2. doi: 10.1158/2159-8290.CD-NB2014-028. [DOI] [PubMed] [Google Scholar]
  • 6.Agarwal N, Di Lorenzo G, Sonpavde G, Bellmunt J. New agents for prostate cancer. Annals of oncology : official journal of the European Society for Medical Oncology / ESMO. 2014 doi: 10.1093/annonc/mdu038. [DOI] [PubMed] [Google Scholar]
  • 7.Ezzell EE, Chang KS, George BJ. New agents in the arsenal to fight castrate-resistant prostate cancer. Current oncology reports. 2013;15:239–48. doi: 10.1007/s11912-013-0305-9. [DOI] [PubMed] [Google Scholar]
  • 8.Saad F, Miller K. Treatment options in castration-resistant prostate cancer: current therapies and emerging docetaxel-based regimens. Urologic oncology. 2014;32:70–9. doi: 10.1016/j.urolonc.2013.01.005. [DOI] [PubMed] [Google Scholar]
  • 9.Beltran H, Beer TM, Carducci MA, et al. New therapies for castration-resistant prostate cancer: efficacy and safety. European urology. 2011;60:279–90. doi: 10.1016/j.eururo.2011.04.038. [DOI] [PubMed] [Google Scholar]
  • 10.Dayyani F, Gallick GE, Logothetis CJ, Corn PG. Novel therapies for metastatic castrate- resistant prostate cancer. Journal of the National Cancer Institute. 2011;103:1665–75. doi: 10.1093/jnci/djr362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Antonarakis ES, Carducci MA. Future directions in castrate-resistant prostate cancer therapy. Clinical genitourinary cancer. 2010;8:37–46. doi: 10.3816/CGC.2010.n.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Cai C, He HH, Chen S, et al. Androgen receptor gene expression in prostate cancer is directly suppressed by the androgen receptor through recruitment of lysine-specific demethylase 1. Cancer cell. 2011;20:457–71. doi: 10.1016/j.ccr.2011.09.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Yuan X, Cai C, Chen S, Chen S, Yu Z, Balk SP. Androgen receptor functions in castration- resistant prostate cancer and mechanisms of resistance to new agents targeting the androgen axis. Oncogene. 2013 doi: 10.1038/onc.2013.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Taplin ME, Bubley GJ, Ko YJ, et al. Selection for androgen receptor mutations in prostate cancers treated with androgen antagonist. Cancer research. 1999;59:2511–5. [PubMed] [Google Scholar]
  • 15.Guo Z, Dai B, Jiang T, et al. Regulation of androgen receptor activity by tyrosine phosphorylation. Cancer cell. 2006;10:309–19. doi: 10.1016/j.ccr.2006.08.021. [DOI] [PubMed] [Google Scholar]
  • 16.Sharma NL, Massie CE, Ramos-Montoya A, et al. The androgen receptor induces a distinct transcriptional program in castration-resistant prostate cancer in man. Cancer cell. 2013;23:35–47. doi: 10.1016/j.ccr.2012.11.010. [DOI] [PubMed] [Google Scholar]
  • 17.Hsieh AC, Liu Y, Edlind MP, et al. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature. 2012;485:55–61. doi: 10.1038/nature10912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Brenner JC, Chinnaiyan AM. Disruptive events in the life of prostate cancer. Cancer cell. 2011;19:301–3. doi: 10.1016/j.ccr.2011.02.020. [DOI] [PubMed] [Google Scholar]
  • 19.Carver BS, Chapinski C, Wongvipat J, et al. Reciprocal feedback regulation of PI3K and androgen receptor signaling in PTEN-deficient prostate cancer. Cancer cell. 2011;19:575–86. doi: 10.1016/j.ccr.2011.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Gonzalez-Billalabeitia E, Seitzer N, Song SJ, et al. Vulnerabilities of PTEN-p53-deficient prostate cancers to compound PARP/PI3K inhibition. Cancer discovery. 2014 doi: 10.1158/2159-8290.CD-13-0230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Chin YM, Yuan X, Balk SP, Toker A. Pten-deficient tumors depend on akt2 for maintenance and survival. Cancer discovery. 2014 doi: 10.1158/2159-8290.CD-13-0873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Grasso CS, Wu YM, Robinson DR, et al. The mutational landscape of lethal castration-resistant prostate cancer. Nature. 2012;487:239–43. doi: 10.1038/nature11125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Hong SW, Shin JS, Moon JH, et al. NVP-BEZ235, a dual PI3K/mTOR inhibitor, induces cell death through alternate routes in prostate cancer cells depending on the PTEN genotype. Apoptosis : an international journal on programmed cell death. 2014;19:895–904. doi: 10.1007/s10495-014-0973-4. [DOI] [PubMed] [Google Scholar]
  • 24.Nakabayashi M, Werner L, Courtney KD, et al. Phase II trial of RAD001 and bicalutamide for castration-resistant prostate cancer. BJU international. 2012;110:1729–35. doi: 10.1111/j.1464-410X.2012.11456.x. [DOI] [PubMed] [Google Scholar]
  • 25.Templeton AJ, Dutoit V, Cathomas R, et al. Phase 2 trial of single-agent everolimus in chemotherapy-naive patients with castration-resistant prostate cancer (SAKK 08/08). European urology. 2013;64:150–8. doi: 10.1016/j.eururo.2013.03.040. [DOI] [PubMed] [Google Scholar]
  • 26.de Muga S, Hernandez S, Agell L, et al. Molecular alterations of EGFR and PTEN in prostate cancer: association with high-grade and advanced-stage carcinomas. Modern pathology : an official journal of the United States and Canadian Academy of Pathology, Inc. 2010;23:703–12. doi: 10.1038/modpathol.2010.45. [DOI] [PubMed] [Google Scholar]
  • 27.Mimeault M, Johansson SL, Batra SK. Pathobiological implications of the expression of EGFR, pAkt, NF-kappaB and MIC-1 in prostate cancer stem cells and their progenies. PloS one. 2012;7:e31919. doi: 10.1371/journal.pone.0031919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Aggarwal RR, Ryan CJ, Chan JM. Insulin-like growth factor pathway: a link between androgen deprivation therapy (ADT), insulin resistance, and disease progression in patients with prostate cancer? Urologic oncology. 2013;31:522–30. doi: 10.1016/j.urolonc.2011.05.001. [DOI] [PubMed] [Google Scholar]
  • 29.Fahrenholtz CD, Beltran PJ, Burnstein KL. Targeting IGF-IR with ganitumab inhibits tumorigenesis and increases durability of response to androgen-deprivation therapy in VCaP prostate cancer xenografts. Molecular cancer therapeutics. 2013;12:394–404. doi: 10.1158/1535-7163.MCT-12-0648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Armstrong K, Ahmad I, Kalna G, et al. Upregulated FGFR1 expression is associated with the transition of hormone-naive to castrate-resistant prostate cancer. British journal of cancer. 2011;105:1362–9. doi: 10.1038/bjc.2011.367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Mathew P, Thall PF, Bucana CD, et al. Platelet-derived growth factor receptor inhibition and chemotherapy for castration-resistant prostate cancer with bone metastases. Clinical cancer research : an official journal of the American Association for Cancer Research. 2007;13:5816–24. doi: 10.1158/1078-0432.CCR-07-1269. [DOI] [PubMed] [Google Scholar]
  • 32.Najy AJ, Jung YS, Won JJ, et al. Cediranib inhibits both the intraosseous growth of PDGF D-positive prostate cancer cells and the associated bone reaction. The Prostate. 2012;72:1328–38. doi: 10.1002/pros.22481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Knudsen BS, Gmyrek GA, Inra J, et al. High expression of the Met receptor in prostate cancer metastasis to bone. Urology. 2002;60:1113–7. doi: 10.1016/s0090-4295(02)01954-4. [DOI] [PubMed] [Google Scholar]
  • 34.Lee RJ, Smith MR. Targeting MET and vascular endothelial growth factor receptor signaling in castration-resistant prostate cancer. Cancer journal. 2013;19:90–8. doi: 10.1097/PPO.0b013e318281e280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Varkaris A, Corn PG, Gaur S, Dayyani F, Logothetis CJ, Gallick GE. The role of HGF/c-Met signaling in prostate cancer progression and c-Met inhibitors in clinical trials. Expert opinion on investigational drugs. 2011;20:1677–84. doi: 10.1517/13543784.2011.631523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Mulholland DJ, Kobayashi N, Ruscetti M, et al. Pten loss and RAS/MAPK activation cooperate to promote EMT and metastasis initiated from prostate cancer stem/progenitor cells. Cancer research. 2012;72:1878–89. doi: 10.1158/0008-5472.CAN-11-3132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Sridhar SS, Hotte SJ, Chin JL, et al. A multicenter phase II clinical trial of lapatinib (GW572016) in hormonally untreated advanced prostate cancer. American journal of clinical oncology. 2010;33:609–13. doi: 10.1097/COC.0b013e3181beac33. [DOI] [PubMed] [Google Scholar]
  • 38.Whang YE, Armstrong AJ, Rathmell WK, et al. A phase II study of lapatinib, a dual EGFR and HER-2 tyrosine kinase inhibitor, in patients with castration-resistant prostate cancer. Urologic oncology. 2013;31:82–6. doi: 10.1016/j.urolonc.2010.09.018. [DOI] [PubMed] [Google Scholar]
  • 39.Cathomas R, Rothermundt C, Klingbiel D, et al. Efficacy of cetuximab in metastatic castration-resistant prostate cancer might depend on EGFR and PTEN expression: results from a phase II trial (SAKK 08/07). Clinical cancer research : an official journal of the American Association for Cancer Research. 2012;18:6049–57. doi: 10.1158/1078-0432.CCR-12-2219. [DOI] [PubMed] [Google Scholar]
  • 40.de Bono JS, Piulats JM, Pandha HS, et al. Phase II randomized study of figitumumab plus docetaxel and docetaxel alone with crossover for metastatic castration-resistant prostate cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2014;20:1925–34. doi: 10.1158/1078-0432.CCR-13-1869. [DOI] [PubMed] [Google Scholar]
  • 41.Margel D, Urbach DR, Lipscombe LL, et al. Metformin use and all-cause and prostate cancer-specific mortality among men with diabetes. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2013;31:3069–75. doi: 10.1200/JCO.2012.46.7043. [DOI] [PubMed] [Google Scholar]
  • 42.Cairns RA, Harris IS, Mak TW. Regulation of cancer cell metabolism. Nature reviews Cancer. 2011;11:85–95. doi: 10.1038/nrc2981. [DOI] [PubMed] [Google Scholar]
  • 43.Koppenol WH, Bounds PL, Dang CV. Otto Warburg's contributions to current concepts of cancer metabolism. Nature reviews Cancer. 2011;11:325–37. doi: 10.1038/nrc3038. [DOI] [PubMed] [Google Scholar]
  • 44.Giatromanolaki A, Koukourakis MI, Koutsopoulos A, Mendrinos S, Sivridis E. The metabolic interactions between tumor cells and tumor-associated stroma (TAS) in prostatic cancer. Cancer biology & therapy. 2012;13:1284–9. doi: 10.4161/cbt.21785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Altenberg B, Greulich KO. Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics. 2004;84:1014–20. doi: 10.1016/j.ygeno.2004.08.010. [DOI] [PubMed] [Google Scholar]
  • 46.Quinn BJ, Kitagawa H, Memmott RM, Gills JJ, Dennis PA. Repositioning metformin for cancer prevention and treatment. Trends in endocrinology and metabolism: TEM. 2013;24:469–80. doi: 10.1016/j.tem.2013.05.004. [DOI] [PubMed] [Google Scholar]
  • 47.Tomlins SA, Rhodes DR, Perner S, et al. Recurrent fusion of TMPRSS2 and ETS transcription factor genes in prostate cancer. Science. 2005;310:644–8. doi: 10.1126/science.1117679. [DOI] [PubMed] [Google Scholar]
  • 48.Tomlins SA, Laxman B, Dhanasekaran SM, et al. Distinct classes of chromosomal rearrangements create oncogenic ETS gene fusions in prostate cancer. Nature. 2007;448:595–9. doi: 10.1038/nature06024. [DOI] [PubMed] [Google Scholar]
  • 49.Helgeson BE, Tomlins SA, Shah N, et al. Characterization of TMPRSS2:ETV5 and SLC45A3:ETV5 gene fusions in prostate cancer. Cancer research. 2008;68:73–80. doi: 10.1158/0008-5472.CAN-07-5352. [DOI] [PubMed] [Google Scholar]
  • 50.Tomlins SA, Mehra R, Rhodes DR, et al. TMPRSS2:ETV4 gene fusions define a third molecular subtype of prostate cancer. Cancer research. 2006;66:3396–400. doi: 10.1158/0008-5472.CAN-06-0168. [DOI] [PubMed] [Google Scholar]
  • 51.Wang XS, Shankar S, Dhanasekaran SM, et al. Characterization of KRAS rearrangements in metastatic prostate cancer. Cancer discovery. 2011;1:35–43. doi: 10.1158/2159-8274.CD-10-0022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Taylor BS, Schultz N, Hieronymus H, et al. Integrative genomic profiling of human prostate cancer. Cancer cell. 2010;18:11–22. doi: 10.1016/j.ccr.2010.05.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Lee WH, Isaacs WB, Bova GS, Nelson WG. CG island methylation changes near the GSTP1 gene in prostatic carcinoma cells detected using the polymerase chain reaction: a new prostate cancer biomarker. Cancer epidemiology, biomarkers & prevention : a publication of the American Association for Cancer Research, cosponsored by the American Society of Preventive Oncology. 1997;6:443–50. [PubMed] [Google Scholar]
  • 54.Kong D, Heath E, Chen W, et al. Epigenetic silencing of miR-34a in human prostate cancer cells and tumor tissue specimens can be reversed by BR-DIM treatment. American journal of translational research. 2012;4:14–23. [PMC free article] [PubMed] [Google Scholar]
  • 55.Wang X, Gao H, Ren L, Gu J, Zhang Y, Zhang Y. Demethylation of the miR-146a promoter by 5-Aza-2′-deoxycytidine correlates with delayed progression of castration-resistant prostate cancer. BMC cancer. 2014;14:308. doi: 10.1186/1471-2407-14-308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Kirmizis A, Bartley SM, Kuzmichev A, et al. Silencing of human polycomb target genes is associated with methylation of histone H3 Lys 27. Genes Dev. 2004;18:1592–605. doi: 10.1101/gad.1200204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Varambally S, Dhanasekaran SM, Zhou M, et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature. 2002;419:624–9. doi: 10.1038/nature01075. [DOI] [PubMed] [Google Scholar]
  • 58.Varambally S, Yu J, Laxman B, et al. Integrative genomic and proteomic analysis of prostate cancer reveals signatures of metastatic progression. Cancer cell. 2005;8:393–406. doi: 10.1016/j.ccr.2005.10.001. [DOI] [PubMed] [Google Scholar]
  • 59.Bachmann IM, Halvorsen OJ, Collett K, et al. EZH2 expression is associated with high proliferation rate and aggressive tumor subgroups in cutaneous melanoma and cancers of the endometrium, prostate, and breast. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2006;24:268–73. doi: 10.1200/JCO.2005.01.5180. [DOI] [PubMed] [Google Scholar]
  • 60.Berger MF, Lawrence MS, Demichelis F, et al. The genomic complexity of primary human prostate cancer. Nature. 2011;470:214–20. doi: 10.1038/nature09744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Bradley D, Rathkopf D, Dunn R, et al. Vorinostat in advanced prostate cancer patients progressing on prior chemotherapy (National Cancer Institute Trial 6862): trial results and interleukin-6 analysis: a study by the Department of Defense Prostate Cancer Clinical Trial Consortium and University of Chicago Phase 2 Consortium. Cancer. 2009;115:5541–9. doi: 10.1002/cncr.24597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Schneider BJ, Kalemkerian GP, Bradley D, et al. Phase I study of vorinostat (suberoylanilide hydroxamic acid, NSC 701852) in combination with docetaxel in patients with advanced and relapsed solid malignancies. Investigational new drugs. 2012;30:249–57. doi: 10.1007/s10637-010-9503-6. [DOI] [PubMed] [Google Scholar]
  • 63.Rathkopf DE, Picus J, Hussain A, et al. A phase 2 study of intravenous panobinostat in patients with castration-resistant prostate cancer. Cancer chemotherapy and pharmacology. 2013;72:537–44. doi: 10.1007/s00280-013-2224-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Whyte WA, Orlando DA, Hnisz D, et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell. 2013;153:307–19. doi: 10.1016/j.cell.2013.03.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Filippakopoulos P, Qi J, Picaud S, et al. Selective inhibition of BET bromodomains. Nature. 2010;468:1067–73. doi: 10.1038/nature09504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Chapuy B, McKeown MR, Lin CY, et al. Discovery and characterization of super-enhancer-associated dependencies in diffuse large B cell lymphoma. Cancer cell. 2013;24:777–90. doi: 10.1016/j.ccr.2013.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Loven J, Hoke HA, Lin CY, et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell. 2013;153:320–34. doi: 10.1016/j.cell.2013.03.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Cho H, Herzka T, Zheng W, et al. RapidCaP, a novel GEM model for metastatic prostate cancer analysis and therapy, reveals myc as a driver of Pten-mutant metastasis. Cancer discovery. 2014;4:318–33. doi: 10.1158/2159-8290.CD-13-0346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Asangani IA, Dommeti VL, Wang X, et al. Therapeutic targeting of BET bromodomain proteins in castration-resistant prostate cancer. Nature. 2014 doi: 10.1038/nature13229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Wyce A, Ganji G, Smitheman KN, et al. BET inhibition silences expression of MYCN and BCL2 and induces cytotoxicity in neuroblastoma tumor models. PloS one. 2013;8:e72967. doi: 10.1371/journal.pone.0072967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Bubendorf L, Schopfer A, Wagner U, et al. Metastatic patterns of prostate cancer: an autopsy study of 1,589 patients. Human pathology. 2000;31:578–83. doi: 10.1053/hp.2000.6698. [DOI] [PubMed] [Google Scholar]
  • 72.Polyak K, Weinberg RA. Transitions between epithelial and mesenchymal states: acquisition of malignant and stem cell traits. Nature reviews Cancer. 2009;9:265–73. doi: 10.1038/nrc2620. [DOI] [PubMed] [Google Scholar]
  • 73.Zhu Z, Sanchez-Sweatman O, Huang X, et al. Anoikis and metastatic potential of cloudman S91 melanoma cells. Cancer research. 2001;61:1707–16. [PubMed] [Google Scholar]
  • 74.Xu J, Lamouille S, Derynck R. TGF-beta-induced epithelial to mesenchymal transition. Cell research. 2009;19:156–72. doi: 10.1038/cr.2009.5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Valcourt U, Kowanetz M, Niimi H, Heldin CH, Moustakas A. TGF-beta and the Smad signaling pathway support transcriptomic reprogramming during epithelial-mesenchymal cell transition. Molecular biology of the cell. 2005;16:1987–2002. doi: 10.1091/mbc.E04-08-0658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Tiwari N, Tiwari VK, Waldmeier L, et al. Sox4 is a master regulator of epithelial mesenchymal transition by controlling ezh2 expression and epigenetic reprogramming. Cancer cell. 2013;23:768–83. doi: 10.1016/j.ccr.2013.04.020. [DOI] [PubMed] [Google Scholar]
  • 77.Wang L, Li Y, Yang X, et al. ERG-SOX4 interaction promotes epithelial-mesenchymal transition in prostate cancer cells. The Prostate. 2014;74:647–58. doi: 10.1002/pros.22783. [DOI] [PubMed] [Google Scholar]
  • 78.Zhang J, Liang Q, Lei Y, et al. SOX4 induces epithelial-mesenchymal transition and contributes to breast cancer progression. Cancer research. 2012;72:4597–608. doi: 10.1158/0008-5472.CAN-12-1045. [DOI] [PubMed] [Google Scholar]
  • 79.Huang Z, Hurley PJ, Simons BW, et al. Sox9 is required for prostate development and prostate cancer initiation. Oncotarget. 2012;3:651–63. doi: 10.18632/oncotarget.531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Cai C, Wang H, He HH, et al. ERG induces androgen receptor-mediated regulation of SOX9 in prostate cancer. The Journal of clinical investigation. 2013;123:1109–22. doi: 10.1172/JCI66666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Thomsen MK, Ambroisine L, Wynn S, et al. SOX9 elevation in the prostate promotes proliferation and cooperates with PTEN loss to drive tumor formation. Cancer research. 2010;70:979–87. doi: 10.1158/0008-5472.CAN-09-2370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Wang H, He L, Ma F, et al. SOX9 regulates low density lipoprotein receptor-related protein 6 (LRP6) and T-cell factor 4 (TCF4) expression and Wnt/beta-catenin activation in breast cancer. The Journal of biological chemistry. 2013;288:6478–87. doi: 10.1074/jbc.M112.419184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Scharer CD, McCabe CD, Ali-Seyed M, Berger MF, Bulyk ML, Moreno CS. Genome-wide promoter analysis of the SOX4 transcriptional network in prostate cancer cells. Cancer research. 2009;69:709–17. doi: 10.1158/0008-5472.CAN-08-3415. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Sinner D, Kordich JJ, Spence JR, et al. Sox17 and Sox4 differentially regulate beta-catenin/T-cell factor activity and proliferation of colon carcinoma cells. Mol Cell Biol. 2007;27:7802–15. doi: 10.1128/MCB.02179-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Liu P, Ramachandran S, Ali Seyed M, et al. Sex-determining region Y box 4 is a transforming oncogene in human prostate cancer cells. Cancer research. 2006;66:4011–9. doi: 10.1158/0008-5472.CAN-05-3055. [DOI] [PubMed] [Google Scholar]
  • 86.Moreno CS. The Sex-determining region Y-box 4 and homeobox C6 transcriptional networks in prostate cancer progression: crosstalk with the Wnt, Notch, and PI3K pathways. The American journal of pathology. 2010;176:518–27. doi: 10.2353/ajpath.2010.090657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Kypta RM, Waxman J. Wnt/beta-catenin signalling in prostate cancer. Nature reviews Urology. 2012 doi: 10.1038/nrurol.2012.116. [DOI] [PubMed] [Google Scholar]
  • 88.Zheng D, Decker KF, Zhou T, et al. Role of WNT7B-induced noncanonical pathway in advanced prostate cancer. Molecular cancer research : MCR. 2013;11:482–93. doi: 10.1158/1541-7786.MCR-12-0520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Wu L, Zhao JC, Kim J, Jin HJ, Wang CY, Yu J. ERG is a critical regulator of Wnt/LEF1 signaling in prostate cancer. Cancer research. 2013;73:6068–79. doi: 10.1158/0008-5472.CAN-13-0882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Wang G, Wang J, Sadar MD. Crosstalk between the androgen receptor and beta-catenin in castrate-resistant prostate cancer. Cancer research. 2008;68:9918–27. doi: 10.1158/0008-5472.CAN-08-1718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Lee E, Madar A, David G, Garabedian MJ, Dasgupta R, Logan SK. Inhibition of androgen receptor and beta-catenin activity in prostate cancer. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:15710–5. doi: 10.1073/pnas.1218168110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Carvalho FL, Simons BW, Eberhart CG, Berman DM. Notch signaling in prostate cancer: A moving target. The Prostate. 2014;74:933–45. doi: 10.1002/pros.22811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Efstathiou E, Karlou M, Wen S, et al. Integrated Hedgehog signaling is induced following castration in human and murine prostate cancers. The Prostate. 2013;73:153–61. doi: 10.1002/pros.22550. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Gowda PS, Deng JD, Mishra S, et al. Inhibition of hedgehog and androgen receptor signaling pathways produced synergistic suppression of castration-resistant prostate cancer progression. Molecular cancer research : MCR. 2013;11:1448–61. doi: 10.1158/1541-7786.MCR-13-0278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Domingo-Domenech J, Vidal SJ, Rodriguez-Bravo V, et al. Suppression of acquired docetaxel resistance in prostate cancer through depletion of notch-and hedgehog-dependent tumor-initiating cells. Cancer cell. 2012;22:373–88. doi: 10.1016/j.ccr.2012.07.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Martens-Uzunova ES, Bottcher R, Croce CM, Jenster G, Visakorpi T, Calin GA. Long noncoding RNA in prostate, bladder, and kidney cancer. European urology. 2014;65:1140–51. doi: 10.1016/j.eururo.2013.12.003. [DOI] [PubMed] [Google Scholar]
  • 97.Crea F, Watahiki A, Quagliata L, et al. Identification of a long non-coding RNA as a novel biomarker and potential therapeutic target for metastatic prostate cancer. Oncotarget. 2014;5:764–74. doi: 10.18632/oncotarget.1769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Prensner JR, Chinnaiyan AM. The emergence of lncRNAs in cancer biology. Cancer discovery. 2011;1:391–407. doi: 10.1158/2159-8290.CD-11-0209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Cech TR, Steitz JA. The noncoding RNA revolution-trashing old rules to forge new ones. Cell. 2014;157:77–94. doi: 10.1016/j.cell.2014.03.008. [DOI] [PubMed] [Google Scholar]
  • 100.Geisler S, Coller J. RNA in unexpected places: long non-coding RNA functions in diverse cellular contexts. Nature reviews Molecular cell biology. 2013;14:699–712. doi: 10.1038/nrm3679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Aguilo F, Zhou MM, Walsh MJ. Long noncoding RNA, polycomb, and the ghosts haunting INK4b-ARF-INK4a expression. Cancer research. 2011;71:5365–9. doi: 10.1158/0008-5472.CAN-10-4379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Pickard MR, Mourtada-Maarabouni M, Williams GT. Long non-coding RNA GAS5 regulates apoptosis in prostate cancer cell lines. Biochimica et biophysica acta. 2013;1832:1613–23. doi: 10.1016/j.bbadis.2013.05.005. [DOI] [PubMed] [Google Scholar]
  • 103.Prensner JR, Iyer MK, Sahu A, et al. The long noncoding RNA SChLAP1 promotes aggressive prostate cancer and antagonizes the SWI/SNF complex. Nature genetics. 2013;45:1392–8. doi: 10.1038/ng.2771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Ren S, Liu Y, Xu W, et al. Long noncoding RNA MALAT-1 is a new potential therapeutic target for castration resistant prostate cancer. The Journal of urology. 2013;190:2278–87. doi: 10.1016/j.juro.2013.07.001. [DOI] [PubMed] [Google Scholar]
  • 105.Takayama K, Horie-Inoue K, Katayama S, et al. Androgen-responsive long noncoding RNA CTBP1-AS promotes prostate cancer. The EMBO journal. 2013;32:1665–80. doi: 10.1038/emboj.2013.99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Wang L, Han S, Jin G, et al. Linc00963: a novel, long non-coding RNA involved in the transition of prostate cancer from androgen-dependence to androgen-independence. International journal of oncology. 2014;44:2041–9. doi: 10.3892/ijo.2014.2363. [DOI] [PubMed] [Google Scholar]
  • 107.Yang L, Lin C, Jin C, et al. lncRNA-dependent mechanisms of androgen-receptor-regulated gene activation programs. Nature. 2013;500:598–602. doi: 10.1038/nature12451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Prensner JR, Sahu A, Iyer MK, et al. The IncRNAs PCGEM1 and PRNCR1 are not implicated in castration resistant prostate cancer. Oncotarget. 2014;5:1434–8. doi: 10.18632/oncotarget.1846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Adair TH, Montani JP. Angiogenesis. San Rafael (CA): 2010. [Google Scholar]
  • 110.Karagiannis GS, Saraon P, Jarvi KA, Diamandis EP. Proteomic signatures of angiogenesis in androgen-independent prostate cancer. The Prostate. 2014;74:260–72. doi: 10.1002/pros.22747. [DOI] [PubMed] [Google Scholar]
  • 111.Mak P, Leav I, Pursell B, et al. ERbeta impedes prostate cancer EMT by destabilizing HIF-1alpha and inhibiting VEGF-mediated snail nuclear localization: implications for Gleason grading. Cancer cell. 2010;17:319–32. doi: 10.1016/j.ccr.2010.02.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Small AC, Oh WK. Bevacizumab treatment of prostate cancer. Expert opinion on biological therapy. 2012;12:1241–9. doi: 10.1517/14712598.2012.704015. [DOI] [PubMed] [Google Scholar]
  • 113.Olsson A, Bjork A, Vallon-Christersson J, Isaacs JT, Leanderson T. Tasquinimod (ABR-215050), a quinoline-3-carboxamide anti-angiogenic agent, modulates the expression of thrombospondin-1 in human prostate tumors. Molecular cancer. 2010;9:107. doi: 10.1186/1476-4598-9-107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Krupitskaya Y, Wakelee HA. Ramucirumab, a fully human mAb to the transmembrane signaling tyrosine kinase VEGFR-2 for the potential treatment of cancer. Current opinion in investigational drugs. 2009;10:597–605. [PubMed] [Google Scholar]
  • 115.Kelly WK, Halabi S, Carducci M, et al. Randomized, double-blind, placebo-controlled phase III trial comparing docetaxel and prednisone with or without bevacizumab in men with metastatic castration-resistant prostate cancer: CALGB 90401. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2012;30:1534–40. doi: 10.1200/JCO.2011.39.4767. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Tannock IF, Fizazi K, Ivanov S, et al. Aflibercept versus placebo in combination with docetaxel and prednisone for treatment of men with metastatic castration-resistant prostate cancer (VENICE): a phase 3, double-blind randomised trial. Lancet Oncol. 2013;14:760–8. doi: 10.1016/S1470-2045(13)70184-0. [DOI] [PubMed] [Google Scholar]
  • 117.Yakes FM, Chen J, Tan J, et al. Cabozantinib (XL184), a novel MET and VEGFR2 inhibitor, simultaneously suppresses metastasis, angiogenesis, and tumor growth. Molecular cancer therapeutics. 2011;10:2298–308. doi: 10.1158/1535-7163.MCT-11-0264. [DOI] [PubMed] [Google Scholar]
  • 118.Smith DC, Smith MR, Sweeney C, et al. Cabozantinib in patients with advanced prostate cancer: results of a phase II randomized discontinuation trial. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2013;31:412–9. doi: 10.1200/JCO.2012.45.0494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Basch E, Autio KA, Smith MR, et al. Effects of Cabozantinib on Pain and Narcotic Use in Patients with Castration-resistant Prostate Cancer: Results from a Phase 2 Nonrandomized Expansion Cohort. European urology. 2014 doi: 10.1016/j.eururo.2014.02.013. [DOI] [PubMed] [Google Scholar]
  • 120.Nguyen HM, Ruppender N, Zhang X, et al. Cabozantinib inhibits growth of androgen-sensitive and castration-resistant prostate cancer and affects bone remodeling. PloS one. 2013;8:e78881. doi: 10.1371/journal.pone.0078881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Crawford ED, Stone NN, Yu EY, et al. Challenges and recommendations for early identification of metastatic disease in prostate cancer. Urology. 2014;83:664–9. doi: 10.1016/j.urology.2013.10.026. [DOI] [PubMed] [Google Scholar]
  • 122.Andriole GL, Crawford ED, Grubb RL, 3rd, et al. Prostate cancer screening in the randomized Prostate, Lung, Colorectal, and Ovarian Cancer Screening Trial: mortality results after 13 years of follow-up. Journal of the National Cancer Institute. 2012;104:125–32. doi: 10.1093/jnci/djr500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Schroder FH, Hugosson J, Roobol MJ, et al. Prostate-cancer mortality at 11 years of follow-up. The New England journal of medicine. 2012;366:981–90. doi: 10.1056/NEJMoa1113135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Miller AB. New data on prostate-cancer mortality after PSA screening. The New England journal of medicine. 2012;366:1047–8. doi: 10.1056/NEJMe1200185. [DOI] [PubMed] [Google Scholar]
  • 125.Chou R, Croswell JM, Dana T, et al. Screening for prostate cancer: a review of the evidence for the U.S. Preventive Services Task Force. Annals of internal medicine. 2011;155:762–71. doi: 10.7326/0003-4819-155-11-201112060-00375. [DOI] [PubMed] [Google Scholar]
  • 126.Chou R, LeFevre ML. Prostate cancer screening--the evidence, the recommendations, and the clinical implications. JAMA : the journal of the American Medical Association. 2011;306:2721–2. doi: 10.1001/jama.2011.1891. [DOI] [PubMed] [Google Scholar]
  • 127.Leyten GH, Hessels D, Jannink SA, et al. Prospective multicentre evaluation of PCA3 and TMPRSS2-ERG gene fusions as diagnostic and prognostic urinary biomarkers for prostate cancer. European urology. 2014;65:534–42. doi: 10.1016/j.eururo.2012.11.014. [DOI] [PubMed] [Google Scholar]
  • 128.Danila DC, Anand A, Sung CC, et al. TMPRSS2-ERG status in circulating tumor cells as a predictive biomarker of sensitivity in castration-resistant prostate cancer patients treated with abiraterone acetate. European urology. 2011;60:897–904. doi: 10.1016/j.eururo.2011.07.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Martinez-Pineiro L, Schalken JA, Cabri P, Maisonobe P, de la Taille A, Triptocare Study G Evaluation of urinary prostate cancer antigen-3 (PCA3) and TMPRSS2-ERG score changes when starting androgen-deprivation therapy with triptorelin 6-month formulation in patients with locally advanced and metastatic prostate cancer. BJU international. 2013 doi: 10.1111/bju.12542. [DOI] [PubMed] [Google Scholar]
  • 130.Park K, Dalton JT, Narayanan R, et al. TMPRSS2:ERG gene fusion predicts subsequent detection of prostate cancer in patients with high-grade prostatic intraepithelial neoplasia. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2014;32:206–11. doi: 10.1200/JCO.2013.49.8386. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Qu X, Randhawa G, Friedman C, et al. A three-marker FISH panel detects more genetic aberrations of AR, PTEN and TMPRSS2/ERG in castration-resistant or metastatic prostate cancers than in primary prostate tumors. PloS one. 2013;8:e74671. doi: 10.1371/journal.pone.0074671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Vaananen RM, Lilja H, Kauko L, et al. Cancer-associated changes in the expression of TMPRSS2-ERG, PCA3, and SPINK1 in histologically benign tissue from cancerous vs noncancerous prostatectomy specimens. Urology. 2014;83:511, e1–7. doi: 10.1016/j.urology.2013.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Yao Y, Wang H, Li B, Tang Y. Evaluation of the TMPRSS2:ERG fusion for the detection of prostate cancer: a systematic review and meta-analysis. Tumour biology : the journal of the International Society for Oncodevelopmental Biology and Medicine. 2014;35:2157–66. doi: 10.1007/s13277-013-1286-x. [DOI] [PubMed] [Google Scholar]
  • 134.Dijkstra S, Birker IL, Smit FP, et al. Prostate cancer biomarker profiles in urinary sediments and exosomes. The Journal of urology. 2014;191:1132–8. doi: 10.1016/j.juro.2013.11.001. [DOI] [PubMed] [Google Scholar]
  • 135.Irshad S, Bansal M, Castillo-Martin M, et al. A molecular signature predictive of indolent prostate cancer. Science translational medicine. 2013;5:202ra122. doi: 10.1126/scitranslmed.3006408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Aytes A, Mitrofanova A, Lefebvre C, et al. Cross-Species Regulatory Network Analysis Identifies a Synergistic Interaction between FOXM1 and CENPF that Drives Prostate Cancer Malignancy. Cancer cell. 2014;25:638–51. doi: 10.1016/j.ccr.2014.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Glinsky GV, Glinskii AB, Stephenson AJ, Hoffman RM, Gerald WL. Gene expression profiling predicts clinical outcome of prostate cancer. The Journal of clinical investigation. 2004;113:913–23. doi: 10.1172/JCI20032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Sboner A, Demichelis F, Calza S, et al. Molecular sampling of prostate cancer: a dilemma for predicting disease progression. BMC medical genomics. 2010;3:8. doi: 10.1186/1755-8794-3-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Long Q, Xu J, Osunkoya AO, et al. Global Transcriptome Analysis of Formalin-Fixed Prostate Cancer Specimens Identifies Biomarkers of Disease Recurrence. Cancer research. 2014;74:3228–37. doi: 10.1158/0008-5472.CAN-13-2699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Cuzick J, Swanson GP, Fisher G, et al. Prognostic value of an RNA expression signature derived from cell cycle proliferation genes in patients with prostate cancer: a retrospective study. Lancet Oncol. 2011;12:245–55. doi: 10.1016/S1470-2045(10)70295-3. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES