Next Article in Journal
Trends in Anti-Tumor Effects of Pseudomonas aeruginosa Mannose-Sensitive-Hemagglutinin (PA-MSHA): An Overview of Positive and Negative Effects
Previous Article in Journal
The Importance of Th2 Immune Responses in Mediating the Progression of Gastritis-Associated Metaplasia to Gastric Cancer
Previous Article in Special Issue
ID2 Promotes Lineage Transition of Prostate Cancer through FGFR and JAK-STAT Signaling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms of Prostate Cancer Development in the Precision Medicine Era: A Comprehensive Review

Department of Urology, Iwate Medical University, Iwate 028-3694, Japan
*
Author to whom correspondence should be addressed.
Cancers 2024, 16(3), 523; https://doi.org/10.3390/cancers16030523
Submission received: 24 November 2023 / Revised: 21 January 2024 / Accepted: 22 January 2024 / Published: 25 January 2024
(This article belongs to the Special Issue Molecular Mechanisms in Prostate Cancer Development)

Abstract

:

Simple Summary

Prostate cancer (PCa) is characterized by various genomic alterations that play a pivotal role in carcinogenesis. Efforts in precision medicine aimed at improving diagnosis, prevention, and surveillance based on genetic alterations are advancing. Notably, no tumor markers surpass prostate-specific antigen in specificity, and existing treatments primarily target the androgen receptor axis, with exceptions for patients with alterations in homologous recombination repair-related genes, such as BRCA1/2 and ATM, who may benefit from poly (ADP-ribose) polymerase inhibitors. In order to delineate the current state of research on PCa, we provide an overview of cutting-edge genomic research, including genome alterations, cancer immunology, and non-coding RNAs, in PCa. These aspects are relevant to comprehending the molecular mechanisms underlying health disparities, PCa initiation and progression, and drug resistance in the precision medicine era.

Abstract

The progression of prostate cancer (PCa) relies on the activation of the androgen receptor (AR) by androgens. Despite efforts to block this pathway through androgen deprivation therapy, resistance can occur through several mechanisms, including the abnormal activation of AR, resulting in castration-resistant PCa following the introduction of treatment. Mutations, amplifications, and splicing variants in AR-related genes have garnered attention in this regard. Furthermore, recent large-scale next-generation sequencing analysis has revealed the critical roles of AR and AR-related genes, as well as the DNA repair, PI3K, and cell cycle pathways, in the onset and progression of PCa. Moreover, research on epigenomics and microRNA has increasingly become popular; however, it has not translated into the development of effective therapeutic strategies. Additionally, treatments targeting homologous recombination repair mutations and the PI3K/Akt pathway have been developed and are increasingly accessible, and multiple clinical trials have investigated the efficacy of immune checkpoint inhibitors. In this comprehensive review, we outline the status of PCa research in genomics and briefly explore potential future developments in the field of epigenetic modifications and microRNAs.

Graphical Abstract

1. Introduction

Prostate cancer (PCa) ranks as the fourth most commonly diagnosed cancer globally, accounting for 7.3% of all cases. In 2020, PCa held the position of the second most prevalent cancer (at 14.1%) and the fifth leading cause of cancer-related mortality (at 6.8%) among men [1]. Factors such as advanced age, African descent, and a family history of the disease are associated with poorer prognoses for PCa [2].
Age-adjusted incidence rates for PCa in Europe, America, and Oceania range from 59 to 73 cases per 100,000 men, surpassing those in Africa and Asia, where rates are 29.7 and 13.6 cases per 100,000 men, respectively. However, age-standardized mortality rates are highest in Africa (at 16.3 cases per 100,000 men), followed by Latin America, the Caribbean, Europe, Oceania, Northern America, and Asia (with rates of 4.1–14.2 per 100,000 men) [3].
Treatment for PCa primarily involves androgen deprivation therapy (ADT), androgen receptor (AR)-signaling inhibitors (ARSIs), and chemotherapy. Following initial hormone therapy, metastatic castration-sensitive PCa (mCSPC) often progresses to metastatic castration-resistant PCa (mCRPC) within approximately 1–2 years. The median survival time and cancer-specific mortality for mCRPC are 22 months (95% confidence interval [CI]: 21.0–22.9) and 30 months (95% CI: 28.3–31.7), respectively [4]. In one study, the 1- and 5-year survival rates for mCRPC were estimated at 58, 37, and 58% and 17, 4, and 9% in German, Swedish, and US cohorts, respectively [5]. In another study, the median overall survival (OS) for chemotherapy-naïve mCRPC was significantly longer in the ARSI group (treated with abiraterone acetate or enzalutamide; 34.7 and 35.3 months, respectively) than in the placebo group [6,7]. In Japanese patients with mCRPC, the median progression-free survival (PFS) and OS based on prostate-specific antigen (PSA) were 5.6 (95% CI: 4.7–6.8) months and 26.1 (95% CI: 22.1–30.5) months, respectively [8]. The median OS of diagnosed CRPC in patients with mCRPC or non-metastatic CRPC (m0CRPC) was 29.2 (95% CI: 15.3–37.4) months and 41.7 (95% CI: 16.1–not estimable) months, respectively, with no discernible difference between de novo (28.1 months; 95% CI: 15.3–38.2) and metachronous (33.3 months; 95% CI: 14.8–65.9) mCRPC at Iwate Medical University Hospital (our data).
Recently, precision medicine targeting driver genes, such as treatment employing poly (ADP-ribose) polymerase (PARP) inhibitors for BRCA mutations identified through companion diagnostics, has been realized. Next-generation sequencing (NGS) technology has facilitated comprehensive and exhaustive genome analysis, revealing a multitude of genetic mutations associated with cancer initiation and progression. Furthermore, NGS has enabled routine clinical examination of individual genetic abnormalities. Consequently, personalized treatments based on genetic anomalies have been developed, even in the context of PCa, with accumulating evidence supporting this approach. Recent reports have highlighted the utility of Decipher Prostate testing in determining the treatment approach for recurrence after radical prostatectomy [9]. Despite extensive genome-wide sequencing studies on PCa, no biomarker has surpassed PSA for several decades.
Therefore, in this review, we aim to provide an overview of recent advances in NGS findings, potential biomarkers for various treatments, and the current status and challenges of precision medicine for PCa based on genetic anomalies.

2. Genomic Alterations

In PCa, the involvement of driver genes is highly diverse in terms of both carcinogenesis and disease progression. This diversity varies significantly among different cases as well as across various stages of PCa. Considering the variation in driver genes, precision medicine assumes profound significance in the context of PCa (Table 1). The most prevalent genetic anomalies observed in PCa include point mutations in genes such as SPOP, FOXA1, and TP53, as well as copy number alterations (CNAs) in AR, MYC, RB1, PTEN, CHD1, and fusion genes associated with the ETS (E26 transformation-specific) family [10,11,12,13,14,15,16,17]. Alterations in pathways involving AR, TP53, the cell cycle, and MYC are strongly associated with the development of castration resistance and OS. By contrast, alterations in the Wnt pathway and SPOP mutations are indicative of a more favorable prognosis [18,19,20,21,22,23]. Notably, changes in AR and PI3K/Akt pathway components appear to result from continuous systemic therapy [24,25]. Conversely, PTEN loss, SPOP mutations, and alterations in DNA damage-repair mechanisms, including homologous recombination repair (HRR) and mismatch repair (MMR), tend to manifest early in prostate tumorigenesis, prior to metastatic progression, and generally remain stable over time [24,26].
The prevalence of specific alterations varies in different clinical contexts. For instance, alterations in TP53 (35% vs. 29%), BRCA2 (10% vs. 4%), PIK3CA (8% vs. 2%), RB1 (7% vs. 3%), and APC (11% vs. 9%) are more frequent in high-volume diseases than in low-volume diseases. Alterations in CDK12, PTEN, and FOXA1 are similarly detected between CSPC and CRPC, whereas those in ATM and RB1 exhibit a similar prevalence between de novo and recurrent metastatic disease, with CDK12 (6% vs. 1%) and FOXA1 (17% vs. 10%) alterations more commonly observed in the de novo metastatic state. The PI3K/Akt pathway is affected in 35–40% of mCSPC cases [18,19,20,21,22,23,28,31,32,33,34]. In mCSPC, TP53 and PTEN mutations, or CNAs, have been observed in 32% and 20% of cases, respectively [35].
In mCRPC, the most frequently aberrant genes include AR (62.7%), the ETS family (56.7%), TP53 (53.3%), and PTEN (40.7%). The frequency of alterations in these genes is significantly higher than that observed in localized cancer [36]. In mCRPC, non-AR-related clinically actionable alterations include aberrations in the PI3K/Akt pathway (49%), DNA repair pathway (19%), RAF kinases (3%), CDK inhibitors (7%), and Wnt pathway (5%) [36]. Notably, alterations in AR, TP53, RB1, PTEN, and ATM are enriched in mCRPC when compared with those in mCSPC [28].
Large-scale genome profiling targeting Asian populations has revealed genetic abnormalities (such as those in FOXA1, ZNF292, and CHD1) that differ in frequency from those previously reported in Caucasian populations. This suggests potential variation in driver genes for PCa among different ethnicities. Conversely, AR abnormalities are relatively rare in localized PCa, irrespective of ethnicity [37].

2.1. Fusion Genes from the ETS Family

Among the most prevalent genomic alterations in PCa are fusions involving androgen-regulated promoters with ERG and other members of the ETS family of transcription factors. ETS fusions are observed in over 50% of cases [10,36,38,39]. In particular, the TMPRSS2–ERG fusion represents the most frequent molecular alteration in PCa [40], occurring in 40–50% cases, equivalent to >100,000 cases annually in the United States alone [41]. Notably, Asian patients with PCa exhibit fewer TMPRSS2–ERG fusions than do Caucasian or African-American patients [42,43,44]. TMPRSS2 expression is upregulated by androgenic hormones in PCa cells but downregulated in androgen-independent PCa tissue [44]. Genetic polymorphisms associated with PCa susceptibility may vary by race and ethnicity. For instance, Takata et al. [45] identified five SNPs (rs13385191, rs12653946, rs1983891, rs339331, and rs9600079) linked to PCa susceptibility through a genome-wide association study of a Japanese population. However, only two SNPs (rs13385191 and rs1983891) were associated with PCa susceptibility in a Latino population [46]. The TMPRSS2 Met160Val polymorphism is a genetic risk factor for sporadic PCa in the Japanese population [44]. In the early stages of PCa, frequent genomic alterations include TMPRSS2–ERG fusion in 40–60% of patients [10,47]. The presence of TMPRSS2–ERG in PCa has been associated with more aggressive biological behavior and a poorer prognosis [48,49].

2.2. AR and AR-Related Genes

The dysregulation of AR and AR pathway-related genes plays a central role in PCa pathogenesis. Abnormalities in the AR pathway include gene amplifications, mutations, splicing variants, and ligand-independent AR activity. Furthermore, these abnormalities can result from coding regions as well as AR enhancer region amplifications and genomic structural rearrangements capable of producing constitutively active truncated AR isoforms, notably AR splice variant 7 (AR-V7) [50].
Over 70% of PCa cases exhibit AR pathway aberrations, including those involving SPOP or FOXA1, with the majority resulting from direct alterations affecting AR through amplifications and mutations [36]. Alterations in androgen receptor signaling are more frequently observed in metastatic samples, with AR amplification or mutation being the most common alteration (50–80%). SPOP mutations, while present, are relatively less frequent in metastatic samples [10,36,51,52]. AR alterations are rarely observed prior to ADT, although they are enriched in patients with mCRPC who have a higher disease burden and prior exposure to ARSIs [53]. By contrast, changes in AR copy number status are not evident in patients treated with docetaxel or cabazitaxel [54,55]. Additionally, truncated AR splice variants were previously believed to be primarily expressed in mCRPC, and the presence of AR-V7 was associated with resistance to hormone therapy and taxane chemotherapy [56,57]. The detection of AR gene amplification in plasma cell-free DNA has consistently been linked to shorter PFS and OS in patients treated with ARSIs [58,59,60]. In a past study, the AR activity of tumors was assessed by calculating the AR activity score based on the expression patterns of 20 genes that were experimentally validated as AR transcriptional targets. This scoring indicated a wide range of AR activity between PCa cases and among different genomic subtypes [61].
SPOP plays a role in protein degradation by acting as a substrate adaptor, enabling ubiquitination by the Cullin3–RING complex [62]. In PCa, SPOP functions as a tumor-suppressor gene by promoting the degradation of multiple oncogenic substrates, including AR and ERG [63,64]. SPOP mutations represent the most common point mutations in localized PCa [10,12]. The frequency of SPOP mutations in localized disease (10–15%) and metastatic PCa (5–10%) aligns with the observation of a more favorable disease prognosis in de novo mCSPC [23]. Alterations in SPOP are enriched in locoregional and early-stage disease; therefore, tumors with SPOP mutations may exhibit increased sensitivity to ADT [12,26,28]. Furthermore, the potential of SPOP mutations as predictive biomarkers may manifest in enhanced sensitivity to second-generation ARSIs and taxane chemotherapy in mCRPC [65].

2.3. TP53

When comparing clonal alterations between a matched primary localized tumor and a subsequent metastatic tumor from the same patient, tumors acquired at a later stage typically exhibit a higher mutation count than do tumors acquired at an earlier stage. TP53 alterations emerge early in affected patients and are present in all tumors from the same patient. TP53 mutations are clonal, even in cases of both primary localized and subsequent metastatic PCa [28]. A high frequency of TP53 alterations is observed in primary localized tumors, consistent with prior reports of aggressive behavior of PCa caused by TP53 alterations [66,67,68]. The timing of TP53 alterations in tumors, specifically whether these alterations are present in tumors early in their evolution or acquired later during disease progression, remains uncertain. Somatic alterations in BRCA2 have been identified in matched tumors, suggesting that somatic loss-of-function alterations in this gene occur early in tumorigenesis. By contrast, alterations in AR do not occur early in matched samples, which is consistent with treatment-related changes that promote castration resistance [28].

2.4. Mitogen-Activated Protein Kinase (MAPK) Pathway

Genes within the MAPK pathway are affected in 25% of tumors [10]. The presence of RAF kinase fusions in 3% of individuals with mCRPC suggests the potential utility of pan-RAF inhibitors or MEK inhibitors [36].

2.5. PI3K/Akt Pathway

Alterations in the PI3K/Akt pathway constitute the second most frequently observed pathway aberration in metastatic PCa [10,36,69,70]. The PI3K/Akt pathway may be antagonized by several phosphatases, including PTEN and PHLPP. PI3K activation results in the phosphorylation of AKT and its downstream genes, including the mammalian target of rapamycin (mTOR). Phosphorylated AKT is an indicator of PI3K/Akt pathway activation [71,72], and AKT activates nuclear factor-κB (NF-κB) signaling in CRPC, with levels of NF-κB being higher in castration-resistant cell lines than in androgen-dependent cell lines [73]. PTEN, which leads to the downstream activation of the PI3K/Akt and AR signaling pathways, plays a crucial role in tumor suppression, and its loss is associated with poorer clinical outcomes in patients with mCRPC [18,53,74,75]. Considering that PTEN-deleted tumors may rely on PIK3CB owing to feedback inhibition from PIK3CA, the simultaneous loss and mutation of PTEN and PIK3CB could result in increased PI3K/Akt pathway activity. This observation suggests the existence of a subset of tumors for which a combination of PI3K and androgen signaling inhibition might be effective. In the context of the PI3K/Akt pathway, specific inhibitors targeting PIK3CB may be beneficial for individuals with mutations, amplifications, and/or fusions involving the related gene. Notably, some individuals who demonstrate sustained responses (lasting more than 1 year) to PIK3CB-specific inhibition possess activating mutations or amplifications in PIK3CB [27]. PTEN loss occurs in 30–50% of PCa cases [36,51,52] and is recognized as an early genomic event in PCa initiation [10,76]. However, heterogeneity in clonal composition is evident in castration-naïve tumors [77]. In primary PCa, the deletion or mutation of PTEN is observed in 17% of cases [10]. In mCRPC, biallelic loss and hotspot mutations of PTEN (40–60%), as well as amplifications and activating fusions in PIK3CA (5%) and AKT1 (1%), are observed [15,36,78].

2.6. NF-κB Signaling

The NF-κB/Rel proteins include NF-κB2 p52/p100, NF-κB1 p50/p105, c-Rel, RelA/p65, RelB, and others. These proteins function as dimeric transcription factors, primarily regulating the transcription of genes involved in a wide range of biological processes such as apoptosis, angiogenesis, inflammation response, cell survival, inflammation, stress response, B-cell development, and lymphoid tissue formation [79,80,81,82,83,84].
The activation of NF-κB (p65) and Sonic Hedgehog pathways in PCa is positively correlated, and the overexpression of NF-κB(p65), SHh, and GLI1 is observed in aggressive PCa tissue cores featuring a high Gleason score and advanced clinical stage. These pathways play key roles in the progression of advanced PCa and are potential targets for therapeutic intervention [85].
The involvement of NF-κB signaling in the onset and maintenance of CRPC is considerable, irrespective of whether it is due to abnormal AR activity or AR-independent mechanisms [86]. Moreover, NF-κB plays a crucial role in the development of treatment resistance and metastasis in PCa. Specifically, the RelB-activated noncanonical NF-κB pathway could prolong and enhance NF-κB activity [87]. In addition, the NF-κB complex can promote the survival of tumor cells in PCa through anti-apoptotic signaling. The activation of NF-κB is sustained by a positive feedback loop. Such activation is associated with PCa progression to a metastatic phenotype and mCRPC. A novel role of NF-κB kinase (IKK)-α in the progression of NF-κB-independent PCa and its progression to CRPC has been elucidated, revealing a mechanistic link between inflammation and PCa [88]. Additionally, NF-κB is implicated in the activation of the PI3K/Akt pathway and some non-coding RNAs, contributing to PCa progression [73,89,90,91].

2.7. Cell Cycle Pathway

RB1 loss is detected in 21% of PCa cases [36]. Focal amplifications involving CCND1 (9%), as well as less frequent (<5%) events affecting CDKN2A/B, CDKN1B, and CDK4, are observed in mCRPC [36]. Disruptions in the cell cycle, such as CCND1 amplification or CDKN2A/B loss, may lead to an increased responsiveness to CDK4 inhibitors, as observed in other tumor types [27]. Pre-clinical findings indicate that selectively targeting CDK4/6 represents a therapeutic approach for both early-stage and advanced prostate PCa [92]. Mutations or deletions of TP53 and RB1 are more prevalent in metastatic than in localized PCa [10,36], and these alterations are associated with a poorer clinical outcome and aggressive behavior [35,66,67,93,94].

2.8. DNA Repair Pathway

Alterations in the DNA repair pathway are more frequent in mCRPC than in localized PCa or CSPC [10,36]. Genetic abnormalities associated with the DNA damage response (DDR) are observed in approximately 23–28% of mCRPC cases, including both germline and somatic variants [10,11,20,28,29,30].

2.8.1. HRR Genes

Among common gene abnormalities, BRCA2 and ATM mutations constitute a significant majority [28]. In advanced PCa, BRCA2 (7–13%), BRCA1 (0–1%), ATM (5–7%), and CDK12 (4–6%) exhibit frequent mutations [36,51,52,95]. The prevalence of pathological variants in BRCA1/2 and ATM varies across different stages of PCa (Table 1) [10,36,76,96]. ATM alterations are associated with a poor prognosis; germline mutations are linked to higher tumor grades and lethal PCa [97,98,99]. CDK12 alterations are associated with aggressive clinical features and a poor prognosis in mCRPC [100]. In a study of PCa patients who underwent radical prostatectomy or localized radiation therapy, both metastasis-free survival and disease-specific survival were significantly shorter in patients with BRCA mutations than in non-carriers [101]. In localized PCa, BRCA2 germline mutations are associated with early-onset, high-grade tumors and a poor prognosis [101,102]. BRCA2 alterations are enriched in mCRPC relative to their frequency in mCSPC [35]. Notably, approximately half of these pathological variants are attributed to germline mutations, and they are frequently accompanied by evidence of tumor loss-of-heterozygosity [28,103] as well as characteristic patterns of genomic and epigenomic instability [104]. The presence of germline BRCA2 alterations in primary PCa is associated with poor survival outcomes [105]. In addition, Japanese patients with mCRPC-causing germline BRCA1/2 mutations exhibit a significantly poorer prognosis and PFS following ARSI and taxane chemotherapy than non-careers [106]. BRCA2 germline and/or somatic mutations are linked to aggressive pathologic features such as intraductal carcinoma of the prostate [107,108], and patients with these mutations appear to exhibit a poorer response to ARSIs than those seen in patients with wild-type BRCA2 [53,109]. Regardless of patient ethnicity, PCa with germline or somatic HRR abnormalities exhibits increased aggressiveness, leading to enhanced resistance to ARSI and resulting in poor OS for CRPC [8,105,109,110,111,112,113]. Furthermore, the OS of patients with CRPC who underwent ARSI or were administered olaparib varies depending on the specific mutations within HRR-related genes. Therefore, personalized treatment recommendations are likely to be based on the results of genomic profiling (Table 2).

2.8.2. MMR Genes and Microsatellite Instability (MSI)

Tumors with MMR gene deficiencies or MSI often exhibit an augmented antitumor immune response characterized by an increased density of tumor-infiltrating lymphocytes (TILs) [114,115]. This phenomenon is attributed to elevated mutation rates and enhanced levels of neoantigens in MMR-deficient tumors. Such neoantigens arise through various mechanisms, including mutant peptides, frameshift mutations, and insertions or deletions (indels) in coding microsatellites [116,117]. These neoantigens are presented on the cell surface by MHC-I molecules, facilitating the T-cell-mediated elimination of tumor cells. Approximately 3–8% of PCa cases, with a germline mutation rate of approximately 1%, are linked to the deficiency of MMR genes, including MSH2, MSH6, PMS2, and MLH1. Such deficiency leads to hypermutation, MSI, and characteristic trinucleotide mutational signatures [28,36,118,119,120,121,122,123]. MMR genes play crucial roles in the recognition and repair of incorrect insertions, deletions, and base misincorporations during DNA replication or recombination [124]. Among them, germline mutations in genes associated with Lynch syndrome (21.9%) have been observed [118]. Hypermutation occurs in 12% of advanced PCa cases, all of which exhibit mutations in MMR genes and exhibit MSI. Such mutations often involve complex structural rearrangements in MSH2 or MSH6 rather than epigenetic silencing of MLH1 [122]. Although MMR protein loss and MSI are evident in both primary and advanced PCa, only a few MMR mutations have been identified [125,126,127,128,129,130,131,132]. Mutations in MMR genes in PCa are strongly associated with elevated neoantigen expression and TIL accumulation [133].

3. Immune Environment

In the context of PCa, the immune environment and hormone therapy involve macrophages/antigen-presenting cells (APCs), CD8-positive cytotoxic T lymphocytes, CD4-positive helper T lymphocytes, and natural killer (NK) cells, all of which recognize and target cancer cells, contributing to antitumor immunity. Several studies have reported on the expression of these cells within prostate tissues; notably, fewer TILs have been identified in poorly differentiated PCa than in benign prostatic hyperplasia, suggesting a potential link between immune cell deficiency and cancer progression. The infiltration of these immune cells within the tumor microenvironment of PCa tissue has also been associated with prognosis; patients with PCa and a higher number of infiltrating TILs tend to have better outcomes than those with fewer infiltrating TILs [134,135,136]. The examination of CD68-positive tumor-infiltrating macrophages (TAMs) in radical prostatectomy specimens of patients with PCa revealed that cases with a higher number of infiltrating cells were associated with a significantly worse prognosis for PSA recurrence than those with fewer infiltrating cells [137]. In PCa, an efficient tumor immune response is hindered by TILs that are either inactive or belong to the immunosuppressive T regulatory cell (Treg) subtype. Tregs represent subgroups of CD4+CD25+ and CD8+/FoxP3+ T lymphocytes with the ability to directly suppress immune responses. These Tregs are negatively correlated with prognosis in patients with PCa, as they contribute to cancer progression [138,139,140]. ADT is associated with the regulation of the immune system, i.e., the control of male hormones can either promote thymus formation and B-cell proliferation or suppress the tumor-infiltrating Tregs and can induce NK cell infiltration and CD4+ T-cell infiltration within PCa tissues. This regulation of the immune response in PCa may enhance the antitumor effects through the combination of hormone therapy and immunotherapy [141,142,143,144].

4. Epigenetic Modifications

Epigenetic traits represent inheritable characteristics resulting from chromosomal alterations or DNA modifications without alteration of the DNA sequence [145]. Epigenetic modifications, including acetylation, methylation, ubiquitination, and phosphorylation, play pivotal roles in transcription, DNA repair, and replication [146]. Epigenetic regulation is a dynamic and reversible process involving the addition of epigenetic marks to histones or DNA, the recognition or recruitment of these marks, and the removal of these marks by epigenetic writers, readers, and erasers, respectively.
PCa is associated with alterations in DNA methylation, histone modification, chromatin accessibility, and 3D chromatin structure. Such changes contribute to disease development and progression by influencing gene expression, promoting cellular plasticity, and facilitating CRPC emergence under treatment pressure. The dysregulation of chromatin and epigenetics confers the complete range of cancer hallmarks by altering transcriptional regulation, mediating all defined hallmarks of cancer, and promoting cancer cell plasticity [147].
Aberrant histone modifications can lead to the upregulation of oncogenes or the suppression of tumor-suppressor genes. Notably, histone methylation/acetylation and DNA methylation centrally control gene expression, thereby promoting the progression and metastasis of PCa [148,149,150,151,152]. Chromatin regulators such as EZH2 play a critical role in driving epigenetic reprogramming in PCa, being overexpressed in PCa and significantly overexpressed in CRPC [153,154,155]. Bromodomain (BRD)-containing proteins serve as chromatin readers, recognizing mono-acetylated histones and thereby initiating chromatin remodeling and transcription. Dysregulation of BRDs is a common occurrence in cancer, including metastatic PCa [156]. Within the BRD family, BRD4 stands out as an important co-regulator of AR [156,157], with its protein expression escalating in the context of castration resistance and being associated with a poor prognosis. BET (bromodomain and extraterminal) inhibitors have exhibited antitumor activity in CRPC models, attenuating AR signaling and reducing c-MYC expression [158,159]. Moreover, BET inhibitors have been demonstrated to decrease AR-V7 levels by regulating the splicing factors required for AR-V7 generation [160].

4.1. DNA Methylation

The contribution of DNA methylation to PCa development was first reported in the early 1990s [161]. Approximately 22% of tumors in PCa are associated with hypermethylation [162]. DNA methylation is linked to gene silencing and occurs when a methyl group is added to the C5 position of the cytosine residues in CpG dinucleotides [163]. DNMT (DNA methyltransferase) enzymes catalyze the addition of 5-methyl cytosine to DNA, which can be removed by DNA demethylases from the ten-eleven translocation (TET) family [164]. In addition, approximately 60% of all gene promoters are associated with CpG islands [165]. Consequently, abnormal DNA hypermethylation at CpG islands can lead to gene silencing, including the inactivation of tumor-suppressor genes [47,166]. The downregulation of TET2 has been implicated in the regulation of AR signaling and the development of PCa [167,168]. The diminished expression of both TET2 and 5-hydroxymethylcytosine, associated with PCa progression, has been identified as a potential prognostic biomarker [169,170].
Cytidine methylation of regulatory sequences near GSTP1, which encodes an enzyme responsible for safeguarding DNA against oxidants and carcinogens, has been associated with prostate tumorigenesis [171]. In PCa, GSTP1 is methylated in its promoter region, leading to the diminished expression of GSTP1 in tumor cells. Hypermethylation of the GSTP1 promoter occurs in approximately 75% of pre-invasive high-grade prostatic intraepithelial neoplasms and in over 90% of prostate tumors and persists through all stages of PCa progression [172]. In patients with localized disease, pre-operative circulating methylated GSTP1 in plasma has been demonstrated to predict PSA recurrence and tumor aggressiveness [173,174,175].
Promoter hypermethylation can lead to the silencing of not only tumor-suppressor genes but also key receptors, such as AR and ESR1. Additionally, cell adhesion genes (such as CD44 or CDH1), cell cycle genes (such as CCND2, CDKN1B, or SFN), and apoptosis-related genes (such as ASC, BCL2, DAPK, or PTGS2) can be affected by this process [176]. Although global hypomethylation is common in primary tumors, it becomes more pronounced in metastatic PCa [177]. Notably, 22% of mCRPC tumors exhibit distinct epigenomic features characterized by hypermethylation and the enrichment of somatic mutations involving TET2, DNMT3B, IDH1, and BRAF [162].
DNMT inhibitors such as azacytidine and decitabine have been developed to target aberrant DNA hypermethylation; furthermore, their efficacy against PCa has been clinically evaluated (NSC127716). In a phase 1/2 study of azacitidine in combination with docetaxel in mCRPC, a PSA response was observed in 52% (10/19) of patients, with no dose-limiting toxicity reported [178].

4.2. Histone Methylation

Histones undergo methylation when one, two, or three methyl groups are added from S-adenosylmethionine to the side chains of arginine, lysine, and histidine residues. Histone methylations, such as H3K4me1, H3K9me2, and H3K9me3, are reduced in PCa tissues when compared with their levels in normal prostate tissues [179]. However, H3K27me3 marks are strongly enriched in the promoter regions of tumor-suppressor genes in metastatic PCa when compared with those in localized PCa and normal prostate tissues [150,180].
The increased genomic distribution of H3K27me3 in metastatic PCa can primarily be attributed to the overexpression of EZH2 [154], a histone methyltransferase. EZH2 plays a pivotal role in promoting lineage plasticity and differentiation changes, which are closely associated with neuroendocrine PCa (NEPC) [181]. The polycomb group proteins EZH2 and embryonic ectoderm development (EED), another member of the PRC2 complex, are currently under investigation as therapeutic targets [182]. EED functions as a direct regulator of AR and its downstream genes, working in conjunction with EZH2 in the context of AR-positive PCa [183]. EZH2 represents a promising target, and several EZH2 inhibitors have been identified in early-stage clinical studies [184]. These EZH2 inhibitors are currently under evaluation in clinical trials, both as standalone treatments and in combination with ARSIs or immunotherapy for PCa [185].
Conversely, histone demethylases are enzymes responsible for removing methyl groups from histones. LSD1 (KDM1A), a multiple histone demethylase, is overexpressed in patients with advanced PCa [186]. LSD1, which specifically demethylates H3K4me1 and H3K4me2 [187], interacts with AR, activating AR-dependent transcription or a subset of cell cycle gene expression [188,189]. A small-molecule LSD1 inhibitor, SP2509, has been shown to attenuate tumor growth in CRPC xenograft models [190]. The inhibition of LSD1 disrupts the chromatin binding of FOXA1, a transcription factor for AR [191]. By regulating AR binding and gene expression, LSD1 inhibition not only suppresses tumor growth but also exhibits a synergistic interaction with enzalutamide in CRPC models [190]. These findings lend support to the use of LSD1 as a potential therapeutic target, with the LSD1 inhibitor INCB059872 currently undergoing clinical development. In addition, a clinical trial involving a novel LSD1 inhibitor, CC-90011, has recently been initiated [192].

4.3. Histone Acetylation

Histone acetylation involves the addition of an acetyl group to the lysine residues of histones, leading to the opening and activation of chromatin. Typically, this process is associated with the activation of transcription, whereas histone deacetylation is often linked to gene silencing [193]. Super-enhancers, which are clusters of enhancers marked by high H3K27ac levels, play a crucial role as drivers of oncogenic processes in cancer cells [194,195]. The activation of histone acetyltransferases, such as p300 and CREB-binding protein (CBP), is strongly associated with an increase in H3K27ac modification frequency [196]. Moreover, p300 and CBP play pivotal roles in the regulation of key genes, including AR target genes, in PCa [197]. Inhibitors targeting p300 and CBP, including CCS1477, A-485, and FT-7051, have been developed [198,199]. The role of histone deacetylase inhibitors in CRPC treatment has also been investigated. This exploration involves the downregulation of AR, AR splice variants, and c-Myc expression. Additionally, these inhibitors have been associated with loss-of-function mutations in p300 or CBP, driving synthetic lethality [197]. Clinical trials evaluating the effectiveness of CCS1477 and FT-7051 in treating PCa have commenced recently.
Histone deacetylases (HDACs) are enzymes responsible for removing histone acetylation; thus far, several types of HDACs have been identified in humans [200]. Notably, there are five classes of HDAC inhibitors, including hydroxamic acids, cyclic tetrapeptides, short-chain carboxylic acids, benzamides, and ketoderivatives [201]. HDAC overexpression has been observed in various malignancies, including PCa [202]. Notably, the expression of HDAC1 and HDAC2 is positively correlated with higher Gleason scores in PCa, whereas that of HDAC1/2/3 is positively associated with the proliferative marker Ki67 [203]. Concordantly, HDAC expression is linked to poor clinical outcomes, and HDAC inhibitors are regarded as potential therapeutic options [203]. Several HDAC inhibitors, such as vorinostat, pracinostat, panobinostat, and romidepsin, have undergone testing in phase 2 clinical trials for PCa [204,205]. However, most patients in these trials experienced either toxicity from these agents or disease progression [206]. Clinical trials involving HDAC inhibitors have not yielded significant success, mainly owing to issues such as poor oral bioavailability, lack of drug selectivity, and other mechanisms that require further exploration [206].

5. Non-Coding RNAs

5.1. Long Non-Coding RNAs (lncRNAs) and Micro-RNAs (miRNAs)

lncRNAs have been implicated in prostate carcinogenesis. They exert their influence through various mechanisms, including regulation of the AR expression pathway, epithelial–mesenchymal transition, miRNAs, and the PI3K/Akt/mTOR pathway; these processes are governed by PCa gene expression marker 1 (PCGEM1), PlncRNA-1, colon cancer-associated transcript 2 (CCAT2), C-terminal binding protein 1 antisense (CTBP1-AS), growth arrest-specific 5 (GAS5), long intergenic non-protein-coding RNA 1296 (LINC01296), LOC400891, and LOC440040, etc. [207,208,209,210,211,212,213,214,215,216].
LINC00624 is overexpressed in both PCa tissues and cells and is associated with malignant progression. Furthermore, the relationship between NF-κB and PCa is linked to the expression of LINC00624, as it forms a co-regulatory axis with TEX10, stimulating NF-κB activity. Therefore, LINC00624 is hypothesized to function as an oncogene in PCa progression [90]. The activation of AKT and NF-κB signaling involves a crucial regulatory protein complex comprising PHLPP, FKBP51, and IKKα. The lncRNA PCAT1 is implicated in the regulation of the PHLPP/FKBP51/IKKα complex, directly binding to FKBP51 and excluding PHLPP from the complex. This regulation by PCAT1 contributes to CRPC progression by activating AKT and NF-κB signals [91].
Upregulated AR expression is a hallmark of CRPC and may serve as a mechanism facilitating AR signaling in the presence or absence of low endogenous androgen levels. Specific lncRNAs, such as CTBP1-AS and HOTAIR, exhibit upregulation following ADT, signifying their negative regulation by androgens. This upregulation of lncRNAs results in increased AR expression. CTBP1-AS upregulation leads to the downregulation of CTBP1, which negatively modulates AR by inhibiting androgen-mediated demethylation. Consequently, CTBP1-AS-induced CTBP1 downregulation results in AR upregulation [207]. HOTAIR binds to the N-terminal domain of the AR, hindering the binding of the E3 ubiquitin ligase MDM2, which interacts with the AR through the same domain. This blockade prevents AR ubiquitination and degradation [217]. PCGEM1 and PRNCR1 bind to AR, enhancing AR-mediated gene activation processes. PCGEM1 is reportedly dependent on PRNCR1 for binding. The relationship between AR and PCGEM1/PRNCR1 was examined via in vitro experiments using androgen-sensitive LNCaP cells [218]. Both PCGEM1 and PRNCR1 were significantly upregulated in CRPC cell line models and activated AR in the absence of androgens. The knockdown of either PCGEM1 or PRNCR1 significantly inhibited in vivo tumor growth in a xenograft mouse model [209], but these proteins were not implicated in CRPC or AR signaling. PCGEM1, but not PRNCR1, was associated with PCa. Notably, neither PCGEM1 nor PRNCR1 were associated with poor clinical outcomes [212]. RNA-sequencing data from two distinct androgen-dependent models revealed that PCGEM1 was considerably expressed in PCa, whereas PRNCR1 was minimally expressed [212]. PCGEM1 was implicated in early-stage PCa but not in metastatic PCa, being significantly downregulated in metastatic tumors relative to primary tumors and significantly repressed in patients with poor outcomes [218]. Significantly reduced PCGEM1 expression and AR activity were also observed post-castration [209]. Conversely, the AR stimulation-induced upregulation of PCGEM1 expression was observed in vivo [218].
Interaction with miRNAs is another mechanism by which lncRNAs regulate AR activity. PlncRNA-1 likely protects AR from miRNA-mediated suppression [211] by acting as a sponge for certain AR-targeting miRNAs, including miR-34c and miR-297. PCGEM1 and HOTAIR are reportedly targeted by miR-34a [219] and miR-145 [210], respectively.
In PCa, both miR-145-5p (the guide strand) and miR-145-3p (the passenger strand) are downregulated. The introduction of these microRNAs into PCa cells reportedly suppresses their proliferative, migratory, and invasive capabilities. Furthermore, miR-145-3p regulates the expression of cancer-promoting genes such as MELK, NCAPG, BUB1, and CDK1. Patients with PCa who exhibit the expression of these genes are more commonly associated with CRPC than with CSPC [220].
The cancer-suppressive functionality of both guide and passenger strands of several microRNAs, including miR-30a, miR-139, miR-143, and miR-145, has been established through large-scale cohort analysis using The Cancer Genome Atlas [221]. The miRNA expression profile of PCa using needle biopsy samples from CSPC and normal prostate tissues revealed that miR-222 and miR-31 exhibit cancer-suppressive functions [222].
miR-130b exhibits significant downregulation among miRNAs in PCa tissues, and its expression is significantly reduced in PCa cell lines as well. The decreased expression of miR-130b significantly promotes the proliferation, invasion, and tubule formation of PCa cells, whereas its overexpression suppresses angiogenesis in PCa. miR-130b directly regulates TNF-α, inhibiting the NF-κB signaling pathway and its downstream gene, VEGFA. Furthermore, VEGFA, by reducing the expression of miR-130b, contributes to a feedback loop involving miR-130b/TNF-α/NF-κB/VEGFA that is associated with angiogenesis in PCa [89]. Moreover, elevated miR-130b levels are reportedly correlated with race disparity and PSA levels/failure in African-American patients. miR-130b is significantly overexpressed in this population when compared with its levels in tissue samples and cell lines from European-American patients. Elevated miR-130b levels reportedly act as an independent prognostic biomarker for African-American patients, highlighting their potential role as a risk factor for poor prognosis in such individuals with PCa. FHIT and CDKN1B are direct functional targets of miR-130b. Additionally, other miR-130b target genes, including TP53INP2, PTEN, FOXO1, KDM2A, BTG1, and CDKN1A, are upregulated in anti-miR-130b-stable cells when compared with those in controls [223].
In PCa, reports of liquid biopsy using upregulated or downregulated microRNAs in the blood (serum, plasma, exosomes) and urine as indicators have been increasing [224] [225]. Several studies suggest the detection of cell-free microRNAs, such as miR-141, miR-375, miR-107, and miR-221, in the blood as useful biomarkers for PCa [226]. Combining miR-17-3p and miR-1185-2-3p for PCa diagnosis has exhibited >90% sensitivity and specificity, highlighting its promise in clinical applications and circumventing the need for unnecessary prostate biopsies [227].

5.2. Role of miRNAs in PCa Drug Resistance

Growing evidence suggests that miRNAs play a pivotal role in anti-AR drug resistance. For example, miR-23b and miR-27b sensitize CRPC cells to flutamide by targeting CCNG1 [228]. miR-221 and miR-222, upregulated in CRPC cells, maintain resistance in PCa [229], and miR-663 affects castration resistance by modulating the AR signal [230]. In mCRPC, patients with higher levels of miR-375 and miR-3687 in whole blood exhibit a shorter time to progression with enzalutamide treatment [231]. Other miRNAs target AR and its splicing variant, AR-V7, to resensitize drug-resistant PCa cells. For instance, miR-212 inhibits hnRNPH1, thereby reducing AR and AR-V7 expression, whereas miR-361-3p enhances enzalutamide sensitivity by targeting AR-V7 [232,233,234]. In addition, miRNAs can serve as prognostic biomarkers in CRPC. High miR-1290 and miR-375 expression is associated with poor survival rates and helps to predict the CRPC stage [234].
Additionally, miRNAs play a key role in taxane resistance. Several miRNAs, including miR-34b-3p, miR-199a, miR-200b-3p, and miR-375, have been implicated in paclitaxel resistance [235,236]. Docetaxel resistance is associated with miR-375 and miR-323 upregulation, whereas miR-195, miR-204, miR-143, and miR-200b can enhance docetaxel sensitivity in PCa [237,238,239,240]. Furthermore, elevated levels of six plasma microRNAs (miR-132, miR-200a, miR-200b, miR-200c, miR-375, and miR-429) have been identified as predictors of decreased OS with docetaxel treatment [241]. Such findings indicate the potential of circulating miRNA expression in the blood as a complementary tool to PSA measurements both in diagnosis and during treatment and highlight the complex roles of miRNAs in the development of resistance to antiandrogen and taxane therapies in PCa.
In the context of PCa, there is an increasing research interest in developing a minimally invasive liquid biopsy using blood or urine to assess treatment progress. However, selecting appropriate miRNA indicators poses a major challenge. Notably, significant variation in miRNA levels among different research facilities is apparent when reviewing research on miRNA levels in urine from patients with PCa [242,243,244,245,246] (Table 3). This inconsistent pattern confounds the utilization of these miRNAs for universal diagnostics. Therefore, further research on the development of liquid biopsy for PCa and the standardization of specimen collection methods, RNA extraction techniques, and detection systems is warranted.

6. Prostate-Specific Membrane Antigen (PSMA)

PSMA is a type II transmembrane protein expressed in most high-grade, metastatic, androgen-insensitive, clinically significant cases of PCa. More than 90% of PCa cases exhibit PSMA overexpression at levels reported to be 100–1000 times higher than those in normal prostate tissue [247,248,249,250,251]; however, PSMA is scarcely expressed in benign or hyperplastic prostate tissue [252]. The organs reported to exhibit high PSMA expression, apart from the prostate, are the kidney, bladder, salivary glands, ganglion, liver, spleen, duodenum, and colon [253,254,255]. PSMA is a favorable target for molecular imaging [256,257,258,259,260,261,262] and therapy [263,264,265,266,267] in PCa. The US Food and Drug Administration (US FDA) has approved three types of radionuclides for positron emission tomography (PET) imaging of PSMA, namely F-18 piflufolastat PSMA, Ga-68 PSMA-11, and F-18 flotufolastat PSMA, for patients with PCa. PSMA PET-computed tomography (PET/CT) or magnetic resonance imaging (PET/MRI) are recommended for detecting metastases and restaging PCa in cases of biochemical recurrence [268]. In the Advanced Prostate Cancer Consensus Conference (APCCC 2022), 77% of panelists reached a consensus to perform PSMA PET/CT on the majority of patients with clinically localized high-risk PCa. However, 92% of panelists voted not to recommend PSMA PET/CT or MRI for patients with clinically localized favorable intermediate-risk (NCCN definition) PCa [269]. Determining superiority between 68Ga and 18F-PSMA PET/CT is challenging because their accumulation in normal organs is similar and their diagnostic accuracy does not differ significantly [260,261,262,270]. In clinical practice, a lack of high-level evidence remains in several aspects of PSMA PET/CT or MRI, and we look forward to future research contributing such evidence.
Radiation therapy utilizing the labeling of PSMA ligands to target PSMA-positive cells and the surrounding microenvironment involves both alpha and beta radionuclides such as 225Ac-PSMA and 177Lu-PSMA, respectively [255,271,272,273]. In phase 3 international collaborative non-blinded trial evaluating 177Lu-PSMA-617 in patients with mCRPC who had a treatment history involving one or more ARSI treatments and 1–2 taxane chemotherapy regimens and were PSMA-positive on PSMA PET/CT, 177Lu-PSMA-617 with standard treatment extended both the image-based PFS (median 8.7 vs. 3.4 months, hazard ratio 0.40, 99.2% CI 0.29–0.57, p < 0.001) and OS (median 15.3 vs. 11.3 months, hazard ratio 0.62, 95% CI 0.52–0.74, p < 0.001) significantly compared to those of standard treatment alone [263]. Therefore, 177Lu-labeled radiopharmaceutical 177Lu-PSMA-617 was approved by the US FDA and European Medicines Agency in 2022 as a last-line therapy for mCRPC. Currently, a 225Ac-labeled PSMA-targeting inhibitor is also being developed. Numerous clinical trials targeting localized CSPC (ClinicalTrials.gov number: NCT04430192, NCT04443062, and NCT05162573), mCSPC (Table 4), and CRPC (Table 5) are currently underway.

7. Precision Medicine: Genome-Based Therapy

7.1. PARP Inhibitors (PARPis) for mCRPC with Alterations in HRR Genes

PARP recognizes single-strand DNA breaks and facilitates the transportation of base-excision repair proteins for genetic repair. Conversely, several proteins, including BRCA, play a pivotal role in HRR, particularly in the repair of double-strand DNA breaks. DNA repair can still occur in cases where a malfunction occurs in either PARP or BRCA. However, the administration of PARPis prevents DNA repair in cases where cancer cells exhibit BRCA abnormalities [96,275,276,277,278]. PARPis traps PARP-1 on damaged DNA, inhibiting auto-poly(ADP-ribosyl)ation and enhancing DNA affinity for the catalytic site, ultimately leading to cell death [96,279,280,281]. Notably, tumor cells with BRCA mutations are considerably more sensitive to PARPi treatment, exhibiting an approximately 1000-fold higher sensitivity than wild-type BRCA cells [282]. Among all HRR genes, the biallelic loss of BRCA2 appears to yield the most significant benefit from PARP inhibition [283]. Durable responses to one class of treatment following progression to the other class appear to be infrequent [284], likely owing to shared resistance mechanisms, including BRCA reversion mutations [285]. Assuming that abnormalities in ATM and BRCA1/2 confer increased sensitivity to PARPis, 19.3% of individuals with mCRPC may likely benefit from this therapy [122]. However, olaparib is ineffective in CRPC with ATM mutations [286].
PARPis, which were approved by the US FDA in 2023 and are used for monotherapy or combination therapy with other drugs for mCRPC, belong to four types: olaparib, niraparib, rucaparib, and talazoparib. Veliparib is currently undergoing clinical trials [287,288]. Preclinical studies have indicated variations in the ability of PARPi to trap PARP enzymes across different tumor cells, including PCa cells, with talazoparib exhibiting the highest PARP-trapping capacity among the tested inhibitors, followed by niraparib, olaparib, and rucaparib [289]. In all trials involving PARPis, both radiographic PFS (rPFS) and OS were longer in patients with HRR mutations, specifically BRCA1/2 mutations, than in those without (Table 6) [288,290,291,292,293,294,295,296]. The TALAPRO-1 trial provided evidence of sustained antitumor activity of talazoparib in heavily pretreated patients with mCRPC and DDR-HRR mutations [297]. However, olaparib superiority was not observed in patients with alterations in other HRR genes in the PROfound trial (Table 2). Overall, although all PARPis share a similar mechanism of action, their differences in molecular structure and potency for trapping PARP enzymes may impact their clinical efficacy and potential use in different types of cancer.
Common adverse events reported with PARPi treatments include anemia and nausea. The incidence of anemia as an adverse event was similar across clinical trials of each PARPi, excluding veliparib (all-grade anemia: 46.0–65.8% and grade ≥ 3: 15.1–46.5%); however, the rate of anemia incidence was higher for talazoparib than for the others [288,291,292,293,294,295,296,298]. The rate of anemia incidence (all-grade and grade ≥ 3 anemia) due to veliparib was the lowest; however, an important consideration is that this trial had a small sample size (n = 74) [288]. The frequency of nausea as an adverse event varied across trials (all grade: 2.7–53.2% and grade ≥ 3: 0.5–4.2%); therefore, there may be specific characteristics associated with the occurrence of nausea due to PARPis (Table 7). Further research is required to elucidate the relationship between PARP-trapping potency and clinical toxicity profiles.

7.2. Akt Inhibitor for mCRPC with PTEN Deficiency

PTEN typically exerts tumor-suppressor effects by inhibiting the PI3K/Akt/mTOR signaling pathway. However, PTEN deficiency is present in approximately 50% of patients with mCRPC. Upon PTEN protein deficiency, signal transduction through PI3K and Akt occurs via genomic or non-genomic pathways, leading to the dysregulation of the cell cycle and cell proliferation, the enhancement of metabolic pathways such as those for glycolysis and protein translation, and the promotion of vascular neogenesis and cell survival pathways. These processes ultimately culminate in tumor formation and growth [299]. In patients with mCRPC, the PI3K/Akt/mTOR signaling pathway is likely activated as a mechanism for acquiring resistance to ARSIs [299]. Therefore, a treatment approach that concurrently inhibits the PI3K/Akt/mTOR and AR signaling pathways has been proposed to potentially improve the prognosis of mCRPC patients with PTEN protein deficiency [299,300].
Ipataseltib is a novel, cooperative, and selective ATP-competitive small-molecule inhibitor targeting all three isoforms of Akt [301]. A phase 3 trial (IPATentia150 trial: NCT03072238) evaluating the combination therapy of ipatasertib and apalutamide revealed that the median rPFS of mCRPC patients with PTEN deficiency was 18.5 months (95% CI: 16.3–2.1 months) in the ipatasertib + apalutamide group and 16.5 months (95% CI: 13.9–17.0 months) in the placebo + apalutamide group. This combination therapy significantly increased rPFS by 23% (HR: 0.77, 95% CI: 0.61–0.98, p = 0.0335) [302,303]. Akt inhibitors are considered potential novel agents representing the precision medicine era for mCRPC, and a phase 3 randomized CAPItello-280 trial (NCT05348577) is underway to evaluate the efficacy and safety of capivasertib and docetaxel administration as compared to that of placebo and docetaxel administration in the treatment of patients with mCRPC [304]. Furthermore, the randomized CAPItello-281 trial (NCT04493853) is underway to evaluate the efficacy of capivasertib, a potent selective inhibitor of all three Akt isoforms (AKT1/2/3), in combination with abiraterone in patients with mCSPC whose tumors are characterized by PTEN deficiency [305].

7.3. Immune Checkpoint Inhibitors (ICIs) for mCRPC

The programmed cell death 1 (PD-1)/programmed cell death ligand 1 (PD-L1) pathway is pivotal in regulating T-cell activity [306,307,308]. Several clinical studies have assessed the efficacy of ICIs, including anti-PD-L1 (nivolumab, atezolizumab, durvalumab, and avelumab), anti-PD1 (pembrolizumab and nivolumab), and anti-CTLA4 (ipilimumab) antibodies. Although preliminary studies have indicated limited anticancer activity of ICIs [309,310], patient selection based on deficiencies in MMR genes is considered crucial, as this patient cohort can potentially respond to ICIs. The FDA has approved the use of the anti-PD1 antibody pembrolizumab for treating cancers, including PCa with MMR mutations or MSI [311]. However, responses to ICIs in MMR-mutated or MSI PCa cases are not consistent.
Strong correlations were reported between immunohistochemical PD-L1 expression and the efficacy of PD-1 blockade in various malignant tumor cohorts. However, within these cohorts, only two patients had CRPC [312]; their samples revealed negative PD-L1 expression, suggesting the lack of a significant association. High PD-1 expression was observed in infiltrating T lymphocytes, specifically in CD8-positive prostate tissues, suggesting that they may not effectively mount an immune response [313]. A study involving 17 radical prostatectomy specimens investigated the expression of FoxP3, PD-1, and B7-H1 and revealed clusters of lymphocytes that were positive for all three proteins in the proximity of PCa lesions, collectively contributing to the negative regulation of the tumor immune response [139].
In one study, over 50% of patients with CRPC-associated, MSI-high tumors, or MMR mutants achieved a 50% reduction in PSA level after pembrolizumab treatment [118]. However, the rationale behind the lack of response to ICI therapy in some patients with PCa and MMR loss/MSI-high status remains unclear. Nonetheless, interest in combining ICIs with other therapies remains high. Concurrently, promising results were observed from a phase 1/2 clinical trial of pembrolizumab plus docetaxel, ARSIs, or PARP inhibitors in patients with mCRPC [314]. Furthermore, several phase 3 clinical trials evaluating the efficacy of pembrolizumab combined with docetaxel, enzalutamide, and olaparib are ongoing.
A phase 1/2 non-randomized trial (CA184-107) targeted patients with mCRPC who had not undergone more than one regimen of chemotherapy. In this trial, ipilimumab was administered as a single agent or in combination with external beam radiation therapy. The effectiveness of ipilimumab as monotherapy was reportedly between 14.3% and 16% in terms of the PSA50% response rate [315]. In another phase 3 randomized clinical trial (CA184-043) conducted with patients with mCRPC and bone metastases progressing after docetaxel chemotherapy, a single fraction of radiation therapy was administered to target bone metastases. Following this, patients were randomly assigned to receive either ipilimumab or a placebo [316]. While no significant difference in OS was observed between the ipilimumab and placebo groups (11.2 vs. 10 months, p = 0.053), the PFS was notably improved in the ipilimumab group compared with that in the placebo group (4 vs. 3.1 months, HR 0.70, p < 0.0001). Furthermore, the PSA response rate was significantly higher in the ipilimumab group than in the placebo group (13.1 vs. 5.2%). These findings highlight the potential benefits of ipilimumab in the treatment of patients with mCRPC and bone metastases following docetaxel chemotherapy [317,318]. Immunotherapy with ipilimumab is expected to be more effective in earlier stages of CRPC with minimal disease progression. New trials involving ipilimumab, either as a standalone treatment or in combination, are currently underway in chemotherapy-naïve PCa (NCT01057810 CA184-095) and neoadjuvant settings (NCT01194271) as phase 3 trials.
Notably, up to 10% of patients with mCRPC present with CDK12 aberration, which is associated with the response to ICIs [95]. CDK12, which forms a complex with cyclin K, is critical for DNA repair during gene translation; consequently, its inactivation leads to focal tandem duplications that promote gene fusions or mutations, thus enhancing neoantigen production and the tumor immune response [95]. One-half (2/4) of patients with mCRPC-causing CDK12 mutations showed PSA responses after administration of a PD1 inhibitor [95]. A phase 2 clinical trial to evaluate the efficacy of ipilimumab and nivolumab for CDK12-mutated mCRPC is ongoing [319]. Notably, CDK12 is mutated in 5% of patients with mCSPC and has been associated with improved responses to ICIs [121,320,321].
Defective MMR is associated with a higher tumor grade, a greater likelihood of developing de novo metastatic PCa, and a poorer response to ARSIs [322]. The identification of MMR-deficient PCa may have immediate clinical relevance considering recent data suggesting the sensitivity of such tumors to immune checkpoint blockade in other malignancies [323,324,325,326]. Favorable responses to ICIs such as pembrolizumab have been reported in MMR-deficient, MMR-mutated, MSI-high, or high-tumor-mutational-burden mCRPC [327,328,329,330,331,332,333]. However, prospective trials comparing ICIs to standard systemic therapy in MMR-deficient PCa are currently lacking [334,335].
An examination of PD-L1 expression in PCa cells and PD-1 expression in TILs among 16 patients with CRPC revealed high PD-1/PD-L1 expression in three patients. However, no TIL infiltration was observed in four cases, and PD-L1 expression was not detected in four of the eight cases [336]. The examination of PD-L1 expression (using clone EPR1161) in 873 hormone therapy-naïve radical prostatectomy specimens confirmed its expression in over 50% of the cases. Furthermore, high PD-L1 expression was significantly associated with cell proliferation, as evidenced by Ki-67, AR expression, and, notably, high PSA recurrence rates [337]. Collectively, poorer prognosis in cases with high PD-L1 expression in hormone therapy-naïve patients indicates the potential for improving outcomes in patients with PCa at a higher risk of recurrence, even before reaching the CRPC stage, by targeting the tumor immune response through PD-1/PD-L1 control. The analysis of blood samples and enzalutamide-resistant PCa cell lines in patients with CRPC who developed resistance to enzalutamide revealed increased numbers of PD-L1/L2-positive circulating dendritic cells in the bloodstream of enzalutamide-resistant patients with CRPC. Additionally, a pronounced increase in PD-L1 expression was observed in enzalutamide-resistant PCa cell lines derived from LNCap [338]. Many monotherapies and combination therapies with anti-PD-1, anti-PD-L1, and anti-CTLA-4 antibodies have not shown promising therapeutic effects or gained approval for PCa; nonetheless, significant effectiveness has been observed in some cases. Therefore, establishing biomarkers that can predict such effectiveness could help establish these therapies as a promising option. Additionally, combination therapy trials using ICIs and various treatments are currently underway and hold the promise of establishing effective treatment methods (Table 8, Table 9 and Table 10).

8. Conclusions

PCa is characterized by its androgen dependency, with treatment primarily based on androgen signal deprivation. However, the involvement of driver genes in both the initiation and progression of PCa is highly diverse among cases. Precision medicine, which involves understanding individual genetic alterations and proposing tailored treatment strategies, is considered crucial in the context of PCa.
The development of treatments based on genetic mutations such as PARPis for HRR gene mutations, AKT inhibitors for patients with genetic mutations in the PI3K/Akt pathway, and ICI for mutations in MMR genes and cases of MSI-high or high-tumor-mutational-burden CSPC and CRPC is anticipated. Monitoring the treatment progress of patients with PCa using liquid biopsy methods such as miRNA in blood or urine is a promising field that could significantly impact routine PCa treatment.

Author Contributions

Conceptualization, S.M., R.T. and W.O.; Data curation, S.M.; Writing—original draft preparation, S.M.; Writing—review and editing, S.M.; Supervision, W.O.; Project administration, S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data can be found within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. Center, M.M.; Jemal, A.; Lortet-Tieulent, J.; Ward, E.; Ferlay, J.; Brawley, O.; Bray, F. International variation in prostate cancer incidence and mortality rates. Eur. Urol. 2012, 61, 1079–1092. [Google Scholar] [CrossRef]
  3. Wang, L.; Lu, B.; He, M.; Wang, Y.; Wang, Z.; Du, L. Prostate Cancer Incidence and Mortality: Global Status and Temporal Trends in 89 Countries from 2000 to 2019. Front. Public Health 2022, 10, 811044. [Google Scholar] [CrossRef]
  4. Gandaglia, G.; Karakiewicz, P.I.; Briganti, A.; Passoni, N.M.; Schiffmann, J.; Trudeau, V.; Graefen, M.; Montorsi, F.; Sun, M. Impact of the Site of Metastases on Survival in Patients with Metastatic Prostate Cancer. Eur. Urol. 2015, 68, 325–334. [Google Scholar] [CrossRef]
  5. Vassilev, Z.P.; Gabarró, M.S.; Kaye, J.A.; Saltus, C.W.; Riedel, O.; Scholle, O.; Mehtälä, J.; Korhonen, P.; Garbe, E.; Zong, J. Incidence of second primary malignancies in metastatic castration-resistant prostate cancer: Results from observational studies in three countries. Future Oncol. 2020, 16, 1889–1901. [Google Scholar] [CrossRef]
  6. Ryan, C.J.; Smith, M.R.; Fizazi, K.; Saad, F.; Mulders, P.F.; Sternberg, C.N.; Miller, K.; Logothetis, C.J.; Shore, N.D.; Small, E.J.; et al. Abiraterone acetate plus prednisone versus placebo plus prednisone in chemotherapy-naive men with metastatic castration-resistant prostate cancer (COU-AA-302): Final overall survival analysis of a randomised, double-blind, placebo-controlled phase 3 study. Lancet Oncol. 2015, 16, 152–160. [Google Scholar] [CrossRef]
  7. Beer, T.M.; Armstrong, A.J.; Rathkopf, D.; Loriot, Y.; Sternberg, C.N.; Higano, C.S.; Iversen, P.; Evans, C.P.; Kim, C.S.; Kimura, G.; et al. Enzalutamide in Men with Chemotherapy-naïve Metastatic Castration-resistant Prostate Cancer: Extended Analysis of the Phase 3 PREVAIL Study. Eur. Urol. 2017, 71, 151–154. [Google Scholar] [CrossRef] [PubMed]
  8. Uemura, H.; Oya, M.; Kamoto, T.; Sugimoto, M.; Shinozaki, K.; Morita, K.; Koto, R.; Takahashi, M.; Nii, M.; Shin, E.; et al. The prevalence of gene mutations in homologous recombination repair pathways in Japanese patients with metastatic castration-resistant prostate cancer in real-world clinical practice: The multi-institutional observational ZENSHIN study. Cancer Med. 2022, 12, 5265–5274. [Google Scholar] [CrossRef] [PubMed]
  9. Feng, F.Y.; Huang, H.-C.; Spratt, D.E.; Zhao, S.; Sandler, H.M.; Simko, J.P.; Davicioni, E.; Nguyen, P.L.; Pollack, A.; Efstathiou, J.A.; et al. Validation of a 22-Gene Genomic Classifier in Patients with Recurrent Prostate Cancer: An Ancillary Study of the NRG/RTOG 9601 Randomized Clinical Trial. JAMA Oncol. 2021, 7, 544–552. [Google Scholar] [CrossRef]
  10. Cancer Genome Atlas Research Network. The Molecular Taxonomy of Primary Prostate Cancer. Cell 2015, 163, 1011–1025. [Google Scholar] [CrossRef]
  11. Baca, S.C.; Prandi, D.; Lawrence, M.S.; Mosquera, J.M.; Romanel, A.; Drier, Y.; Park, K.; Kitabayashi, N.; MacDonald, T.Y.; Ghandi, M.; et al. Punctuated evolution of prostate cancer genomes. Cell 2013, 153, 666–677. [Google Scholar] [CrossRef]
  12. Barbieri, C.E.; Baca, S.C.; Lawrence, M.S.; Demichelis, F.; Blattner, M.; Theurillat, J.P.; White, T.A.; Stojanov, P.; Van Allen, E.; Stransky, N.; et al. Exome sequencing identifies recurrent SPOP, FOXA1 and MED12 mutations in prostate cancer. Nat. Genet. 2012, 44, 685–689. [Google Scholar] [CrossRef] [PubMed]
  13. Berger, M.F.; Lawrence, M.S.; Demichelis, F.; Drier, Y.; Cibulskis, K.; Sivachenko, A.Y.; Sboner, A.; Esgueva, R.; Pflueger, D.; Sougnez, C.; et al. The genomic complexity of primary human prostate cancer. Nature 2011, 470, 214–220. [Google Scholar] [CrossRef] [PubMed]
  14. Cooper, C.S.; Eeles, R.; Wedge, D.C.; Van Loo, P.; Gundem, G.; Alexandrov, L.B.; Kremeyer, B.; Butler, A.; Lynch, A.G.; Camacho, N.; et al. Analysis of the genetic phylogeny of multifocal prostate cancer identifies multiple independent clonal expansions in neoplastic and morphologically normal prostate tissue. Nat. Genet. 2015, 47, 367–372. [Google Scholar] [CrossRef] [PubMed]
  15. Taylor, B.S.; Schultz, N.; Hieronymus, H.; Gopalan, A.; Xiao, Y.; Carver, B.S.; Arora, V.K.; Kaushik, P.; Cerami, E.; Reva, B.; et al. Integrative genomic profiling of human prostate cancer. Cancer Cell 2010, 18, 11–22. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, X.S.; Shankar, S.; Dhanasekaran, S.M.; Ateeq, B.; Sasaki, A.T.; Jing, X.; Robinson, D.; Cao, Q.; Prensner, J.R.; Yocum, A.K.; et al. Characterization of KRAS rearrangements in metastatic prostate cancer. Cancer Discov. 2011, 1, 35–43. [Google Scholar] [CrossRef] [PubMed]
  17. Tomlins, S.A.; Laxman, B.; Dhanasekaran, S.M.; Helgeson, B.E.; Cao, X.; Morris, D.S.; Menon, A.; Jing, X.; Cao, Q.; Han, B.; et al. Distinct classes of chromosomal rearrangements create oncogenic ETS gene fusions in prostate cancer. Nature 2007, 448, 595–599. [Google Scholar] [CrossRef]
  18. Kohli, M.; Tan, W.; Zheng, T.; Wang, A.; Montesinos, C.; Wong, C.; Du, P.; Jia, S.; Yadav, S.; Horvath, L.G.; et al. Clinical and genomic insights into circulating tumor DNA-based alterations across the spectrum of metastatic hormone-sensitive and castrate-resistant prostate cancer. eBioMedicine 2020, 54, 102728. [Google Scholar] [CrossRef] [PubMed]
  19. Stopsack, K.H.; Nandakumar, S.; Wibmer, A.G.; Haywood, S.; Weg, E.S.; Barnett, E.S.; Kim, C.J.; Carbone, E.A.; Vasselman, S.E.; Nguyen, B.; et al. Oncogenic Genomic Alterations, Clinical Phenotypes, and Outcomes in Metastatic Castration-Sensitive Prostate Cancer. Clin. Cancer Res. 2020, 26, 3230–3238. [Google Scholar] [CrossRef]
  20. Mateo, J.; Seed, G.; Bertan, C.; Rescigno, P.; Dolling, D.; Figueiredo, I.; Miranda, S.; Nava Rodrigues, D.; Gurel, B.; Clarke, M.; et al. Genomics of lethal prostate cancer at diagnosis and castration resistance. J. Clin. Investig. 2020, 130, 1743–1751. [Google Scholar] [CrossRef]
  21. Hamid, A.A.; Gray, K.P.; Shaw, G.; MacConaill, L.E.; Evan, C.; Bernard, B.; Loda, M.; Corcoran, N.M.; Van Allen, E.M.; Choudhury, A.D.; et al. Compound Genomic Alterations of TP53, PTEN, and RB1 Tumor Suppressors in Localized and Metastatic Prostate Cancer. Eur. Urol. 2019, 76, 89–97. [Google Scholar] [CrossRef]
  22. Deek, M.P.; Van der Eecken, K.; Phillips, R.; Parikh, N.R.; Isaacsson Velho, P.; Lotan, T.L.; Kishan, A.U.; Maurer, T.; Boutros, P.C.; Hovens, C.; et al. The Mutational Landscape of Metastatic Castration-sensitive Prostate Cancer: The Spectrum Theory Revisited. Eur. Urol. 2021, 80, 632–640. [Google Scholar] [CrossRef]
  23. Swami, U.; Isaacsson Velho, P.; Nussenzveig, R.; Chipman, J.; Sacristan Santos, V.; Erickson, S.; Dharmaraj, D.; Alva, A.S.; Vaishampayan, U.N.; Esther, J.; et al. Association of SPOP Mutations with Outcomes in Men with De Novo Metastatic Castration-sensitive Prostate Cancer. Eur. Urol. 2020, 78, 652–656. [Google Scholar] [CrossRef] [PubMed]
  24. Annala, M.; Taavitsainen, S.; Khalaf, D.J.; Vandekerkhove, G.; Beja, K.; Sipola, J.; Warner, E.W.; Herberts, C.; Wong, A.; Fu, S.; et al. Evolution of Castration-Resistant Prostate Cancer in ctDNA during Sequential Androgen Receptor Pathway Inhibition. Clin. Cancer Res. 2021, 27, 4610–4623. [Google Scholar] [CrossRef] [PubMed]
  25. Gundem, G.; Van Loo, P.; Kremeyer, B.; Alexandrov, L.B.; Tubio, J.M.C.; Papaemmanuil, E.; Brewer, D.S.; Kallio, H.M.L.; Högnäs, G.; Annala, M.; et al. The evolutionary history of lethal metastatic prostate cancer. Nature 2015, 520, 353–357. [Google Scholar] [CrossRef] [PubMed]
  26. Boysen, G.; Barbieri, C.E.; Prandi, D.; Blattner, M.; Chae, S.S.; Dahija, A.; Nataraj, S.; Huang, D.; Marotz, C.; Xu, L.; et al. SPOP mutation leads to genomic instability in prostate cancer. eLife 2015, 4, e09207. [Google Scholar] [CrossRef] [PubMed]
  27. Finn, R.S.; Crown, J.P.; Lang, I.; Boer, K.; Bondarenko, I.M.; Kulyk, S.O.; Ettl, J.; Patel, R.; Pinter, T.; Schmidt, M.; et al. The cyclin-dependent kinase 4/6 inhibitor palbociclib in combination with letrozole versus letrozole alone as first-line treatment of oestrogen receptor-positive, HER2-negative, advanced breast cancer (PALOMA-1/TRIO-18): A randomised phase 2 study. Lancet Oncol. 2015, 16, 25–35. [Google Scholar] [CrossRef] [PubMed]
  28. Abida, W.; Armenia, J.; Gopalan, A.; Brennan, R.; Walsh, M.; Barron, D.; Danila, D.; Rathkopf, D.; Morris, M.; Slovin, S.; et al. Prospective Genomic Profiling of Prostate Cancer Across Disease States Reveals Germline and Somatic Alterations That May Affect Clinical Decision Making. JCO Precis. Oncol. 2017, 2017, 1–16. [Google Scholar] [CrossRef] [PubMed]
  29. Davies, A.; Azad, A.A.; Kwan, E.M. Genomic Alterations to Guide Treatment Selection in Metastatic Prostate Cancer. Crit. Rev. Oncog. 2022, 27, 61–80. [Google Scholar] [CrossRef] [PubMed]
  30. He, Y.; Xu, W.; Xiao, Y.T.; Huang, H.; Gu, D.; Ren, S. Targeting signaling pathways in prostate cancer: Mechanisms and clinical trials. Signal Transduct. Target. Ther. 2022, 7, 198. [Google Scholar] [CrossRef]
  31. Gilson, C.; Ingleby, F.; Gilbert, D.C.; Parry, M.A.; Atako, N.B.; Ali, A.; Hoyle, A.; Clarke, N.W.; Gannon, M.; Wanstall, C.; et al. Genomic Profiles of De Novo High- and Low-Volume Metastatic Prostate Cancer: Results from a 2-Stage Feasibility and Prevalence Study in the STAMPEDE Trial. JCO Precis. Oncol. 2020, 4, 882–897. [Google Scholar] [CrossRef]
  32. Vandekerkhove, G.; Struss, W.J.; Annala, M.; Kallio, H.M.L.; Khalaf, D.; Warner, E.W.; Herberts, C.; Ritch, E.; Beja, K.; Loktionova, Y.; et al. Circulating Tumor DNA Abundance and Potential Utility in De Novo Metastatic Prostate Cancer. Eur. Urol. 2019, 75, 667–675. [Google Scholar] [CrossRef]
  33. Shenderov, E.; Isaacsson Velho, P.; Awan, A.H.; Wang, H.; Mirkheshti, N.; Lotan, T.L.; Carducci, M.A.; Pardoll, D.M.; Eisenberger, M.A.; Antonarakis, E.S. Genomic and clinical characterization of pulmonary-only metastatic prostate cancer: A unique molecular subtype. Prostate 2019, 79, 1572–1579. [Google Scholar] [CrossRef]
  34. Fan, L.; Fei, X.; Zhu, Y.; Pan, J.; Sha, J.; Chi, C.; Gong, Y.; Du, X.; Zhou, L.; Dong, B.; et al. Comparative Analysis of Genomic Alterations across Castration Sensitive and Castration Resistant Prostate Cancer via Circulating Tumor DNA Sequencing. J. Urol. 2021, 205, 461–469. [Google Scholar] [CrossRef]
  35. Van der Eecken, K.; Vanwelkenhuyzen, J.; Deek, M.P.; Tran, P.T.; Warner, E.; Wyatt, A.W.; Kwan, E.M.; Verbeke, S.; Van Dorpe, J.; Fonteyne, V.; et al. Tissue- and Blood-derived Genomic Biomarkers for Metastatic Hormone-sensitive Prostate Cancer: A Systematic Review. Eur. Urol. Oncol. 2021, 4, 914–923. [Google Scholar] [CrossRef] [PubMed]
  36. 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]
  37. Li, J.; Xu, C.; Lee, H.J.; Ren, S.; Zi, X.; Zhang, Z.; Wang, H.; Yu, Y.; Yang, C.; Gao, X.; et al. A genomic and epigenomic atlas of prostate cancer in Asian populations. Nature 2020, 580, 93–99. [Google Scholar] [CrossRef] [PubMed]
  38. Sboner, A.; Habegger, L.; Pflueger, D.; Terry, S.; Chen, D.Z.; Rozowsky, J.S.; Tewari, A.K.; Kitabayashi, N.; Moss, B.J.; Chee, M.S.; et al. FusionSeq: A modular framework for finding gene fusions by analyzing paired-end RNA-sequencing data. Genome Biol. 2010, 11, R104. [Google Scholar] [CrossRef] [PubMed]
  39. Wang, K.; Singh, D.; Zeng, Z.; Coleman, S.J.; Huang, Y.; Savich, G.L.; He, X.; Mieczkowski, P.; Grimm, S.A.; Perou, C.M.; et al. MapSplice: Accurate mapping of RNA-seq reads for splice junction discovery. Nucleic Acids Res. 2010, 38, e178. [Google Scholar] [CrossRef] [PubMed]
  40. Tomlins, S.A.; Rhodes, D.R.; Perner, S.; Dhanasekaran, S.M.; Mehra, R.; Sun, X.W.; Varambally, S.; Cao, X.; Tchinda, J.; Kuefer, R.; et al. Recurrent fusion of TMPRSS2 and ETS transcription factor genes in prostate cancer. Science 2005, 310, 644–648. [Google Scholar] [CrossRef] [PubMed]
  41. Tomlins, S.A.; Bjartell, A.; Chinnaiyan, A.M.; Jenster, G.; Nam, R.K.; Rubin, M.A.; Schalken, J.A. ETS gene fusions in prostate cancer: From discovery to daily clinical practice. Eur. Urol. 2009, 56, 275–286. [Google Scholar] [CrossRef]
  42. Magi-Galluzzi, C.; Tsusuki, T.; Elson, P.; Simmerman, K.; LaFargue, C.; Esgueva, R.; Klein, E.; Rubin, M.A.; Zhou, M. TMPRSS2-ERG gene fusion prevalence and class are significantly different in prostate cancer of Caucasian, African-American and Japanese patients. Prostate 2011, 71, 489–497. [Google Scholar] [CrossRef]
  43. Miyagi, Y.; Sasaki, T.; Fujinami, K.; Sano, J.; Senga, Y.; Miura, T.; Kameda, Y.; Sakuma, Y.; Nakamura, Y.; Harada, M.; et al. ETS family-associated gene fusions in Japanese prostate cancer: Analysis of 194 radical prostatectomy samples. Modern Pathol. 2010, 23, 1492–1498. [Google Scholar] [CrossRef]
  44. Maekawa, S.; Suzuki, M.; Arai, T.; Suzuki, M.; Kato, M.; Morikawa, T.; Kasuya, Y.; Kume, H.; Kitamura, T.; Homma, Y. TMPRSS2 Met160Val polymorphism: Significant association with sporadic prostate cancer, but not with latent prostate cancer in Japanese men. Int. J. Urol. 2014, 21, 1234–1238. [Google Scholar] [CrossRef]
  45. Takata, R.; Akamatsu, S.; Kubo, M.; Takahashi, A.; Hosono, N.; Kawaguchi, T.; Tsunoda, T.; Inazawa, J.; Kamatani, N.; Ogawa, O.; et al. Genome-wide association study identifies five new susceptibility loci for prostate cancer in the Japanese population. Nat. Genet. 2010, 42, 751–754. [Google Scholar] [CrossRef] [PubMed]
  46. Cheng, I.; Chen, G.K.; Nakagawa, H.; He, J.; Wan, P.; Laurie, C.C.; Shen, J.; Sheng, X.; Pooler, L.C.; Crenshaw, A.T.; et al. Evaluating genetic risk for prostate cancer among Japanese and Latinos. Cancer Epidemiol. Biomark. Prev. 2012, 21, 2048–2058. [Google Scholar] [CrossRef] [PubMed]
  47. Fraser, M.; Sabelnykova, V.Y.; Yamaguchi, T.N.; Heisler, L.E.; Livingstone, J.; Huang, V.; Shiah, Y.J.; Yousif, F.; Lin, X.; Masella, A.P.; et al. Genomic hallmarks of localized, non-indolent prostate cancer. Nature 2017, 541, 359–364. [Google Scholar] [CrossRef] [PubMed]
  48. Nam, R.K.; Sugar, L.; Yang, W.; Srivastava, S.; Klotz, L.H.; Yang, L.Y.; Stanimirovic, A.; Encioiu, E.; Neill, M.; Loblaw, D.A.; et al. Expression of the TMPRSS2:ERG fusion gene predicts cancer recurrence after surgery for localised prostate cancer. Br. J. Cancer 2007, 97, 1690–1695. [Google Scholar] [CrossRef] [PubMed]
  49. Rajput, A.B.; Miller, M.A.; De Luca, A.; Boyd, N.; Leung, S.; Hurtado-Coll, A.; Fazli, L.; Jones, E.C.; Palmer, J.B.; Gleave, M.E.; et al. Frequency of the TMPRSS2:ERG gene fusion is increased in moderate to poorly differentiated prostate cancers. J. Clin. Pathol. 2007, 60, 1238–1243. [Google Scholar] [CrossRef] [PubMed]
  50. Takeda, D.Y.; Spisák, S.; Seo, J.H.; Bell, C.; O’Connor, E.; Korthauer, K.; Ribli, D.; Csabai, I.; Solymosi, N.; Szállási, Z.; et al. A Somatically Acquired Enhancer of the Androgen Receptor Is a Noncoding Driver in Advanced Prostate Cancer. Cell 2018, 174, 422–432.e13. [Google Scholar] [CrossRef] [PubMed]
  51. 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] [PubMed]
  52. 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]
  53. Annala, M.; Vandekerkhove, G.; Khalaf, D.; Taavitsainen, S.; Beja, K.; Warner, E.W.; Sunderland, K.; Kollmannsberger, C.; Eigl, B.J.; Finch, D.; et al. Circulating Tumor DNA Genomics Correlate with Resistance to Abiraterone and Enzalutamide in Prostate Cancer. Cancer Discov. 2018, 8, 444–457. [Google Scholar] [CrossRef] [PubMed]
  54. Conteduca, V.; Jayaram, A.; Romero-Laorden, N.; Wetterskog, D.; Salvi, S.; Gurioli, G.; Scarpi, E.; Castro, E.; Marin-Aguilera, M.; Lolli, C.; et al. Plasma Androgen Receptor and Docetaxel for Metastatic Castration-resistant Prostate Cancer. Eur. Urol. 2019, 75, 368–373. [Google Scholar] [CrossRef] [PubMed]
  55. Conteduca, V.; Castro, E.; Wetterskog, D.; Scarpi, E.; Jayaram, A.; Romero-Laorden, N.; Olmos, D.; Gurioli, G.; Lolli, C.; Sáez, M.I.; et al. Plasma AR status and cabazitaxel in heavily treated metastatic castration-resistant prostate cancer. Eur. J. Cancer 2019, 116, 158–168. [Google Scholar] [CrossRef] [PubMed]
  56. Antonarakis, E.S.; Lu, C.; Wang, H.; Luber, B.; Nakazawa, M.; Roeser, J.C.; Chen, Y.; Mohammad, T.A.; Chen, Y.; Fedor, H.L.; et al. AR-V7 and resistance to enzalutamide and abiraterone in prostate cancer. N. Engl. J. Med. 2014, 371, 1028–1038. [Google Scholar] [CrossRef] [PubMed]
  57. Antonarakis, E.S.; Lu, C.; Luber, B.; Wang, H.; Chen, Y.; Nakazawa, M.; Nadal, R.; Paller, C.J.; Denmeade, S.R.; Carducci, M.A.; et al. Androgen Receptor Splice Variant 7 and Efficacy of Taxane Chemotherapy in Patients with Metastatic Castration-Resistant Prostate Cancer. JAMA Oncol. 2015, 1, 582–591. [Google Scholar] [CrossRef]
  58. Azad, A.A.; Volik, S.V.; Wyatt, A.W.; Haegert, A.; Le Bihan, S.; Bell, R.H.; Anderson, S.A.; McConeghy, B.; Shukin, R.; Bazov, J.; et al. Androgen Receptor Gene Aberrations in Circulating Cell-Free DNA: Biomarkers of Therapeutic Resistance in Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2015, 21, 2315–2324. [Google Scholar] [CrossRef]
  59. Wyatt, A.W.; Azad, A.A.; Volik, S.V.; Annala, M.; Beja, K.; McConeghy, B.; Haegert, A.; Warner, E.W.; Mo, F.; Brahmbhatt, S.; et al. Genomic Alterations in Cell-Free DNA and Enzalutamide Resistance in Castration-Resistant Prostate Cancer. JAMA Oncol. 2016, 2, 1598–1606. [Google Scholar] [CrossRef]
  60. Conteduca, V.; Wetterskog, D.; Sharabiani, M.T.A.; Grande, E.; Fernandez-Perez, M.P.; Jayaram, A.; Salvi, S.; Castellano, D.; Romanel, A.; Lolli, C.; et al. Androgen receptor gene status in plasma DNA associates with worse outcome on enzalutamide or abiraterone for castration-resistant prostate cancer: A multi-institution correlative biomarker study. Ann. Oncol. 2017, 28, 1508–1516. [Google Scholar] [CrossRef]
  61. Hieronymus, H.; Lamb, J.; Ross, K.N.; Peng, X.P.; Clement, C.; Rodina, A.; Nieto, M.; Du, J.; Stegmaier, K.; Raj, S.M.; et al. Gene expression signature-based chemical genomic prediction identifies a novel class of HSP90 pathway modulators. Cancer Cell 2006, 10, 321–330. [Google Scholar] [CrossRef] [PubMed]
  62. Zhuang, M.; Calabrese, M.F.; Liu, J.; Waddell, M.B.; Nourse, A.; Hammel, M.; Miller, D.J.; Walden, H.; Duda, D.M.; Seyedin, S.N.; et al. Structures of SPOP-substrate complexes: Insights into molecular architectures of BTB-Cul3 ubiquitin ligases. Mol. Cell 2009, 36, 39–50. [Google Scholar] [CrossRef] [PubMed]
  63. Gan, W.; Dai, X.; Lunardi, A.; Li, Z.; Inuzuka, H.; Liu, P.; Varmeh, S.; Zhang, J.; Cheng, L.; Sun, Y.; et al. SPOP Promotes Ubiquitination and Degradation of the ERG Oncoprotein to Suppress Prostate Cancer Progression. Mol. Cell 2015, 59, 917–930. [Google Scholar] [CrossRef] [PubMed]
  64. An, J.; Wang, C.; Deng, Y.; Yu, L.; Huang, H. Destruction of full-length androgen receptor by wild-type SPOP, but not prostate-cancer-associated mutants. Cell Rep. 2014, 6, 657–669. [Google Scholar] [CrossRef] [PubMed]
  65. Boysen, G.; Rodrigues, D.N.; Rescigno, P.; Seed, G.; Dolling, D.; Riisnaes, R.; Crespo, M.; Zafeiriou, Z.; Sumanasuriya, S.; Bianchini, D.; et al. SPOP-Mutated/CHD1-Deleted Lethal Prostate Cancer and Abiraterone Sensitivity. Clin. Cancer Res. 2018, 24, 5585–5593. [Google Scholar] [CrossRef] [PubMed]
  66. Zhou, Z.; Flesken-Nikitin, A.; Corney, D.C.; Wang, W.; Goodrich, D.W.; Roy-Burman, P.; Nikitin, A.Y. Synergy of p53 and Rb deficiency in a conditional mouse model for metastatic prostate cancer. Cancer Res. 2006, 66, 7889–7898. [Google Scholar] [CrossRef]
  67. Mu, P.; Zhang, Z.; Benelli, M.; Karthaus, W.R.; Hoover, E.; Chen, C.C.; Wongvipat, J.; Ku, S.Y.; Gao, D.; Cao, Z.; et al. SOX2 promotes lineage plasticity and antiandrogen resistance in TP53- and RB1-deficient prostate cancer. Science 2017, 355, 84–88. [Google Scholar] [CrossRef]
  68. Aparicio, A.M.; Shen, L.; Tapia, E.L.; Lu, J.F.; Chen, H.C.; Zhang, J.; Wu, G.; Wang, X.; Troncoso, P.; Corn, P.; et al. Combined Tumor Suppressor Defects Characterize Clinically Defined Aggressive Variant Prostate Cancers. Clin. Cancer Res. 2016, 22, 1520–1530. [Google Scholar] [CrossRef]
  69. Pourmand, G.; Ziaee, A.A.; Abedi, A.R.; Mehrsai, A.; Alavi, H.A.; Ahmadi, A.; Saadati, H.R. Role of PTEN gene in progression of prostate cancer. Urol. J. 2007, 4, 95–100. [Google Scholar]
  70. Reid, A.H.; Attard, G.; Ambroisine, L.; Fisher, G.; Kovacs, G.; Brewer, D.; Clark, J.; Flohr, P.; Edwards, S.; Berney, D.M.; et al. Molecular characterisation of ERG, ETV1 and PTEN gene loci identifies patients at low and high risk of death from prostate cancer. Br. J. Cancer 2010, 102, 678–684. [Google Scholar] [CrossRef]
  71. Brognard, J.; Sierecki, E.; Gao, T.; Newton, A.C. PHLPP and a second isoform, PHLPP2, differentially attenuate the amplitude of Akt signaling by regulating distinct Akt isoforms. Mol. Cell 2007, 25, 917–931. [Google Scholar] [CrossRef] [PubMed]
  72. Mulholland, D.J.; Tran, L.M.; Li, Y.; Cai, H.; Morim, A.; Wang, S.; Plaisier, S.; Garraway, I.P.; Huang, J.; Graeber, T.G.; et al. Cell autonomous role of PTEN in regulating castration-resistant prostate cancer growth. Cancer Cell 2011, 19, 792–804. [Google Scholar] [CrossRef] [PubMed]
  73. Gasparian, A.V.; Yao, Y.J.; Kowalczyk, D.; Lyakh, L.A.; Karseladze, A.; Slaga, T.J.; Budunova, I.V. The role of IKK in constitutive activation of NF-kappaB transcription factor in prostate carcinoma cells. J. Cell Sci. 2002, 115, 141–151. [Google Scholar] [CrossRef] [PubMed]
  74. Ferraldeschi, R.; Nava Rodrigues, D.; Riisnaes, R.; Miranda, S.; Figueiredo, I.; Rescigno, P.; Ravi, P.; Pezaro, C.; Omlin, A.; Lorente, D.; et al. PTEN protein loss and clinical outcome from castration-resistant prostate cancer treated with abiraterone acetate. Eur. Urol. 2015, 67, 795–802. [Google Scholar] [CrossRef] [PubMed]
  75. Carver, B.S.; Chapinski, C.; Wongvipat, J.; Hieronymus, H.; Chen, Y.; Chandarlapaty, S.; Arora, V.K.; Le, C.; Koutcher, J.; Scher, H.; et al. Reciprocal feedback regulation of PI3K and androgen receptor signaling in PTEN-deficient prostate cancer. Cancer Cell 2011, 19, 575–586. [Google Scholar] [CrossRef] [PubMed]
  76. 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]
  77. Prandi, D.; Baca, S.C.; Romanel, A.; Barbieri, C.E.; Mosquera, J.M.; Fontugne, J.; Beltran, H.; Sboner, A.; Garraway, L.A.; Rubin, M.A.; et al. Unraveling the clonal hierarchy of somatic genomic aberrations. Genome Biol. 2014, 15, 439. [Google Scholar] [CrossRef]
  78. Jamaspishvili, T.; Berman, D.M.; Ross, A.E.; Scher, H.I.; De Marzo, A.M.; Squire, J.A.; Lotan, T.L. Clinical implications of PTEN loss in prostate cancer. Nat. Rev. Urol. 2018, 15, 222–234. [Google Scholar] [CrossRef]
  79. Hayden, M.S.; Ghosh, S. Shared principles in NF-kappaB signaling. Cell 2008, 132, 344–362. [Google Scholar] [CrossRef]
  80. Perkins, N.D. Post-translational modifications regulating the activity and function of the nuclear factor kappa B pathway. Oncogene 2006, 25, 6717–6730. [Google Scholar] [CrossRef]
  81. Sun, S.C. The noncanonical NF-κB pathway. Immunol. Rev. 2012, 246, 125–140. [Google Scholar] [CrossRef] [PubMed]
  82. Gilmore, T.D. The Rel/NF-kappaB signal transduction pathway: Introduction. Oncogene 1999, 18, 6842–6844. [Google Scholar] [CrossRef] [PubMed]
  83. Li, Q.; Verma, I.M. NF-kappaB regulation in the immune system. Nat. Rev. Immunol. 2002, 2, 725–734. [Google Scholar] [CrossRef]
  84. Gilmore, T.D. Introduction to NF-kappaB: Players, pathways, perspectives. Oncogene 2006, 25, 6680–6684. [Google Scholar] [CrossRef] [PubMed]
  85. Vecchiotti, D.; Verzella, D.; Di Vito Nolfi, M.; D’Andrea, D.; Flati, I.; Di Francesco, B.; Cornice, J.; Alesse, E.; Capece, D.; Zazzeroni, F. Elevated NF-κB/SHh/GLI1 Signature Denotes a Worse Prognosis and Represent a Novel Potential Therapeutic Target in Advanced Prostate Cancer. Cells 2022, 11, 2118. [Google Scholar] [CrossRef] [PubMed]
  86. Thomas-Jardin, S.E.; Dahl, H.; Nawas, A.F.; Bautista, M.; Delk, N.A. NF-κB signaling promotes castration-resistant prostate cancer initiation and progression. Pharmacol. Ther. 2020, 211, 107538. [Google Scholar] [CrossRef]
  87. Wang, X.; Fang, Y.; Sun, W.; Xu, Z.; Zhang, Y.; Wei, X.; Ding, X.; Xu, Y. Endocrinotherapy resistance of prostate and breast cancer: Importance of the NF-κB pathway (Review). Int. J. Oncol. 2020, 56, 1064–1074. [Google Scholar] [CrossRef]
  88. Nguyen, D.P.; Li, J.; Yadav, S.S.; Tewari, A.K. Recent insights into NF-κB signalling pathways and the link between inflammation and prostate cancer. BJU Int. 2014, 114, 168–176. [Google Scholar] [CrossRef]
  89. Mu, H.Q.; He, Y.H.; Wang, S.B.; Yang, S.; Wang, Y.J.; Nan, C.J.; Bao, Y.F.; Xie, Q.P.; Chen, Y.H. MiR-130b/TNF-α/NF-κB/VEGFA loop inhibits prostate cancer angiogenesis. Clin. Transl. Oncol. 2020, 22, 111–121. [Google Scholar] [CrossRef]
  90. Zhou, J.; Li, J.; Qian, C.; Qiu, F.; Shen, Q.; Tong, R.; Yang, Q.; Xu, J.; Zheng, B.; Lv, J.; et al. LINC00624/TEX10/NF-κB axis promotes proliferation and migration of human prostate cancer cells. Biochem. Biophys. Res. Commun. 2022, 601, 1–8. [Google Scholar] [CrossRef] [PubMed]
  91. Shang, Z.; Yu, J.; Sun, L.; Tian, J.; Zhu, S.; Zhang, B.; Dong, Q.; Jiang, N.; Flores-Morales, A.; Chang, C.; et al. LncRNA PCAT1 activates AKT and NF-κB signaling in castration-resistant prostate cancer by regulating the PHLPP/FKBP51/IKKα complex. Nucleic Acids Res. 2019, 47, 4211–4225. [Google Scholar] [CrossRef]
  92. Comstock, C.E.; Augello, M.A.; Goodwin, J.F.; de Leeuw, R.; Schiewer, M.J.; Ostrander, W.F., Jr.; Burkhart, R.A.; McClendon, A.K.; McCue, P.A.; Trabulsi, E.J.; et al. Targeting cell cycle and hormone receptor pathways in cancer. Oncogene 2013, 32, 5481–5491. [Google Scholar] [CrossRef]
  93. Chen, Z.; Trotman, L.C.; Shaffer, D.; Lin, H.K.; Dotan, Z.A.; Niki, M.; Koutcher, J.A.; Scher, H.I.; Ludwig, T.; Gerald, W.; et al. Crucial role of p53-dependent cellular senescence in suppression of Pten-deficient tumorigenesis. Nature 2005, 436, 725–730. [Google Scholar] [CrossRef]
  94. Hong, M.K.; Macintyre, G.; Wedge, D.C.; Van Loo, P.; Patel, K.; Lunke, S.; Alexandrov, L.B.; Sloggett, C.; Cmero, M.; Marass, F.; et al. Tracking the origins and drivers of subclonal metastatic expansion in prostate cancer. Nat. Commun. 2015, 6, 6605. [Google Scholar] [CrossRef]
  95. Wu, Y.M.; Cieślik, M.; Lonigro, R.J.; Vats, P.; Reimers, M.A.; Cao, X.; Ning, Y.; Wang, L.; Kunju, L.P.; de Sarkar, N.; et al. Inactivation of CDK12 Delineates a Distinct Immunogenic Class of Advanced Prostate Cancer. Cell 2018, 173, 1770–1782.e14. [Google Scholar] [CrossRef]
  96. Mateo, J.; Carreira, S.; Sandhu, S.; Miranda, S.; Mossop, H.; Perez-Lopez, R.; Nava Rodrigues, D.; Robinson, D.; Omlin, A.; Tunariu, N.; et al. DNA-Repair Defects and Olaparib in Metastatic Prostate Cancer. N. Engl. J. Med. 2015, 373, 1697–1708. [Google Scholar] [CrossRef] [PubMed]
  97. Neeb, A.; Herranz, N.; Arce-Gallego, S.; Miranda, S.; Buroni, L.; Yuan, W.; Athie, A.; Casals, T.; Carmichael, J.; Rodrigues, D.N.; et al. Advanced Prostate Cancer with ATM Loss: PARP and ATR Inhibitors. Eur. Urol. 2021, 79, 200–211. [Google Scholar] [CrossRef] [PubMed]
  98. Wu, Y.; Yu, H.; Li, S.; Wiley, K.; Zheng, S.L.; LaDuca, H.; Gielzak, M.; Na, R.; Sarver, B.A.J.; Helfand, B.T.; et al. Rare Germline Pathogenic Mutations of DNA Repair Genes Are Most Strongly Associated with Grade Group 5 Prostate Cancer. Eur. Urol. Oncol. 2020, 3, 224–230. [Google Scholar] [CrossRef] [PubMed]
  99. Na, R.; Zheng, S.L.; Han, M.; Yu, H.; Jiang, D.; Shah, S.; Ewing, C.M.; Zhang, L.; Novakovic, K.; Petkewicz, J.; et al. Germline Mutations in ATM and BRCA1/2 Distinguish Risk for Lethal and Indolent Prostate Cancer and are Associated with Early Age at Death. Eur. Urol. 2017, 71, 740–747. [Google Scholar] [CrossRef]
  100. de Bono, J.S.; De Giorgi, U.; Rodrigues, D.N.; Massard, C.; Bracarda, S.; Font, A.; Arranz Arija, J.A.; Shih, K.C.; Radavoi, G.D.; Xu, N.; et al. Randomized Phase II Study Evaluating Akt Blockade with Ipatasertib, in Combination with Abiraterone, in Patients with Metastatic Prostate Cancer with and without PTEN Loss. Clin. Cancer Res. 2019, 25, 928–936. [Google Scholar] [CrossRef]
  101. Castro, E.; Goh, C.; Leongamornlert, D.; Saunders, E.; Tymrakiewicz, M.; Dadaev, T.; Govindasami, K.; Guy, M.; Ellis, S.; Frost, D.; et al. Effect of BRCA Mutations on Metastatic Relapse and Cause-specific Survival After Radical Treatment for Localised Prostate Cancer. Eur. Urol. 2015, 68, 186–193. [Google Scholar] [CrossRef]
  102. Kote-Jarai, Z.; Leongamornlert, D.; Saunders, E.; Tymrakiewicz, M.; Castro, E.; Mahmud, N.; Guy, M.; Edwards, S.; O’Brien, L.; Sawyer, E.; et al. BRCA2 is a moderate penetrance gene contributing to young-onset prostate cancer: Implications for genetic testing in prostate cancer patients. Br. J. Cancer 2011, 105, 1230–1234. [Google Scholar] [CrossRef]
  103. Jonsson, P.; Bandlamudi, C.; Cheng, M.L.; Srinivasan, P.; Chavan, S.S.; Friedman, N.D.; Rosen, E.Y.; Richards, A.L.; Bouvier, N.; Selcuklu, S.D.; et al. Tumour lineage shapes BRCA-mediated phenotypes. Nature 2019, 571, 576–579. [Google Scholar] [CrossRef]
  104. Taylor, R.A.; Fraser, M.; Livingstone, J.; Espiritu, S.M.; Thorne, H.; Huang, V.; Lo, W.; Shiah, Y.J.; Yamaguchi, T.N.; Sliwinski, A.; et al. Germline BRCA2 mutations drive prostate cancers with distinct evolutionary trajectories. Nat. Commun. 2017, 8, 13671. [Google Scholar] [CrossRef] [PubMed]
  105. Castro, E.; Goh, C.; Olmos, D.; Saunders, E.; Leongamornlert, D.; Tymrakiewicz, M.; Mahmud, N.; Dadaev, T.; Govindasami, K.; Guy, M.; et al. Germline BRCA mutations are associated with higher risk of nodal involvement, distant metastasis, and poor survival outcomes in prostate cancer. J. Clin. Oncol. 2013, 31, 1748–1757. [Google Scholar] [CrossRef] [PubMed]
  106. Maekawa, S.; Takata, R.; Sekiguchi, K.; Kagabu, M.; Toyoshima, M.; Tamada, S.; Takahashi, K.; Ikarashi, D.; Matsuura, T.; Kato, R.; et al. Prevalence of germline BRCA1/2 pathogenic variants in Japanese patients treated with castration-resistant prostate cancer and efficacy of CRPC treatment in real-world clinical practice. Jpn. J. Clin. Oncol. 2023, hyad185. [Google Scholar] [CrossRef] [PubMed]
  107. Risbridger, G.P.; Taylor, R.A.; Clouston, D.; Sliwinski, A.; Thorne, H.; Hunter, S.; Li, J.; Mitchell, G.; Murphy, D.; Frydenberg, M.; et al. Patient-derived xenografts reveal that intraductal carcinoma of the prostate is a prominent pathology in BRCA2 mutation carriers with prostate cancer and correlates with poor prognosis. Eur. Urol. 2015, 67, 496–503. [Google Scholar] [CrossRef] [PubMed]
  108. Porter, L.H.; Hashimoto, K.; Lawrence, M.G.; Pezaro, C.; Clouston, D.; Wang, H.; Papargiris, M.; Thorne, H.; Li, J.; Ryan, A.; et al. Intraductal carcinoma of the prostate can evade androgen deprivation, with emergence of castrate-tolerant cells. BJU Int. 2018, 121, 971–978. [Google Scholar] [CrossRef] [PubMed]
  109. Castro, E.; Romero-Laorden, N.; Del Pozo, A.; Lozano, R.; Medina, A.; Puente, J.; Piulats, J.M.; Lorente, D.; Saez, M.I.; Morales-Barrera, R.; et al. PROREPAIR-B: A Prospective Cohort Study of the Impact of Germline DNA Repair Mutations on the Outcomes of Patients with Metastatic Castration-Resistant Prostate Cancer. J. Clin. Oncol. 2019, 37, 490–503. [Google Scholar] [CrossRef] [PubMed]
  110. Kimura, H.; Mizuno, K.; Shiota, M.; Narita, S.; Terada, N.; Fujimoto, N.; Ogura, K.; Hatano, S.; Iwasaki, Y.; Hakozaki, N.; et al. Prognostic significance of pathogenic variants in BRCA1, BRCA2, ATM and PALB2 genes in men undergoing hormonal therapy for advanced prostate cancer. Br. J. Cancer 2022, 127, 1680–1690. [Google Scholar] [CrossRef]
  111. Marshall, C.H.; Fu, W.; Wang, H.; Baras, A.S.; Lotan, T.L.; Antonarakis, E.S. Prevalence of DNA repair gene mutations in localized prostate cancer according to clinical and pathologic features: Association of Gleason score and tumor stage. Prostate Cancer Prostatic Dis. 2019, 22, 59–65. [Google Scholar] [CrossRef]
  112. Momozawa, Y.; Iwasaki, Y.; Hirata, M.; Liu, X.; Kamatani, Y.; Takahashi, A.; Sugano, K.; Yoshida, T.; Murakami, Y.; Matsuda, K.; et al. Germline Pathogenic Variants in 7636 Japanese Patients with Prostate Cancer and 12,366 Controls. J. Natl. Cancer Inst. 2020, 112, 369–376. [Google Scholar] [CrossRef]
  113. Leongamornlert, D.; Saunders, E.; Dadaev, T.; Tymrakiewicz, M.; Goh, C.; Jugurnauth-Little, S.; Kozarewa, I.; Fenwick, K.; Assiotis, I.; Barrowdale, D.; et al. Frequent germline deleterious mutations in DNA repair genes in familial prostate cancer cases are associated with advanced disease. Br. J. Cancer 2014, 110, 1663–1672. [Google Scholar] [CrossRef]
  114. Turajlic, S.; Litchfield, K.; Xu, H.; Rosenthal, R.; McGranahan, N.; Reading, J.L.; Wong, Y.N.S.; Rowan, A.; Kanu, N.; Al Bakir, M.; et al. Insertion-and-deletion-derived tumour-specific neoantigens and the immunogenic phenotype: A pan-cancer analysis. Lancet Oncol. 2017, 18, 1009–1021. [Google Scholar] [CrossRef]
  115. Willis, J.A.; Reyes-Uribe, L.; Chang, K.; Lipkin, S.M.; Vilar, E. Immune Activation in Mismatch Repair-Deficient Carcinogenesis: More Than Just Mutational Rate. Clin. Cancer Res. 2020, 26, 11–17. [Google Scholar] [CrossRef]
  116. Schumacher, T.N.; Scheper, W.; Kvistborg, P. Cancer Neoantigens. Annu. Rev. Immunol. 2019, 37, 173–200. [Google Scholar] [CrossRef]
  117. Maby, P.; Tougeron, D.; Hamieh, M.; Mlecnik, B.; Kora, H.; Bindea, G.; Angell, H.K.; Fredriksen, T.; Elie, N.; Fauquembergue, E.; et al. Correlation between Density of CD8+ T-cell Infiltrate in Microsatellite Unstable Colorectal Cancers and Frameshift Mutations: A Rationale for Personalized Immunotherapy. Cancer Res. 2015, 75, 3446–3455. [Google Scholar] [CrossRef] [PubMed]
  118. Abida, W.; Cheng, M.L.; Armenia, J.; Middha, S.; Autio, K.A.; Vargas, H.A.; Rathkopf, D.; Morris, M.J.; Danila, D.C.; Slovin, S.F.; et al. Analysis of the Prevalence of Microsatellite Instability in Prostate Cancer and Response to Immune Checkpoint Blockade. JAMA Oncol. 2019, 5, 471–478. [Google Scholar] [CrossRef] [PubMed]
  119. Alexandrov, L.B.; Nik-Zainal, S.; Wedge, D.C.; Aparicio, S.A.; Behjati, S.; Biankin, A.V.; Bignell, G.R.; Bolli, N.; Borg, A.; Børresen-Dale, A.L.; et al. Signatures of mutational processes in human cancer. Nature 2013, 500, 415–421. [Google Scholar] [CrossRef] [PubMed]
  120. Alexandrov, L.B.; Nik-Zainal, S.; Wedge, D.C.; Campbell, P.J.; Stratton, M.R. Deciphering signatures of mutational processes operative in human cancer. Cell Rep. 2013, 3, 246–259. [Google Scholar] [CrossRef] [PubMed]
  121. Antonarakis, E.S.; Isaacsson Velho, P.; Fu, W.; Wang, H.; Agarwal, N.; Sacristan Santos, V.; Maughan, B.L.; Pili, R.; Adra, N.; Sternberg, C.N.; et al. CDK12-Altered Prostate Cancer: Clinical Features and Therapeutic Outcomes to Standard Systemic Therapies, Poly (ADP-Ribose) Polymerase Inhibitors, and PD-1 Inhibitors. JCO Precis. Oncol. 2020, 4, 370–381. [Google Scholar] [CrossRef]
  122. Pritchard, C.C.; Morrissey, C.; Kumar, A.; Zhang, X.; Smith, C.; Coleman, I.; Salipante, S.J.; Milbank, J.; Yu, M.; Grady, W.M.; et al. Complex MSH2 and MSH6 mutations in hypermutated microsatellite unstable advanced prostate cancer. Nat. Commun. 2014, 5, 4988. [Google Scholar] [CrossRef]
  123. Pritchard, C.C.; Mateo, J.; Walsh, M.F.; De Sarkar, N.; Abida, W.; Beltran, H.; Garofalo, A.; Gulati, R.; Carreira, S.; Eeles, R.; et al. Inherited DNA-Repair Gene Mutations in Men with Metastatic Prostate Cancer. N. Engl. J. Med. 2016, 375, 443–453. [Google Scholar] [CrossRef]
  124. Jiricny, J. The multifaceted mismatch-repair system. Nat. Rev. Mol. Cell Biol. 2006, 7, 335–346. [Google Scholar] [CrossRef] [PubMed]
  125. Jarzen, J.; Diamanduros, A.; Scarpinato, K.D. Mismatch repair proteins in recurrent prostate cancer. Adv. Clin. Chem. 2013, 60, 65–84. [Google Scholar]
  126. Burger, M.; Burger, S.J.; Denzinger, S.; Wild, P.J.; Wieland, W.F.; Blaszyk, H.; Obermann, E.C.; Stoehr, R.; Hartmann, A. Elevated microsatellite instability at selected tetranucleotide repeats does not correlate with clinicopathologic features of bladder cancer. Eur. Urol. 2006, 50, 770–775; discussion 776. [Google Scholar] [CrossRef]
  127. Chen, Y.; Wang, J.; Fraig, M.M.; Metcalf, J.; Turner, W.R.; Bissada, N.K.; Watson, D.K.; Schweinfest, C.W. Defects of DNA mismatch repair in human prostate cancer. Cancer Res. 2001, 61, 4112–4121. [Google Scholar]
  128. Velasco, A.; Hewitt, S.M.; Albert, P.S.; Hossein, M.; Rosenberg, H.; Martinez, C.; Sagalowsky, A.I.; McConnell, J.D.; Marston, W.; Leach, F.S. Differential expression of the mismatch repair gene hMSH2 in malignant prostate tissue is associated with cancer recurrence. Cancer 2002, 94, 690–699. [Google Scholar] [CrossRef] [PubMed]
  129. Sun, X.; Chen, C.; Vessella, R.L.; Dong, J.T. Microsatellite instability and mismatch repair target gene mutations in cell lines and xenografts of prostate cancer. Prostate 2006, 66, 660–666. [Google Scholar] [CrossRef] [PubMed]
  130. Chen, Y.; Wang, J.; Fraig, M.M.; Henderson, K.; Bissada, N.K.; Watson, D.K.; Schweinfest, C.W. Alterations in PMS2, MSH2 and MLH1 expression in human prostate cancer. Int. J. Oncol. 2003, 22, 1033–1043. [Google Scholar] [CrossRef] [PubMed]
  131. Dahiya, R.; Lee, C.; McCarville, J.; Hu, W.; Kaur, G.; Deng, G. High frequency of genetic instability of microsatellites in human prostatic adenocarcinoma. Int. J. Cancer 1997, 72, 762–767. [Google Scholar] [CrossRef]
  132. Watanabe, M.; Imai, H.; Shiraishi, T.; Shimazaki, J.; Kotake, T.; Yatani, R. Microsatellite instability in human prostate cancer. Br. J. Cancer 1995, 72, 562–564. [Google Scholar] [CrossRef] [PubMed]
  133. Nava Rodrigues, D.; Rescigno, P.; Liu, D.; Yuan, W.; Carreira, S.; Lambros, M.B.; Seed, G.; Mateo, J.; Riisnaes, R.; Mullane, S.; et al. Immunogenomic analyses associate immunological alterations with mismatch repair defects in prostate cancer. J. Clin. Investig. 2018, 128, 4441–4453. [Google Scholar] [CrossRef] [PubMed]
  134. Ebelt, K.; Babaryka, G.; Figel, A.M.; Pohla, H.; Buchner, A.; Stief, C.G.; Eisenmenger, W.; Kirchner, T.; Schendel, D.J.; Noessner, E. Dominance of CD4+ lymphocytic infiltrates with disturbed effector cell characteristics in the tumor microenvironment of prostate carcinoma. Prostate 2008, 68, 1–10. [Google Scholar] [CrossRef] [PubMed]
  135. Gannon, P.O.; Poisson, A.O.; Delvoye, N.; Lapointe, R.; Mes-Masson, A.M.; Saad, F. Characterization of the intra-prostatic immune cell infiltration in androgen-deprived prostate cancer patients. J. Immunol. Methods 2009, 348, 9–17. [Google Scholar] [CrossRef]
  136. Hussein, M.R.; Al-Assiri, M.; Musalam, A.O. Phenotypic characterization of the infiltrating immune cells in normal prostate, benign nodular prostatic hyperplasia and prostatic adenocarcinoma. Exp. Mol. Pathol. 2009, 86, 108–113. [Google Scholar] [CrossRef]
  137. Shirotake, S.; Miyajima, A.; Kosaka, T.; Tanaka, N.; Kikuchi, E.; Mikami, S.; Okada, Y.; Oya, M. Regulation of monocyte chemoattractant protein-1 through angiotensin II type 1 receptor in prostate cancer. Am. J. Pathol. 2012, 180, 1008–1016. [Google Scholar] [CrossRef]
  138. Davidsson, S.; Ohlson, A.L.; Andersson, S.O.; Fall, K.; Meisner, A.; Fiorentino, M.; Andrén, O.; Rider, J.R. CD4 helper T cells, CD8 cytotoxic T cells, and FOXP3+ regulatory T cells with respect to lethal prostate cancer. Mod. Pathol. 2013, 26, 448–455. [Google Scholar] [CrossRef] [PubMed]
  139. Ebelt, K.; Babaryka, G.; Frankenberger, B.; Stief, C.G.; Eisenmenger, W.; Kirchner, T.; Schendel, D.J.; Noessner, E. Prostate cancer lesions are surrounded by FOXP3+, PD-1+ and B7-H1+ lymphocyte clusters. Eur. J. Cancer 2009, 45, 1664–1672. [Google Scholar] [CrossRef]
  140. Kiniwa, Y.; Miyahara, Y.; Wang, H.Y.; Peng, W.; Peng, G.; Wheeler, T.M.; Thompson, T.C.; Old, L.J.; Wang, R.F. CD8+ Foxp3+ regulatory T cells mediate immunosuppression in prostate cancer. Clin. Cancer Res. 2007, 13, 6947–6958. [Google Scholar] [CrossRef]
  141. Sutherland, J.S.; Goldberg, G.L.; Hammett, M.V.; Uldrich, A.P.; Berzins, S.P.; Heng, T.S.; Blazar, B.R.; Millar, J.L.; Malin, M.A.; Chidgey, A.P.; et al. Activation of thymic regeneration in mice and humans following androgen blockade. J. Immunol. 2005, 175, 2741–2753. [Google Scholar] [CrossRef]
  142. Wilson, C.A.; Mrose, S.A.; Thomas, D.W. Enhanced production of B lymphocytes after castration. Blood 1995, 85, 1535–1539. [Google Scholar] [CrossRef] [PubMed]
  143. Drake, C.G.; Doody, A.D.; Mihalyo, M.A.; Huang, C.T.; Kelleher, E.; Ravi, S.; Hipkiss, E.L.; Flies, D.B.; Kennedy, E.P.; Long, M.; et al. Androgen ablation mitigates tolerance to a prostate/prostate cancer-restricted antigen. Cancer Cell 2005, 7, 239–249. [Google Scholar] [CrossRef]
  144. Mercader, M.; Bodner, B.K.; Moser, M.T.; Kwon, P.S.; Park, E.S.; Manecke, R.G.; Ellis, T.M.; Wojcik, E.M.; Yang, D.; Flanigan, R.C.; et al. T cell infiltration of the prostate induced by androgen withdrawal in patients with prostate cancer. Proc. Natl. Acad. Sci. USA 2001, 98, 14565–14570. [Google Scholar] [CrossRef] [PubMed]
  145. Berger, S.L.; Kouzarides, T.; Shiekhattar, R.; Shilatifard, A. An operational definition of epigenetics. Genes Dev. 2009, 23, 781–783. [Google Scholar] [CrossRef] [PubMed]
  146. Dawson, M.A.; Kouzarides, T. Cancer epigenetics: From mechanism to therapy. Cell 2012, 150, 12–27. [Google Scholar] [CrossRef]
  147. Kukkonen, K.; Taavitsainen, S.; Huhtala, L.; Uusi-Makela, J.; Granberg, K.J.; Nykter, M.; Urbanucci, A. Chromatin and Epigenetic Dysregulation of Prostate Cancer Development, Progression, and Therapeutic Response. Cancers 2021, 13, 3325. [Google Scholar] [CrossRef] [PubMed]
  148. Kgatle, M.M.; Kalla, A.A.; Islam, M.M.; Sathekge, M.; Moorad, R. Prostate Cancer: Epigenetic Alterations, Risk Factors, and Therapy. Prostate Cancer 2016, 2016, 5653862. [Google Scholar] [CrossRef]
  149. Ruggero, K.; Farran-Matas, S.; Martinez-Tebar, A.; Aytes, A. Epigenetic Regulation in Prostate Cancer Progression. Curr. Mol. Biol. Rep. 2018, 4, 101–115. [Google Scholar] [CrossRef]
  150. Sugiura, M.; Sato, H.; Kanesaka, M.; Imamura, Y.; Sakamoto, S.; Ichikawa, T.; Kaneda, A. Epigenetic modifications in prostate cancer. Int. J. Urol. 2021, 28, 140–149. [Google Scholar] [CrossRef]
  151. Ruscetti, M.; Dadashian, E.L.; Guo, W.; Quach, B.; Mulholland, D.J.; Park, J.W.; Tran, L.M.; Kobayashi, N.; Bianchi-Frias, D.; Xing, Y.; et al. HDAC inhibition impedes epithelial-mesenchymal plasticity and suppresses metastatic, castration-resistant prostate cancer. Oncogene 2016, 35, 3781–3795. [Google Scholar] [CrossRef]
  152. 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]
  153. Bryant, R.J.; Cross, N.A.; Eaton, C.L.; Hamdy, F.C.; Cunliffe, V.T. EZH2 promotes proliferation and invasiveness of prostate cancer cells. Prostate 2007, 67, 547–556. [Google Scholar] [CrossRef]
  154. Varambally, S.; Dhanasekaran, S.M.; Zhou, M.; Barrette, T.R.; Kumar-Sinha, C.; Sanda, M.G.; Ghosh, D.; Pienta, K.J.; Sewalt, R.G.; Otte, A.P.; et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature 2002, 419, 624–629. [Google Scholar] [CrossRef]
  155. Melling, N.; Thomsen, E.; Tsourlakis, M.C.; Kluth, M.; Hube-Magg, C.; Minner, S.; Koop, C.; Graefen, M.; Heinzer, H.; Wittmer, C.; et al. Overexpression of enhancer of zeste homolog 2 (EZH2) characterizes an aggressive subset of prostate cancers and predicts patient prognosis independently from pre- and postoperatively assessed clinicopathological parameters. Carcinogenesis 2015, 36, 1333–1340. [Google Scholar] [CrossRef]
  156. Urbanucci, A.; Mills, I.G. Bromodomain-containing proteins in prostate cancer. Mol. Cell. Endocrinol. 2018, 462, 31–40. [Google Scholar] [CrossRef]
  157. Fujisawa, T.; Filippakopoulos, P. Functions of bromodomain-containing proteins and their roles in homeostasis and cancer. Nat. Rev. Mol. Cell Biol. 2017, 18, 246–262. [Google Scholar] [CrossRef] [PubMed]
  158. Asangani, I.A.; Dommeti, V.L.; Wang, X.; Malik, R.; Cieslik, M.; Yang, R.; Escara-Wilke, J.; Wilder-Romans, K.; Dhanireddy, S.; Engelke, C.; et al. Therapeutic targeting of BET bromodomain proteins in castration-resistant prostate cancer. Nature 2014, 510, 278–282. [Google Scholar] [CrossRef] [PubMed]
  159. Asangani, I.A.; Wilder-Romans, K.; Dommeti, V.L.; Krishnamurthy, P.M.; Apel, I.J.; Escara-Wilke, J.; Plymate, S.R.; Navone, N.M.; Wang, S.; Feng, F.Y.; et al. BET Bromodomain Inhibitors Enhance Efficacy and Disrupt Resistance to AR Antagonists in the Treatment of Prostate Cancer. Mol. Cancer Res. MCR 2016, 14, 324–331. [Google Scholar] [CrossRef]
  160. Welti, J.; Sharp, A.; Yuan, W.; Dolling, D.; Nava Rodrigues, D.; Figueiredo, I.; Gil, V.; Neeb, A.; Clarke, M.; Seed, G.; et al. Targeting Bromodomain and Extra-Terminal (BET) Family Proteins in Castration-Resistant Prostate Cancer (CRPC). Clin. Cancer Res. 2018, 24, 3149–3162. [Google Scholar] [CrossRef] [PubMed]
  161. Conteduca, V.; Hess, J.; Yamada, Y.; Ku, S.Y.; Beltran, H. Epigenetics in prostate cancer: Clinical implications. Transl. Androl. Urol. 2021, 10, 3104–3116. [Google Scholar] [CrossRef]
  162. Zhao, S.G.; Chen, W.S.; Li, H.; Foye, A.; Zhang, M.; Sjöström, M.; Aggarwal, R.; Playdle, D.; Liao, A.; Alumkal, J.J.; et al. The DNA methylation landscape of advanced prostate cancer. Nat. Genet. 2020, 52, 778–789. [Google Scholar] [CrossRef]
  163. Kulis, M.; Esteller, M. DNA methylation and cancer. Adv. Genet. 2010, 70, 27–56. [Google Scholar] [PubMed]
  164. Rasmussen, K.D.; Helin, K. Role of TET enzymes in DNA methylation, development, and cancer. Genes Dev. 2016, 30, 733–750. [Google Scholar] [CrossRef] [PubMed]
  165. Antequera, F. Structure, function and evolution of CpG island promoters. Cell. Mol. Life Sci. CMLS 2003, 60, 1647–1658. [Google Scholar] [CrossRef] [PubMed]
  166. Gordevičius, J.; Kriščiūnas, A.; Groot, D.E.; Yip, S.M.; Susic, M.; Kwan, A.; Kustra, R.; Joshua, A.M.; Chi, K.N.; Petronis, A.; et al. Cell-Free DNA Modification Dynamics in Abiraterone Acetate-Treated Prostate Cancer Patients. Clin. Cancer Res. 2018, 24, 3317–3324. [Google Scholar] [CrossRef] [PubMed]
  167. Jin, S.G.; Jiang, Y.; Qiu, R.; Rauch, T.A.; Wang, Y.; Schackert, G.; Krex, D.; Lu, Q.; Pfeifer, G.P. 5-Hydroxymethylcytosine is strongly depleted in human cancers but its levels do not correlate with IDH1 mutations. Cancer Res. 2011, 71, 7360–7365. [Google Scholar] [CrossRef] [PubMed]
  168. Nickerson, M.L.; Das, S.; Im, K.M.; Turan, S.; Berndt, S.I.; Li, H.; Lou, H.; Brodie, S.A.; Billaud, J.N.; Zhang, T.; et al. TET2 binds the androgen receptor and loss is associated with prostate cancer. Oncogene 2017, 36, 2172–2183. [Google Scholar] [CrossRef] [PubMed]
  169. Takayama, K.; Misawa, A.; Suzuki, T.; Takagi, K.; Hayashizaki, Y.; Fujimura, T.; Homma, Y.; Takahashi, S.; Urano, T.; Inoue, S. TET2 repression by androgen hormone regulates global hydroxymethylation status and prostate cancer progression. Nat. Commun. 2015, 6, 8219. [Google Scholar] [CrossRef]
  170. Hsu, C.H.; Peng, K.L.; Kang, M.L.; Chen, Y.R.; Yang, Y.C.; Tsai, C.H.; Chu, C.S.; Jeng, Y.M.; Chen, Y.T.; Lin, F.M.; et al. TET1 suppresses cancer invasion by activating the tissue inhibitors of metalloproteinases. Cell Rep. 2012, 2, 568–579. [Google Scholar] [CrossRef]
  171. Stein, R.; Gruenbaum, Y.; Pollack, Y.; Razin, A.; Cedar, H. Clonal inheritance of the pattern of DNA methylation in mouse cells. Proc. Natl. Acad. Sci. USA 1982, 79, 61–65. [Google Scholar] [CrossRef] [PubMed]
  172. Bastian, P.J.; Yegnasubramanian, S.; Palapattu, G.S.; Rogers, C.G.; Lin, X.; De Marzo, A.M.; Nelson, W.G. Molecular biomarker in prostate cancer: The role of CpG island hypermethylation. Eur. Urol. 2004, 46, 698–708. [Google Scholar] [CrossRef] [PubMed]
  173. Delgado-Cruzata, L.; Hruby, G.W.; Gonzalez, K.; McKiernan, J.; Benson, M.C.; Santella, R.M.; Shen, J. DNA methylation changes correlate with Gleason score and tumor stage in prostate cancer. DNA Cell Biol. 2012, 31, 187–192. [Google Scholar] [CrossRef] [PubMed]
  174. Rouprêt, M.; Hupertan, V.; Catto, J.W.; Yates, D.R.; Rehman, I.; Proctor, L.M.; Phillips, J.; Meuth, M.; Cussenot, O.; Hamdy, F.C. Promoter hypermethylation in circulating blood cells identifies prostate cancer progression. Int. J. Cancer 2008, 122, 952–956. [Google Scholar] [CrossRef] [PubMed]
  175. Mahon, K.L.; Qu, W.; Devaney, J.; Paul, C.; Castillo, L.; Wykes, R.J.; Chatfield, M.D.; Boyer, M.J.; Stockler, M.R.; Marx, G.; et al. Methylated Glutathione S-transferase 1 (mGSTP1) is a potential plasma free DNA epigenetic marker of prognosis and response to chemotherapy in castrate-resistant prostate cancer. Br. J. Cancer 2014, 111, 1802–1809. [Google Scholar] [CrossRef] [PubMed]
  176. Graça, I.; Pereira-Silva, E.; Henrique, R.; Packham, G.; Crabb, S.J.; Jerónimo, C. Epigenetic modulators as therapeutic targets in prostate cancer. Clin. Epigenet. 2016, 8, 98. [Google Scholar] [CrossRef]
  177. Yegnasubramanian, S.; Haffner, M.C.; Zhang, Y.; Gurel, B.; Cornish, T.C.; Wu, Z.; Irizarry, R.A.; Morgan, J.; Hicks, J.; DeWeese, T.L.; et al. DNA hypomethylation arises later in prostate cancer progression than CpG island hypermethylation and contributes to metastatic tumor heterogeneity. Cancer Res. 2008, 68, 8954–8967. [Google Scholar] [CrossRef]
  178. Singal, R.; Ramachandran, K.; Gordian, E.; Quintero, C.; Zhao, W.; Reis, I.M. Phase I/II study of azacitidine, docetaxel, and prednisone in patients with metastatic castration-resistant prostate cancer previously treated with docetaxel-based therapy. Clin. Genitourin. Cancer 2015, 13, 22–31. [Google Scholar] [CrossRef]
  179. Ellinger, J.; Kahl, P.; von der Gathen, J.; Rogenhofer, S.; Heukamp, L.C.; Gütgemann, I.; Walter, B.; Hofstädter, F.; Büttner, R.; Müller, S.C.; et al. Global levels of histone modifications predict prostate cancer recurrence. Prostate 2010, 70, 61–69. [Google Scholar] [CrossRef]
  180. Chen, Z.; Wang, L.; Wang, Q.; Li, W. Histone modifications and chromatin organization in prostate cancer. Epigenomics 2010, 2, 551–560. [Google Scholar] [CrossRef]
  181. Dardenne, E.; Beltran, H.; Benelli, M.; Gayvert, K.; Berger, A.; Puca, L.; Cyrta, J.; Sboner, A.; Noorzad, Z.; MacDonald, T.; et al. N-Myc Induces an EZH2-Mediated Transcriptional Program Driving Neuroendocrine Prostate Cancer. Cancer Cell 2016, 30, 563–577. [Google Scholar] [CrossRef]
  182. Italiano, A.; Soria, J.C.; Toulmonde, M.; Michot, J.M.; Lucchesi, C.; Varga, A.; Coindre, J.M.; Blakemore, S.J.; Clawson, A.; Suttle, B.; et al. Tazemetostat, an EZH2 inhibitor, in relapsed or refractory B-cell non-Hodgkin lymphoma and advanced solid tumours: A first-in-human, open-label, phase 1 study. Lancet Oncol. 2018, 19, 649–659. [Google Scholar] [CrossRef]
  183. Liu, Q.; Wang, G.; Li, Q.; Jiang, W.; Kim, J.S.; Wang, R.; Zhu, S.; Wang, X.; Yan, L.; Yi, Y.; et al. Polycomb group proteins EZH2 and EED directly regulate androgen receptor in advanced prostate cancer. Int. J. Cancer 2019, 145, 415–426. [Google Scholar] [CrossRef]
  184. Aggarwal, R.R.; Schweizer, M.T.; Nanus, D.M.; Pantuck, A.J.; Heath, E.I.; Campeau, E.; Attwell, S.; Norek, K.; Snyder, M.; Bauman, L.; et al. A Phase Ib/IIa Study of the Pan-BET Inhibitor ZEN-3694 in Combination with Enzalutamide in Patients with Metastatic Castration-resistant Prostate Cancer. Clin. Cancer Res. 2020, 26, 5338–5347. [Google Scholar] [CrossRef]
  185. Davies, A.; Zoubeidi, A.; Selth, L.A. The epigenetic and transcriptional landscape of neuroendocrine prostate cancer. Endocr.-Relat. Cancer 2020, 27, R35–R50. [Google Scholar] [CrossRef]
  186. Crea, F.; Sun, L.; Mai, A.; Chiang, Y.T.; Farrar, W.L.; Danesi, R.; Helgason, C.D. The emerging role of histone lysine demethylases in prostate cancer. Mol. Cancer 2012, 11, 52. [Google Scholar] [CrossRef]
  187. He, Y.; Zhao, Y.; Wang, L.; Bohrer, L.R.; Pan, Y.; Wang, L.; Huang, H. LSD1 promotes S-phase entry and tumorigenesis via chromatin co-occupation with E2F1 and selective H3K9 demethylation. Oncogene 2018, 37, 534–543. [Google Scholar] [CrossRef] [PubMed]
  188. Metzger, E.; Wissmann, M.; Yin, N.; Müller, J.M.; Schneider, R.; Peters, A.H.; Günther, T.; Buettner, R.; Schüle, R. LSD1 demethylates repressive histone marks to promote androgen-receptor-dependent transcription. Nature 2005, 437, 436–439. [Google Scholar] [CrossRef] [PubMed]
  189. Liang, Y.; Ahmed, M.; Guo, H.; Soares, F.; Hua, J.T.; Gao, S.; Lu, C.; Poon, C.; Han, W.; Langstein, J.; et al. LSD1-Mediated Epigenetic Reprogramming Drives CENPE Expression and Prostate Cancer Progression. Cancer Res. 2017, 77, 5479–5490. [Google Scholar] [CrossRef] [PubMed]
  190. Sehrawat, A.; Gao, L.; Wang, Y.; Bankhead, A., 3rd; McWeeney, S.K.; King, C.J.; Schwartzman, J.; Urrutia, J.; Bisson, W.H.; Coleman, D.J.; et al. LSD1 activates a lethal prostate cancer gene network independently of its demethylase function. Proc. Natl. Acad. Sci. USA 2018, 115, E4179–E4188. [Google Scholar] [CrossRef]
  191. Gao, S.; Chen, S.; Han, D.; Wang, Z.; Li, M.; Han, W.; Besschetnova, A.; Liu, M.; Zhou, F.; Barrett, D.; et al. Chromatin binding of FOXA1 is promoted by LSD1-mediated demethylation in prostate cancer. Nat. Genet. 2020, 52, 1011–1017. [Google Scholar] [CrossRef]
  192. Kanouni, T.; Severin, C.; Cho, R.W.; Yuen, N.Y.; Xu, J.; Shi, L.; Lai, C.; Del Rosario, J.R.; Stansfield, R.K.; Lawton, L.N.; et al. Discovery of CC-90011: A Potent and Selective Reversible Inhibitor of Lysine Specific Demethylase 1 (LSD1). J. Med. Chem. 2020, 63, 14522–14529. [Google Scholar] [CrossRef]
  193. Struhl, K. Histone acetylation and transcriptional regulatory mechanisms. Genes Dev. 1998, 12, 599–606. [Google Scholar] [CrossRef]
  194. Lovén, J.; Hoke, H.A.; Lin, C.Y.; Lau, A.; Orlando, D.A.; Vakoc, C.R.; Bradner, J.E.; Lee, T.I.; Young, R.A. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 2013, 153, 320–334. [Google Scholar] [CrossRef] [PubMed]
  195. Wen, S.; He, Y.; Wang, L.; Zhang, J.; Quan, C.; Niu, Y.; Huang, H. Aberrant activation of super enhancer and choline metabolism drive antiandrogen therapy resistance in prostate cancer. Oncogene 2020, 39, 6556–6571. [Google Scholar] [CrossRef] [PubMed]
  196. Hnisz, D.; Abraham, B.J.; Lee, T.I.; Lau, A.; Saint-André, V.; Sigova, A.A.; Hoke, H.A.; Young, R.A. Super-enhancers in the control of cell identity and disease. Cell 2013, 155, 934–947. [Google Scholar] [CrossRef] [PubMed]
  197. Welti, J.; Sharp, A.; Brooks, N.; Yuan, W.; McNair, C.; Chand, S.N.; Pal, A.; Figueiredo, I.; Riisnaes, R.; Gurel, B.; et al. Targeting the p300/CBP Axis in Lethal Prostate Cancer. Cancer Discov. 2021, 11, 1118–1137. [Google Scholar] [CrossRef] [PubMed]
  198. Lasko, L.M.; Jakob, C.G.; Edalji, R.P.; Qiu, W.; Montgomery, D.; Digiammarino, E.L.; Hansen, T.M.; Risi, R.M.; Frey, R.; Manaves, V.; et al. Discovery of a selective catalytic p300/CBP inhibitor that targets lineage-specific tumours. Nature 2017, 550, 128–132. [Google Scholar] [CrossRef] [PubMed]
  199. He, Z.X.; Wei, B.F.; Zhang, X.; Gong, Y.P.; Ma, L.Y.; Zhao, W. Current development of CBP/p300 inhibitors in the last decade. Eur. J. Med. Chem. 2021, 209, 112861. [Google Scholar] [CrossRef] [PubMed]
  200. Seto, E.; Yoshida, M. Erasers of histone acetylation: The histone deacetylase enzymes. Cold Spring Harb. Perspect. Biol. 2014, 6, a018713. [Google Scholar] [CrossRef]
  201. Roche, J.; Bertrand, P. Inside HDACs with more selective HDAC inhibitors. Eur. J. Med. Chem. 2016, 121, 451–483. [Google Scholar] [CrossRef]
  202. Ropero, S.; Esteller, M. The role of histone deacetylases (HDACs) in human cancer. Mol. Oncol. 2007, 1, 19–25. [Google Scholar] [CrossRef]
  203. Weichert, W.; Röske, A.; Gekeler, V.; Beckers, T.; Stephan, C.; Jung, K.; Fritzsche, F.R.; Niesporek, S.; Denkert, C.; Dietel, M.; et al. Histone deacetylases 1, 2 and 3 are highly expressed in prostate cancer and HDAC2 expression is associated with shorter PSA relapse time after radical prostatectomy. Br. J. Cancer 2008, 98, 604–610. [Google Scholar] [CrossRef]
  204. Kaushik, D.; Vashistha, V.; Isharwal, S.; Sediqe, S.A.; Lin, M.F. Histone deacetylase inhibitors in castration-resistant prostate cancer: Molecular mechanism of action and recent clinical trials. Ther. Adv. Urol. 2015, 7, 388–395. [Google Scholar] [CrossRef]
  205. Ferrari, A.C.; Alumkal, J.J.; Stein, M.N.; Taplin, M.E.; Babb, J.; Barnett, E.S.; Gomez-Pinillos, A.; Liu, X.; Moore, D.; DiPaola, R.; et al. Epigenetic Therapy with Panobinostat Combined with Bicalutamide Rechallenge in Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2019, 25, 52–63. [Google Scholar] [CrossRef] [PubMed]
  206. Rana, Z.; Diermeier, S.; Hanif, M.; Rosengren, R.J. Understanding Failure and Improving Treatment Using HDAC Inhibitors for Prostate Cancer. Biomedicines 2020, 8, 22. [Google Scholar] [CrossRef]
  207. Takayama, K.; Horie-Inoue, K.; Katayama, S.; Suzuki, T.; Tsutsumi, S.; Ikeda, K.; Urano, T.; Fujimura, T.; Takagi, K.; Takahashi, S.; et al. Androgen-responsive long noncoding RNA CTBP1-AS promotes prostate cancer. EMBO J. 2013, 32, 1665–1680. [Google Scholar] [CrossRef] [PubMed]
  208. Zheng, J.; Zhao, S.; He, X.; Zheng, Z.; Bai, W.; Duan, Y.; Cheng, S.; Wang, J.; Liu, X.; Zhang, G. The up-regulation of long non-coding RNA CCAT2 indicates a poor prognosis for prostate cancer and promotes metastasis by affecting epithelial-mesenchymal transition. Biochem. Biophys. Res. Commun. 2016, 480, 508–514. [Google Scholar] [CrossRef]
  209. Yang, L.; Lin, C.; Jin, C.; Yang, J.C.; Tanasa, B.; Li, W.; Merkurjev, D.; Ohgi, K.A.; Meng, D.; Zhang, J.; et al. lncRNA-dependent mechanisms of androgen-receptor-regulated gene activation programs. Nature 2013, 500, 598–602. [Google Scholar] [CrossRef] [PubMed]
  210. He, J.H.; Zhang, J.Z.; Han, Z.P.; Wang, L.; Lv, Y.B.; Li, Y.G. Reciprocal regulation of PCGEM1 and miR-145 promote proliferation of LNCaP prostate cancer cells. J. Exp. Clin. Cancer Res. CR 2014, 33, 72. [Google Scholar] [CrossRef]
  211. Fang, Z.; Xu, C.; Li, Y.; Cai, X.; Ren, S.; Liu, H.; Wang, Y.; Wang, F.; Chen, R.; Qu, M.; et al. A feed-forward regulatory loop between androgen receptor and PlncRNA-1 promotes prostate cancer progression. Cancer Lett. 2016, 374, 62–74. [Google Scholar] [CrossRef]
  212. Prensner, J.R.; Sahu, A.; Iyer, M.K.; Malik, R.; Chandler, B.; Asangani, I.A.; Poliakov, A.; Vergara, I.A.; Alshalalfa, M.; Jenkins, R.B.; et al. The IncRNAs PCGEM1 and PRNCR1 are not implicated in castration resistant prostate cancer. Oncotarget 2014, 5, 1434–1438. [Google Scholar] [CrossRef]
  213. Hung, C.L.; Wang, L.Y.; Yu, Y.L.; Chen, H.W.; Srivastava, S.; Petrovics, G.; Kung, H.J. A long noncoding RNA connects c-Myc to tumor metabolism. Proc. Natl. Acad. Sci. USA 2014, 111, 18697–18702. [Google Scholar] [CrossRef] [PubMed]
  214. Wu, J.; Cheng, G.; Zhang, C.; Zheng, Y.; Xu, H.; Yang, H.; Hua, L. Long noncoding RNA LINC01296 is associated with poor prognosis in prostate cancer and promotes cancer-cell proliferation and metastasis. OncoTargets Ther. 2017, 10, 1843–1852. [Google Scholar] [CrossRef] [PubMed]
  215. Wang, J.; Cheng, G.; Li, X.; Pan, Y.; Qin, C.; Yang, H.; Hua, L.; Wang, Z. Overexpression of long non-coding RNA LOC400891 promotes tumor progression and poor prognosis in prostate cancer. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2016, 37, 9603–9613. [Google Scholar] [CrossRef] [PubMed]
  216. Zhang, C.; Liu, C.; Wu, J.; Zheng, Y.; Xu, H.; Cheng, G.; Hua, L. Upregulation of long noncoding RNA LOC440040 promotes tumor progression and predicts poor prognosis in patients with prostate cancer. OncoTargets Ther. 2017, 10, 4945–4954. [Google Scholar] [CrossRef] [PubMed]
  217. Zhang, A.; Zhao, J.C.; Kim, J.; Fong, K.W.; Yang, Y.A.; Chakravarti, D.; Mo, Y.Y.; Yu, J. LncRNA HOTAIR Enhances the Androgen-Receptor-Mediated Transcriptional Program and Drives Castration-Resistant Prostate Cancer. Cell Rep. 2015, 13, 209–221. [Google Scholar] [CrossRef] [PubMed]
  218. Parolia, A.; Crea, F.; Xue, H.; Wang, Y.; Mo, F.; Ramnarine, V.R.; Liu, H.H.; Lin, D.; Saidy, N.R.; Clermont, P.L.; et al. The long non-coding RNA PCGEM1 is regulated by androgen receptor activity in vivo. Mol. Cancer 2015, 14, 46. [Google Scholar] [CrossRef] [PubMed]
  219. Chiyomaru, T.; Yamamura, S.; Fukuhara, S.; Yoshino, H.; Kinoshita, T.; Majid, S.; Saini, S.; Chang, I.; Tanaka, Y.; Enokida, H.; et al. Genistein inhibits prostate cancer cell growth by targeting miR-34a and oncogenic HOTAIR. PLoS ONE 2013, 8, e70372. [Google Scholar] [CrossRef]
  220. Goto, Y.; Kurozumi, A.; Arai, T.; Nohata, N.; Kojima, S.; Okato, A.; Kato, M.; Yamazaki, K.; Ishida, Y.; Naya, Y.; et al. Impact of novel miR-145-3p regulatory networks on survival in patients with castration-resistant prostate cancer. Br. J. Cancer 2017, 117, 409–420. [Google Scholar] [CrossRef]
  221. Mitra, R.; Adams, C.M.; Jiang, W.; Greenawalt, E.; Eischen, C.M. Pan-cancer analysis reveals cooperativity of both strands of microRNA that regulate tumorigenesis and patient survival. Nat. Commun. 2020, 11, 968. [Google Scholar] [CrossRef]
  222. Fuse, M.; Kojima, S.; Enokida, H.; Chiyomaru, T.; Yoshino, H.; Nohata, N.; Kinoshita, T.; Sakamoto, S.; Naya, Y.; Nakagawa, M.; et al. Tumor suppressive microRNAs (miR-222 and miR-31) regulate molecular pathways based on microRNA expression signature in prostate cancer. J. Hum. Genet. 2012, 57, 691–699. [Google Scholar] [CrossRef]
  223. Hashimoto, Y.; Shiina, M.; Dasgupta, P.; Kulkarni, P.; Kato, T.; Wong, R.K.; Tanaka, Y.; Shahryari, V.; Maekawa, S.; Yamamura, S.; et al. Upregulation of miR-130b Contributes to Risk of Poor Prognosis and Racial Disparity in African-American Prostate Cancer. Cancer Prev. Res. 2019, 12, 585–598. [Google Scholar] [CrossRef] [PubMed]
  224. Mitchell, P.S.; Parkin, R.K.; Kroh, E.M.; Fritz, B.R.; Wyman, S.K.; Pogosova-Agadjanyan, E.L.; Peterson, A.; Noteboom, J.; O’Briant, K.C.; Allen, A.; et al. Circulating microRNAs as stable blood-based markers for cancer detection. Proc. Natl. Acad. Sci. USA 2008, 105, 10513–10518. [Google Scholar] [CrossRef] [PubMed]
  225. Ponti, G.; Maccaferri, M.; Percesepe, A.; Tomasi, A.; Ozben, T. Liquid biopsy with cell free DNA: New horizons for prostate cancer. Crit. Rev. Clin. Lab. Sci. 2021, 58, 60–76. [Google Scholar] [CrossRef] [PubMed]
  226. Endzeliņš, E.; Melne, V.; Kalniņa, Z.; Lietuvietis, V.; Riekstiņa, U.; Llorente, A.; Linē, A. Diagnostic, prognostic and predictive value of cell-free miRNAs in prostate cancer: A systematic review. Mol. Cancer 2016, 15, 41. [Google Scholar] [CrossRef] [PubMed]
  227. Urabe, F.; Matsuzaki, J.; Yamamoto, Y.; Kimura, T.; Hara, T.; Ichikawa, M.; Takizawa, S.; Aoki, Y.; Niida, S.; Sakamoto, H.; et al. Large-scale Circulating microRNA Profiling for the Liquid Biopsy of Prostate Cancer. Clin. Cancer Res. 2019, 25, 3016–3025. [Google Scholar] [CrossRef] [PubMed]
  228. Pimenta, R.C.; Viana, N.I.; Amaral, G.Q.; Park, R.; Morais, D.R.; Pontes, J., Jr.; Guimaraes, V.R.; Camargo, J.A.; Leite, K.R.; Nahas, W.C.; et al. MicroRNA-23b and microRNA-27b plus flutamide treatment enhances apoptosis rate and decreases CCNG1 expression in a castration-resistant prostate cancer cell line. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2018, 40, 1010428318803011. [Google Scholar] [CrossRef] [PubMed]
  229. Sun, T.; Wang, Q.; Balk, S.; Brown, M.; Lee, G.S.; Kantoff, P. The role of microRNA-221 and microRNA-222 in androgen-independent prostate cancer cell lines. Cancer Res. 2009, 69, 3356–3363. [Google Scholar] [CrossRef]
  230. Jiao, L.; Deng, Z.; Xu, C.; Yu, Y.; Li, Y.; Yang, C.; Chen, J.; Liu, Z.; Huang, G.; Li, L.C.; et al. miR-663 induces castration-resistant prostate cancer transformation and predicts clinical recurrence. J. Cell. Physiol. 2014, 229, 834–844. [Google Scholar] [CrossRef]
  231. Benoist, G.E.; van Oort, I.M.; Boerrigter, E.; Verhaegh, G.W.; van Hooij, O.; Groen, L.; Smit, F.; de Mol, P.; Hamberg, P.; Dezentjé, V.O.; et al. Prognostic Value of Novel Liquid Biomarkers in Patients with Metastatic Castration-Resistant Prostate Cancer Treated with Enzalutamide: A Prospective Observational Study. Clin. Chem. 2020, 66, 842–851. [Google Scholar] [CrossRef]
  232. Yang, Y.; Jia, D.; Kim, H.; Abd Elmageed, Z.Y.; Datta, A.; Davis, R.; Srivastav, S.; Moroz, K.; Crawford, B.E.; Moparty, K.; et al. Dysregulation of miR-212 Promotes Castration Resistance through hnRNPH1-Mediated Regulation of AR and AR-V7: Implications for Racial Disparity of Prostate Cancer. Clin. Cancer Res. 2016, 22, 1744–1756. [Google Scholar] [CrossRef] [PubMed]
  233. Liu, B.; Sun, Y.; Tang, M.; Liang, C.; Huang, C.P.; Niu, Y.; Wang, Z.; Chang, C. The miR-361-3p increases enzalutamide (Enz) sensitivity via targeting the ARv7 and MKNK2 to better suppress the Enz-resistant prostate cancer. Cell Death Dis. 2020, 11, 807. [Google Scholar] [CrossRef]
  234. Huang, X.; Yuan, T.; Liang, M.; Du, M.; Xia, S.; Dittmar, R.; Wang, D.; See, W.; Costello, B.A.; Quevedo, F.; et al. Exosomal miR-1290 and miR-375 as prognostic markers in castration-resistant prostate cancer. Eur. Urol. 2015, 67, 33–41. [Google Scholar] [CrossRef]
  235. Samli, H.; Samli, M.; Vatansever, B.; Ardicli, S.; Aztopal, N.; Dincel, D.; Sahin, A.; Balci, F. Paclitaxel resistance and the role of miRNAs in prostate cancer cell lines. World J. Urol. 2019, 37, 1117–1126. [Google Scholar] [CrossRef] [PubMed]
  236. Chen, L.; Cao, H.; Feng, Y. MiR-199a suppresses prostate cancer paclitaxel resistance by targeting YES1. World J. Urol. 2018, 36, 357–365. [Google Scholar] [CrossRef] [PubMed]
  237. Ma, X.; Zou, L.; Li, X.; Chen, Z.; Lin, Q.; Wu, X. MicroRNA-195 regulates docetaxel resistance by targeting clusterin in prostate cancer. Biomed. Pharmacother. 2018, 99, 445–450. [Google Scholar] [CrossRef]
  238. Wu, G.; Wang, J.; Chen, G.; Zhao, X. microRNA-204 modulates chemosensitivity and apoptosis of prostate cancer cells by targeting zinc-finger E-box-binding homeobox 1 (ZEB1). Am. J. Transl. Res. 2017, 9, 3599–3610. [Google Scholar]
  239. Xu, B.; Niu, X.; Zhang, X.; Tao, J.; Wu, D.; Wang, Z.; Li, P.; Zhang, W.; Wu, H.; Feng, N.; et al. miR-143 decreases prostate cancer cells proliferation and migration and enhances their sensitivity to docetaxel through suppression of KRAS. Mol. Cell. Biochem. 2011, 350, 207–213. [Google Scholar] [CrossRef]
  240. Yu, J.; Lu, Y.; Cui, D.; Li, E.; Zhu, Y.; Zhao, Y.; Zhao, F.; Xia, S. miR-200b suppresses cell proliferation, migration and enhances chemosensitivity in prostate cancer by regulating Bmi-1. Oncol. Rep. 2014, 31, 910–918. [Google Scholar] [CrossRef]
  241. Lin, H.M.; Mahon, K.L.; Spielman, C.; Gurney, H.; Mallesara, G.; Stockler, M.R.; Bastick, P.; Briscoe, K.; Marx, G.; Swarbrick, A.; et al. Phase 2 study of circulating microRNA biomarkers in castration-resistant prostate cancer. Br. J. Cancer 2017, 116, 1002–1011. [Google Scholar] [CrossRef]
  242. Jeon, J.; Olkhov-Mitsel, E.; Xie, H.; Yao, C.Q.; Zhao, F.; Jahangiri, S.; Cuizon, C.; Scarcello, S.; Jeyapala, R.; Watson, J.D.; et al. Temporal Stability and Prognostic Biomarker Potential of the Prostate Cancer Urine miRNA Transcriptome. J. Natl. Cancer Inst. 2020, 112, 247–255. [Google Scholar] [CrossRef]
  243. Rodríguez, M.; Bajo-Santos, C.; Hessvik, N.P.; Lorenz, S.; Fromm, B.; Berge, V.; Sandvig, K.; Linē, A.; Llorente, A. Identification of non-invasive miRNAs biomarkers for prostate cancer by deep sequencing analysis of urinary exosomes. Mol. Cancer 2017, 16, 156. [Google Scholar] [CrossRef]
  244. Fredsøe, J.; Rasmussen, A.K.I.; Thomsen, A.R.; Mouritzen, P.; Høyer, S.; Borre, M.; Ørntoft, T.F.; Sørensen, K.D. Diagnostic and Prognostic MicroRNA Biomarkers for Prostate Cancer in Cell-free Urine. Eur. Urol. Focus 2018, 4, 825–833. [Google Scholar] [CrossRef]
  245. Koppers-Lalic, D.; Hackenberg, M.; de Menezes, R.; Misovic, B.; Wachalska, M.; Geldof, A.; Zini, N.; de Reijke, T.; Wurdinger, T.; Vis, A.; et al. Non-invasive prostate cancer detection by measuring miRNA variants (isomiRs) in urine extracellular vesicles. Oncotarget 2016, 7, 22566–22578. [Google Scholar] [CrossRef]
  246. Salido-Guadarrama, A.I.; Morales-Montor, J.G.; Rangel-Escareño, C.; Langley, E.; Peralta-Zaragoza, O.; Cruz Colin, J.L.; Rodriguez-Dorantes, M. Urinary microRNA-based signature improves accuracy of detection of clinically relevant prostate cancer within the prostate-specific antigen grey zone. Mol. Med. Rep. 2016, 13, 4549–4560. [Google Scholar] [CrossRef]
  247. Lenzo, N.P.; Meyrick, D.; Turner, J.H. Review of Gallium-68 PSMA PET/CT Imaging in the Management of Prostate Cancer. Diagnostics 2018, 8, 16. [Google Scholar] [CrossRef] [PubMed]
  248. Bostwick, D.G.; Pacelli, A.; Blute, M.; Roche, P.; Murphy, G.P. Prostate specific membrane antigen expression in prostatic intraepithelial neoplasia and adenocarcinoma: A study of 184 cases. Cancer 1998, 82, 2256–2261. [Google Scholar] [CrossRef]
  249. Kusumi, T.; Koie, T.; Tanaka, M.; Matsumoto, K.; Sato, F.; Kusumi, A.; Ohyama, C.; Kijima, H. Immunohistochemical detection of carcinoma in radical prostatectomy specimens following hormone therapy. Pathol. Int. 2008, 58, 687–694. [Google Scholar] [CrossRef] [PubMed]
  250. Mannweiler, S.; Amersdorfer, P.; Trajanoski, S.; Terrett, J.A.; King, D.; Mehes, G. Heterogeneity of prostate-specific membrane antigen (PSMA) expression in prostate carcinoma with distant metastasis. Pathol. Oncol. Res. POR 2009, 15, 167–172. [Google Scholar] [CrossRef] [PubMed]
  251. Ananias, H.J.; van den Heuvel, M.C.; Helfrich, W.; de Jong, I.J. Expression of the gastrin-releasing peptide receptor, the prostate stem cell antigen and the prostate-specific membrane antigen in lymph node and bone metastases of prostate cancer. Prostate 2009, 69, 1101–1108. [Google Scholar] [CrossRef]
  252. Elsässer-Beile, U.; Bühler, P.; Wolf, P. Targeted therapies for prostate cancer against the prostate specific membrane antigen. Curr. Drug Targets 2009, 10, 118–125. [Google Scholar] [CrossRef]
  253. Tasch, J.; Gong, M.; Sadelain, M.; Heston, W.D. A unique folate hydrolase, prostate-specific membrane antigen (PSMA): A target for immunotherapy? Crit. Rev. Immunol. 2001, 21, 249–261. [Google Scholar] [CrossRef]
  254. Sokoloff, R.L.; Norton, K.C.; Gasior, C.L.; Marker, K.M.; Grauer, L.S. A dual-monoclonal sandwich assay for prostate-specific membrane antigen: Levels in tissues, seminal fluid and urine. Prostate 2000, 43, 150–157. [Google Scholar] [CrossRef]
  255. Violet, J.; Jackson, P.; Ferdinandus, J.; Sandhu, S.; Akhurst, T.; Iravani, A.; Kong, G.; Kumar, A.R.; Thang, S.P.; Eu, P.; et al. Dosimetry of 177Lu-PSMA-617 in Metastatic Castration-Resistant Prostate Cancer: Correlations between Pretherapeutic Imaging and Whole-Body Tumor Dosimetry with Treatment Outcomes. J. Nucl. Med. 2019, 60, 517–523. [Google Scholar] [CrossRef]
  256. Afshar-Oromieh, A.; Avtzi, E.; Giesel, F.L.; Holland-Letz, T.; Linhart, H.G.; Eder, M.; Eisenhut, M.; Boxler, S.; Hadaschik, B.A.; Kratochwil, C.; et al. The diagnostic value of PET/CT imaging with the 68Ga-labelled PSMA ligand HBED-CC in the diagnosis of recurrent prostate cancer. Eur. J. Nucl. Med. Mol. Imaging 2015, 42, 197–209. [Google Scholar] [CrossRef]
  257. Maurer, T.; Gschwend, J.E.; Rauscher, I.; Souvatzoglou, M.; Haller, B.; Weirich, G.; Wester, H.J.; Heck, M.; Kübler, H.; Beer, A.J.; et al. Diagnostic Efficacy of 68Gallium-PSMA Positron Emission Tomography Compared to Conventional Imaging for Lymph Node Staging of 130 Consecutive Patients with Intermediate to High Risk Prostate Cancer. J. Urol. 2016, 195, 1436–1443. [Google Scholar] [CrossRef] [PubMed]
  258. Budäus, L.; Leyh-Bannurah, S.R.; Salomon, G.; Michl, U.; Heinzer, H.; Huland, H.; Graefen, M.; Steuber, T.; Rosenbaum, C. Initial Experience of 68Ga-PSMA PET/CT Imaging in High-risk Prostate Cancer Patients Prior to Radical Prostatectomy. Eur. Urol. 2016, 69, 393–396. [Google Scholar] [CrossRef] [PubMed]
  259. Afshar-Oromieh, A.; Malcher, A.; Eder, M.; Eisenhut, M.; Linhart, H.G.; Hadaschik, B.A.; Holland-Letz, T.; Giesel, F.L.; Kratochwil, C.; Haufe, S.; et al. PET imaging with a [68Ga]gallium-labelled PSMA ligand for the diagnosis of prostate cancer: Biodistribution in humans and first evaluation of tumour lesions. Eur. J. Nucl. Med. Mol. Imaging 2013, 40, 486–495. [Google Scholar] [CrossRef] [PubMed]
  260. Ferreira, G.; Iravani, A.; Hofman, M.S.; Hicks, R.J. Intra-individual comparison of 68Ga-PSMA-11 and 18F-DCFPyL normal-organ biodistribution. Cancer Imaging 2019, 19, 23. [Google Scholar] [CrossRef]
  261. Kroenke, M.; Mirzoyan, L.; Horn, T.; Peeken, J.C.; Wurzer, A.; Wester, H.J.; Makowski, M.; Weber, W.A.; Eiber, M.; Rauscher, I. Matched-Pair Comparison of 68Ga-PSMA-11 and (18)F-rhPSMA-7 PET/CT in Patients with Primary and Biochemical Recurrence of Prostate Cancer: Frequency of Non-Tumor-Related Uptake and Tumor Positivity. J. Nucl. Med. 2021, 62, 1082–1088. [Google Scholar] [CrossRef] [PubMed]
  262. Huang, S.; Ong, S.; McKenzie, D.; Mirabelli, A.; Chen, D.C.; Chengodu, T.; Murphy, D.G.; Hofman, M.S.; Lawrentschuk, N.; Perera, M. Comparison of 18F-based PSMA radiotracers with [68Ga]Ga-PSMA-11 in PET/CT imaging of prostate cancer-a systematic review and meta-analysis. Prostate Cancer Prostatic Dis. 2023. [Google Scholar] [CrossRef] [PubMed]
  263. Sartor, O.; de Bono, J.; Chi, K.N.; Fizazi, K.; Herrmann, K.; Rahbar, K.; Tagawa, S.T.; Nordquist, L.T.; Vaishampayan, N.; El-Haddad, G.; et al. Lutetium-177-PSMA-617 for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2021, 385, 1091–1103. [Google Scholar] [CrossRef] [PubMed]
  264. Zechmann, C.M.; Afshar-Oromieh, A.; Armor, T.; Stubbs, J.B.; Mier, W.; Hadaschik, B.; Joyal, J.; Kopka, K.; Debus, J.; Babich, J.W.; et al. Radiation dosimetry and first therapy results with a 124I/131I-labeled small molecule (MIP-1095) targeting PSMA for prostate cancer therapy. Eur. J. Nucl. Med. Mol. Imaging 2014, 41, 1280–1292. [Google Scholar] [CrossRef] [PubMed]
  265. Ahmadzadehfar, H.; Eppard, E.; Kürpig, S.; Fimmers, R.; Yordanova, A.; Schlenkhoff, C.D.; Gärtner, F.; Rogenhofer, S.; Essler, M. Therapeutic response and side effects of repeated radioligand therapy with 177Lu-PSMA-DKFZ-617 of castrate-resistant metastatic prostate cancer. Oncotarget 2016, 7, 12477–12488. [Google Scholar] [CrossRef]
  266. Heck, M.M.; Retz, M.; D’Alessandria, C.; Rauscher, I.; Scheidhauer, K.; Maurer, T.; Storz, E.; Janssen, F.; Schottelius, M.; Wester, H.J.; et al. Systemic Radioligand Therapy with 177Lu Labeled Prostate Specific Membrane Antigen Ligand for Imaging and Therapy in Patients with Metastatic Castration Resistant Prostate Cancer. J. Urol. 2016, 196, 382–391. [Google Scholar] [CrossRef]
  267. Hofman, M.S.; Emmett, L.; Sandhu, S.; Iravani, A.; Buteau, J.P.; Joshua, A.M.; Goh, J.C.; Pattison, D.A.; Tan, T.H.; Kirkwood, I.D.; et al. Overall survival with [177Lu]Lu-PSMA-617 versus cabazitaxel in metastatic castration-resistant prostate cancer (TheraP): Secondary outcomes of a randomised, open-label, phase 2 trial. Lancet Oncol. 2023, 25, 99–107. [Google Scholar] [CrossRef]
  268. National Comprehensive Cancer Network. NCCN Clinical Practice Guidelines in Prostate Cancer (Version 4.2023). Available online: https://www.nccn.org/professionals/physician_gls/pdf/prostate.pdf (accessed on 23 November 2023).
  269. Gillessen, S.; Bossi, A.; Davis, I.D.; de Bono, J.; Fizazi, K.; James, N.D.; Mottet, N.; Shore, N.; Small, E.; Smith, M.; et al. Management of Patients with Advanced Prostate Cancer. Part I: Intermediate-/High-risk and Locally Advanced Disease, Biochemical Relapse, and Side Effects of Hormonal Treatment: Report of the Advanced Prostate Cancer Consensus Conference 2022. Eur. Urol. 2023, 83, 267–293. [Google Scholar] [CrossRef]
  270. Evangelista, L.; Maurer, T.; van der Poel, H.; Alongi, F.; Kunikowska, J.; Laudicella, R.; Fanti, S.; Hofman, M.S. [68Ga]Ga-PSMA Versus [18F]PSMA Positron Emission Tomography/Computed Tomography in the Staging of Primary and Recurrent Prostate Cancer. A Systematic Review of the Literature. Eur. Urol. Oncol. 2022, 5, 273–282. [Google Scholar] [CrossRef]
  271. Sathekge, M.M.; Bruchertseifer, F.; Vorster, M.; Morgenstern, A.; Lawal, I.O. Global experience with PSMA-based alpha therapy in prostate cancer. Eur. J. Nucl. Med. Mol. Imaging 2021, 49, 30–46. [Google Scholar] [CrossRef]
  272. Afshar-Oromieh, A.; Hetzheim, H.; Kratochwil, C.; Benesova, M.; Eder, M.; Neels, O.C.; Eisenhut, M.; Kübler, W.; Holland-Letz, T.; Giesel, F.L.; et al. The Theranostic PSMA Ligand PSMA-617 in the Diagnosis of Prostate Cancer by PET/CT: Biodistribution in Humans, Radiation Dosimetry, and First Evaluation of Tumor Lesions. J. Nucl. Med. 2015, 56, 1697–1705. [Google Scholar] [CrossRef] [PubMed]
  273. Banerjee, S.; Pillai, M.R.; Knapp, F.F. Lutetium-177 therapeutic radiopharmaceuticals: Linking chemistry, radiochemistry, and practical applications. Chem. Rev. 2015, 115, 2934–2974. [Google Scholar] [CrossRef] [PubMed]
  274. Suman, S.; Parghane, R.V.; Joshi, A.; Prabhash, K.; Talole, S.; Basu, S. Combined 177Lu-PSMA-617 PRLT and abiraterone acetate versus 177Lu-PSMA-617 PRLT monotherapy in metastatic castration-resistant prostate cancer: An observational study comparing the response and durability. Prostate 2021, 81, 1225–1234. [Google Scholar] [CrossRef] [PubMed]
  275. Fong, P.C.; Boss, D.S.; Yap, T.A.; Tutt, A.; Wu, P.; Mergui-Roelvink, M.; Mortimer, P.; Swaisland, H.; Lau, A.; O’Connor, M.J.; et al. Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers. N. Engl. J. Med. 2009, 361, 123–134. [Google Scholar] [CrossRef] [PubMed]
  276. Menear, K.A.; Adcock, C.; Boulter, R.; Cockcroft, X.L.; Copsey, L.; Cranston, A.; Dillon, K.J.; Drzewiecki, J.; Garman, S.; Gomez, S.; et al. 4-[3-(4-cyclopropanecarbonylpiperazine-1-carbonyl)-4-fluorobenzyl]-2H-phthalazin-1-one: A novel bioavailable inhibitor of poly(ADP-ribose) polymerase-1. J. Med. Chem. 2008, 51, 6581–6591. [Google Scholar] [CrossRef] [PubMed]
  277. Slade, D. PARP and PARG inhibitors in cancer treatment. Genes Dev. 2020, 34, 360–394. [Google Scholar] [CrossRef]
  278. D’Andrea, A.D. Mechanisms of PARP inhibitor sensitivity and resistance. DNA Repair 2018, 71, 172–176. [Google Scholar] [CrossRef]
  279. Ray Chaudhuri, A.; Nussenzweig, A. The multifaceted roles of PARP1 in DNA repair and chromatin remodelling. Nat. Rev. Mol. Cell Biol. 2017, 18, 610–621. [Google Scholar] [CrossRef]
  280. Pommier, Y.; O’Connor, M.J.; de Bono, J. Laying a trap to kill cancer cells: PARP inhibitors and their mechanisms of action. Sci. Transl. Med. 2016, 8, 362ps317. [Google Scholar] [CrossRef]
  281. Topatana, W.; Juengpanich, S.; Li, S.; Cao, J.; Hu, J.; Lee, J.; Suliyanto, K.; Ma, D.; Zhang, B.; Chen, M.; et al. Advances in synthetic lethality for cancer therapy: Cellular mechanism and clinical translation. J. Hematol. Oncol. 2020, 13, 118. [Google Scholar] [CrossRef]
  282. Santiago-O’Farrill, J.M.; Weroha, S.J.; Hou, X.; Oberg, A.L.; Heinzen, E.P.; Maurer, M.J.; Pang, L.; Rask, P.; Amaravadi, R.K.; Becker, S.E.; et al. Poly(adenosine diphosphate ribose) polymerase inhibitors induce autophagy-mediated drug resistance in ovarian cancer cells, xenografts, and patient-derived xenograft models. Cancer 2020, 126, 894–907. [Google Scholar] [CrossRef]
  283. Marshall, C.H.; Sokolova, A.O.; McNatty, A.L.; Cheng, H.H.; Eisenberger, M.A.; Bryce, A.H.; Schweizer, M.T.; Antonarakis, E.S. Differential Response to Olaparib Treatment Among Men with Metastatic Castration-resistant Prostate Cancer Harboring BRCA1 or BRCA2 Versus ATM Mutations. Eur. Urol. 2019, 76, 452–458. [Google Scholar] [CrossRef]
  284. Schmid, S.; Omlin, A.; Higano, C.; Sweeney, C.; Martinez Chanza, N.; Mehra, N.; Kuppen, M.C.P.; Beltran, H.; Conteduca, V.; Vargas Pivato de Almeida, D.; et al. Activity of Platinum-Based Chemotherapy in Patients with Advanced Prostate Cancer with and without DNA Repair Gene Aberrations. JAMA Netw. Open 2020, 3, e2021692. [Google Scholar] [CrossRef]
  285. Quigley, D.; Alumkal, J.J.; Wyatt, A.W.; Kothari, V.; Foye, A.; Lloyd, P.; Aggarwal, R.; Kim, W.; Lu, E.; Schwartzman, J.; et al. Analysis of Circulating Cell-Free DNA Identifies Multiclonal Heterogeneity of BRCA2 Reversion Mutations Associated with Resistance to PARP Inhibitors. Cancer Discov. 2017, 7, 999–1005. [Google Scholar] [CrossRef]
  286. Mateo, J.; Hall, E.; Sandhu, S.; Omlin, A.G.; Miranda, S.; Carreira, S.; Goodall, J.; Gillman, A.; Mossop, H.; Ralph, C.; et al. Antitumour Activity of the Parp Inhibitor Olaparib in Unselected Sporadic Castration-Resistant Prostate Cancer (Crpc) in the Toparp Trial. Ann. Oncol. 2014, 25, v1. [Google Scholar] [CrossRef]
  287. Lowrance, W.; Dreicer, R.; Jarrard, D.F.; Scarpato, K.R.; Kim, S.K.; Kirkby, E.; Buckley, D.I.; Griffin, J.C.; Cookson, M.S. Updates to Advanced Prostate Cancer: AUA/SUO Guideline (2023). J. Urol. 2023, 209, 1082–1090. [Google Scholar] [CrossRef]
  288. Hussain, M.; Daignault-Newton, S.; Twardowski, P.W.; Albany, C.; Stein, M.N.; Kunju, L.P.; Siddiqui, J.; Wu, Y.M.; Robinson, D.; Lonigro, R.J.; et al. Targeting Androgen Receptor and DNA Repair in Metastatic Castration-Resistant Prostate Cancer: Results from NCI 9012. J. Clin. Oncol. 2018, 36, 991–999. [Google Scholar] [CrossRef]
  289. Hopkins, T.A.; Shi, Y.; Rodriguez, L.E.; Solomon, L.R.; Donawho, C.K.; DiGiammarino, E.L.; Panchal, S.C.; Wilsbacher, J.L.; Gao, W.; Olson, A.M.; et al. Mechanistic Dissection of PARP1 Trapping and the Impact on In Vivo Tolerability and Efficacy of PARP Inhibitors. Mol. Cancer Res. MCR 2015, 13, 1465–1477. [Google Scholar] [CrossRef] [PubMed]
  290. de Bono, J.; Mateo, J.; Fizazi, K.; Saad, F.; Shore, N.; Sandhu, S.; Chi, K.N.; Sartor, O.; Agarwal, N.; Olmos, D.; et al. Olaparib for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2020, 382, 2091–2102. [Google Scholar] [CrossRef]
  291. Hussain, M.; Mateo, J.; Fizazi, K.; Saad, F.; Shore, N.; Sandhu, S.; Chi, K.N.; Sartor, O.; Agarwal, N.; Olmos, D.; et al. Survival with Olaparib in Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2020, 383, 2345–2357. [Google Scholar] [CrossRef] [PubMed]
  292. Clarke, N.W.; Armstrong, A.J.; Thiery-Vuillemin, A.; Oya, M.; Shore, N.; Loredo, E.; Procopio, G.; Menezes, J.d.; Girotto, G.; Arslan, C.; et al. Abiraterone and Olaparib for Metastatic Castration-Resistant Prostate Cancer. NEJM Evid. 2022, 1, EVIDoa2200043. [Google Scholar] [CrossRef]
  293. Antonarakis, E.S.; Park, S.H.; Goh, J.C.; Shin, S.J.; Lee, J.L.; Mehra, N.; McDermott, R.; Sala-Gonzalez, N.; Fong, P.C.; Greil, R.; et al. Pembrolizumab Plus Olaparib for Patients with Previously Treated and Biomarker-Unselected Metastatic Castration-Resistant Prostate Cancer: The Randomized, Open-Label, Phase III KEYLYNK-010 Trial. J. Clin. Oncol. 2023, 41, 3839–3850. [Google Scholar] [CrossRef] [PubMed]
  294. Chi, K.N.; Rathkopf, D.; Smith, M.R.; Efstathiou, E.; Attard, G.; Olmos, D.; Lee, J.Y.; Small, E.J.; Pereira de Santana Gomes, A.J.; Roubaud, G.; et al. Niraparib and Abiraterone Acetate for Metastatic Castration-Resistant Prostate Cancer. J. Clin. Oncol. 2023, 41, 3339–3351. [Google Scholar] [CrossRef] [PubMed]
  295. Fizazi, K.; Piulats, J.M.; Reaume, M.N.; Ostler, P.; McDermott, R.; Gingerich, J.R.; Pintus, E.; Sridhar, S.S.; Bambury, R.M.; Emmenegger, U.; et al. Rucaparib or Physician’s Choice in Metastatic Prostate Cancer. N. Engl. J. Med. 2023, 388, 719–732. [Google Scholar] [CrossRef] [PubMed]
  296. Agarwal, N.; Azad, A.A.; Carles, J.; Fay, A.P.; Matsubara, N.; Heinrich, D.; Szczylik, C.; De Giorgi, U.; Young Joung, J.; Fong, P.C.C.; et al. Talazoparib plus enzalutamide in men with first-line metastatic castration-resistant prostate cancer (TALAPRO-2): A randomised, placebo-controlled, phase 3 trial. Lancet 2023, 402, 291–303. [Google Scholar] [CrossRef] [PubMed]
  297. de Bono, J.S.; Mehra, N.; Scagliotti, G.V.; Castro, E.; Dorff, T.; Stirling, A.; Stenzl, A.; Fleming, M.T.; Higano, C.S.; Saad, F.; et al. Talazoparib monotherapy in metastatic castration-resistant prostate cancer with DNA repair alterations (TALAPRO-1): An open-label, phase 2 trial. Lancet Oncol. 2021, 22, 1250–1264. [Google Scholar] [CrossRef] [PubMed]
  298. Roubaud, G.; Özgüroğlu, M.; Penel, N.; Matsubara, N.; Mehra, N.; Kolinsky, M.P.; Procopio, G.; Feyerabend, S.; Joung, J.Y.; Gravis, G.; et al. Olaparib tolerability and common adverse-event management in patients with metastatic castration-resistant prostate cancer: Further analyses from the PROfound study. Eur. J. Cancer 2022, 170, 73–84. [Google Scholar] [CrossRef]
  299. Zhang, T.; George, D.J.; Armstrong, A.J. Precision Medicine Approaches When Prostate Cancer Akts Up. Clin. Cancer Res. 2019, 25, 901–903. [Google Scholar] [CrossRef]
  300. Turnham, D.J.; Bullock, N.; Dass, M.S.; Staffurth, J.N.; Pearson, H.B. The PTEN Conundrum: How to Target PTEN-Deficient Prostate Cancer. Cells 2020, 9, 2342. [Google Scholar] [CrossRef]
  301. Blake, J.F.; Xu, R.; Bencsik, J.R.; Xiao, D.; Kallan, N.C.; Schlachter, S.; Mitchell, I.S.; Spencer, K.L.; Banka, A.L.; Wallace, E.M.; et al. Discovery and preclinical pharmacology of a selective ATP-competitive Akt inhibitor (GDC-0068) for the treatment of human tumors. J. Med. Chem. 2012, 55, 8110–8127. [Google Scholar] [CrossRef]
  302. Sweeney, C.; Bracarda, S.; Sternberg, C.N.; Chi, K.N.; Olmos, D.; Sandhu, S.; Massard, C.; Matsubara, N.; Alekseev, B.; Parnis, F.; et al. Ipatasertib plus abiraterone and prednisolone in metastatic castration-resistant prostate cancer (IPATential150): A multicentre, randomised, double-blind, phase 3 trial. Lancet 2021, 398, 131–142. [Google Scholar] [CrossRef]
  303. de Bono, J.S.; Bracarda, S.; Sternberg, C.N.; Chi, K.N.; Olmos, D.; Sandhu, S.; Massard, C.; Matsubara, N.; Alekseev, B.; Gafanov, R.; et al. LBA4 IPATential150: Phase III study of ipatasertib (ipat) plus abiraterone (abi) vs placebo (pbo) plus abi in metastatic castration-resistant prostate cancer (mCRPC). Ann. Oncol. 2020, 31, S1153–S1154. [Google Scholar] [CrossRef]
  304. ClinicalTrials.gov. Home Page. Available online: https://clinicaltrials.gov/study/NCT05348577 (accessed on 17 October 2023).
  305. ClinicalTrials.gov. Home Page. Available online: https://clinicaltrials.gov/study/NCT04493853 (accessed on 17 October 2023).
  306. Pardoll, D.M. The blockade of immune checkpoints in cancer immunotherapy. Nat. Rev. Cancer 2012, 12, 252–264. [Google Scholar] [CrossRef]
  307. Mahoney, K.M.; Rennert, P.D.; Freeman, G.J. Combination cancer immunotherapy and new immunomodulatory targets. Nat. Rev. Drug Discov. 2015, 14, 561–584. [Google Scholar] [CrossRef]
  308. Agata, Y.; Kawasaki, A.; Nishimura, H.; Ishida, Y.; Tsubata, T.; Yagita, H.; Honjo, T. Expression of the PD-1 antigen on the surface of stimulated mouse T and B lymphocytes. Int. Immunol. 1996, 8, 765–772. [Google Scholar] [CrossRef]
  309. Rescigno, P.; de Bono, J.S. Immunotherapy for lethal prostate cancer. Nat. Rev. Urol. 2019, 16, 69–70. [Google Scholar] [CrossRef] [PubMed]
  310. He, Y.; Wang, L.; Wei, T.; Xiao, Y.T.; Sheng, H.; Su, H.; Hollern, D.P.; Zhang, X.; Ma, J.; Wen, S.; et al. FOXA1 overexpression suppresses interferon signaling and immune response in cancer. J. Clin. Investig. 2021, 131, e147025. [Google Scholar] [CrossRef] [PubMed]
  311. Boyiadzis, M.M.; Kirkwood, J.M.; Marshall, J.L.; Pritchard, C.C.; Azad, N.S.; Gulley, J.L. Significance and implications of FDA approval of pembrolizumab for biomarker-defined disease. J. Immunother. Cancer 2018, 6, 35. [Google Scholar] [CrossRef] [PubMed]
  312. Taube, J.M.; Klein, A.; Brahmer, J.R.; Xu, H.; Pan, X.; Kim, J.H.; Chen, L.; Pardoll, D.M.; Topalian, S.L.; Anders, R.A. Association of PD-1, PD-1 ligands, and other features of the tumor immune microenvironment with response to anti-PD-1 therapy. Clin. Cancer Res. 2014, 20, 5064–5074. [Google Scholar] [CrossRef] [PubMed]
  313. Sfanos, K.S.; Bruno, T.C.; Meeker, A.K.; De Marzo, A.M.; Isaacs, W.B.; Drake, C.G. Human prostate-infiltrating CD8+ T lymphocytes are oligoclonal and PD-1+. Prostate 2009, 69, 1694–1703. [Google Scholar] [CrossRef] [PubMed]
  314. Yu, E.Y.; Kolinsky, M.P.; Berry, W.R.; Retz, M.; Mourey, L.; Piulats, J.M.; Appleman, L.J.; Romano, E.; Gravis, G.; Gurney, H.; et al. Pembrolizumab Plus Docetaxel and Prednisone in Patients with Metastatic Castration-resistant Prostate Cancer: Long-term Results from the Phase 1b/2 KEYNOTE-365 Cohort B Study. Eur. Urol. 2022, 82, 22–30. [Google Scholar] [CrossRef]
  315. Slovin, S.F.; Higano, C.S.; Hamid, O.; Tejwani, S.; Harzstark, A.; Alumkal, J.J.; Scher, H.I.; Chin, K.; Gagnier, P.; McHenry, M.B.; et al. Ipilimumab alone or in combination with radiotherapy in metastatic castration-resistant prostate cancer: Results from an open-label, multicenter phase I/II study. Ann. Oncol. 2013, 24, 1813–1821. [Google Scholar] [CrossRef]
  316. Kwon, E.D.; Drake, C.G.; Scher, H.I.; Fizazi, K.; Bossi, A.; van den Eertwegh, A.J.; Krainer, M.; Houede, N.; Santos, R.; Mahammedi, H.; et al. Ipilimumab versus placebo after radiotherapy in patients with metastatic castration-resistant prostate cancer that had progressed after docetaxel chemotherapy (CA184-043): A multicentre, randomised, double-blind, phase 3 trial. Lancet Oncol. 2014, 15, 700–712. [Google Scholar] [CrossRef]
  317. Killock, D. Prostate cancer: Mixed responses to ipilimumab. Nat. Rev. Urol. 2014, 11, 305. [Google Scholar] [CrossRef] [PubMed]
  318. Parker, C. A near miss for prostate cancer immunotherapy. Lancet Oncol. 2014, 15, 669–671. [Google Scholar] [CrossRef]
  319. ClinicalTrials.gov. Home Page. Available online: https://clinicaltrials.gov/study/NCT03570619 (accessed on 17 October 2023).
  320. Nguyen, B.; Mota, J.M.; Nandakumar, S.; Stopsack, K.H.; Weg, E.; Rathkopf, D.; Morris, M.J.; Scher, H.I.; Kantoff, P.W.; Gopalan, A.; et al. Pan-cancer Analysis of CDK12 Alterations Identifies a Subset of Prostate Cancers with Distinct Genomic and Clinical Characteristics. Eur. Urol. 2020, 78, 671–679. [Google Scholar] [CrossRef] [PubMed]
  321. Tosoian, J.J.; Birer, S.R.; Jeffrey Karnes, R.; Zhang, J.; Davicioni, E.; Klein, E.E.; Freedland, S.J.; Weinmann, S.; Trock, B.J.; Dess, R.T.; et al. Performance of clinicopathologic models in men with high risk localized prostate cancer: Impact of a 22-gene genomic classifier. Prostate Cancer Prostatic Dis. 2020, 23, 646–653. [Google Scholar] [CrossRef] [PubMed]
  322. Graham, L.S.; Montgomery, B.; Cheng, H.H.; Yu, E.Y.; Nelson, P.S.; Pritchard, C.; Erickson, S.; Alva, A.; Schweizer, M.T. Mismatch repair deficiency in metastatic prostate cancer: Response to PD-1 blockade and standard therapies. PLoS ONE 2020, 15, e0233260. [Google Scholar] [CrossRef] [PubMed]
  323. Le, D.T.; Uram, J.N.; Wang, H.; Bartlett, B.R.; Kemberling, H.; Eyring, A.D.; Skora, A.D.; Luber, B.S.; Azad, N.S.; Laheru, D.; et al. PD-1 Blockade in Tumors with Mismatch-Repair Deficiency. N. Engl. J. Med. 2015, 372, 2509–2520. [Google Scholar] [CrossRef]
  324. Snyder, A.; Makarov, V.; Merghoub, T.; Yuan, J.; Zaretsky, J.M.; Desrichard, A.; Walsh, L.A.; Postow, M.A.; Wong, P.; Ho, T.S.; et al. Genetic basis for clinical response to CTLA-4 blockade in melanoma. N. Engl. J. Med. 2014, 371, 2189–2199. [Google Scholar] [CrossRef] [PubMed]
  325. Van Allen, E.M.; Miao, D.; Schilling, B.; Shukla, S.A.; Blank, C.; Zimmer, L.; Sucker, A.; Hillen, U.; Foppen, M.H.G.; Goldinger, S.M.; et al. Genomic correlates of response to CTLA-4 blockade in metastatic melanoma. Science 2015, 350, 207–211. [Google Scholar] [CrossRef]
  326. Schweizer, M.T.; Antonarakis, E.S.; Bismar, T.A.; Guedes, L.B.; Cheng, H.H.; Tretiakova, M.S.; Vakar-Lopez, F.; Klemfuss, N.; Konnick, E.Q.; Mostaghel, E.A.; et al. Genomic Characterization of Prostatic Ductal Adenocarcinoma Identifies a High Prevalence of DNA Repair Gene Mutations. JCO Precis. Oncol. 2019, 3, 1–9. [Google Scholar] [CrossRef] [PubMed]
  327. Antonarakis, E.S.; Shaukat, F.; Isaacsson Velho, P.; Kaur, H.; Shenderov, E.; Pardoll, D.M.; Lotan, T.L. Clinical Features and Therapeutic Outcomes in Men with Advanced Prostate Cancer and DNA Mismatch Repair Gene Mutations. Eur. Urol. 2019, 75, 378–382. [Google Scholar] [CrossRef] [PubMed]
  328. Zhu, J.; Tucker, M.; Marin, D.; Gupta, R.T.; Healy, P.; Humeniuk, M.; Jarvis, C.; Zhang, T.; McNamara, M.; George, D.J.; et al. Clinical utility of FoundationOne tissue molecular profiling in men with metastatic prostate cancer. Urol. Oncol. 2019, 37, 813.e1–813.e9. [Google Scholar] [CrossRef] [PubMed]
  329. Graf, R.P.; Fisher, V.; Weberpals, J.; Gjoerup, O.; Tierno, M.B.; Huang, R.S.P.; Sayegh, N.; Lin, D.I.; Raskina, K.; Schrock, A.B.; et al. Comparative Effectiveness of Immune Checkpoint Inhibitors vs Chemotherapy by Tumor Mutational Burden in Metastatic Castration-Resistant Prostate Cancer. JAMA Netw. Open 2022, 5, e225394. [Google Scholar] [CrossRef] [PubMed]
  330. Nourmohammadi Abadchi, S.; Sena, L.A.; Antonarakis, E.S.; Pritchard, C.C.; Eshleman, J.R.; Konnick, E.Q.; Salipante, S.J.; Shenderov, E.; Lotan, T.L. MLH1 Loss in Primary Prostate Cancer. JCO Precis. Oncol. 2023, 7, e2200611. [Google Scholar] [CrossRef] [PubMed]
  331. Venkatachalam, S.; McFarland, T.R.; Agarwal, N.; Swami, U. Immune Checkpoint Inhibitors in Prostate Cancer. Cancers 2021, 13, 2187. [Google Scholar] [CrossRef] [PubMed]
  332. Guedes, L.B.; Antonarakis, E.S.; Schweizer, M.T.; Mirkheshti, N.; Almutairi, F.; Park, J.C.; Glavaris, S.; Hicks, J.; Eisenberger, M.A.; De Marzo, A.M.; et al. MSH2 Loss in Primary Prostate Cancer. Clin. Cancer Res. 2017, 23, 6863–6874. [Google Scholar] [CrossRef] [PubMed]
  333. Chan, T.A.; Yarchoan, M.; Jaffee, E.; Swanton, C.; Quezada, S.A.; Stenzinger, A.; Peters, S. Development of tumor mutation burden as an immunotherapy biomarker: Utility for the oncology clinic. Ann. Oncol. 2019, 30, 44–56. [Google Scholar] [CrossRef]
  334. Lanka, S.M.; Zorko, N.A.; Antonarakis, E.S.; Barata, P.C. Metastatic Castration-Resistant Prostate Cancer, Immune Checkpoint Inhibitors, and Beyond. Curr. Oncol. 2023, 30, 4246–4256. [Google Scholar] [CrossRef]
  335. Rebuzzi, S.E.; Rescigno, P.; Catalano, F.; Mollica, V.; Vogl, U.M.; Marandino, L.; Massari, F.; Pereira Mestre, R.; Zanardi, E.; Signori, A.; et al. Immune Checkpoint Inhibitors in Advanced Prostate Cancer: Current Data and Future Perspectives. Cancers 2022, 14, 1245. [Google Scholar] [CrossRef] [PubMed]
  336. Massari, F.; Ciccarese, C.; Caliò, A.; Munari, E.; Cima, L.; Porcaro, A.B.; Novella, G.; Artibani, W.; Sava, T.; Eccher, A.; et al. Magnitude of PD-1, PD-L1 and T Lymphocyte Expression on Tissue from Castration-Resistant Prostate Adenocarcinoma: An Exploratory Analysis. Target. Oncol. 2016, 11, 345–351. [Google Scholar] [CrossRef] [PubMed]
  337. Gevensleben, H.; Dietrich, D.; Golletz, C.; Steiner, S.; Jung, M.; Thiesler, T.; Majores, M.; Stein, J.; Uhl, B.; Müller, S.; et al. The Immune Checkpoint Regulator PD-L1 Is Highly Expressed in Aggressive Primary Prostate Cancer. Clin. Cancer Res. 2016, 22, 1969–1977. [Google Scholar] [CrossRef] [PubMed]
  338. Bishop, J.L.; Sio, A.; Angeles, A.; Roberts, M.E.; Azad, A.A.; Chi, K.N.; Zoubeidi, A. PD-L1 is highly expressed in Enzalutamide resistant prostate cancer. Oncotarget 2015, 6, 234–242. [Google Scholar] [CrossRef]
Table 1. Frequency of genetic alterations in each stage of prostate cancer.
Table 1. Frequency of genetic alterations in each stage of prostate cancer.
ReferenceStageGene
ARTP53PTENRB1SPOPBRCA2BRCA1ATMCDK12ETS
[11]Localized081711130.36154
[27]Localized036.354.518.209.109.1072.7
[20]mCSPC5332451370.256
[10]mCRPC62.753.3409.3813.30.77.34.756.7
[28]Localized227122126124
mCSPC430187117126
mCRPC5248291851021111
[29]mCSPC0–530–4812–345–87–112–91–23–64–6
mCRPC41–5230–5032–406–136–117–121–25–64–6
[30]Localized1.28170.91231.272.446
mCSPC6332551370.58726
mCRPC594033131112581032
AR, androgen receptor; TP53, tumor protein 53 (p53); PTEN, phosphatase and tensin homolog; RB1, retinoblastoma protein 1; SPOP, speckle-type POZ protein; BRCA2, breast cancer 2 gene; BRCA1, breast cancer 1 gene; ATM, ataxia-telangiectasia mutated; CDK12, cyclin-dependent kinase 12; ETS, E26 transformation-specific transcription factor; mCSPC, metastatic castration-sensitive prostate cancer; mCRPC, metastatic castration-resistant prostate cancer.
Table 2. Overall survival for patients with single gene mutations in HRR-related genes.
Table 2. Overall survival for patients with single gene mutations in HRR-related genes.
GeneTreatmentnMedian OS (Months)HR (95% CI)
BRCA2Olaparib8124.80.59 (0.37–0.95)
ARSI4715.2
CDK12Olaparib6114.10.97 (0.57–1.71)
ARSI2811.5
ATMOlaparib6218.00.93 (0.53–1.75)
ARSI2415.6
BRCA1Olaparib811.70.42 (0.12–1.53)
ARSI59.4
CHECK2Olaparib716.60.87 (0.19–4.44)
ARSI517.1
PPP2R2AOlaparib68.15.11 (1.10–35.73)
ARSI4NE
RAD51BOlaparib4NENE
ARSI13.6
RAD54LOlaparib319.3NE
ARSI25.7
ARSI, androgen receptor signaling inhibitor; CI, confidence interval; HR, hazard ratio; HRR, homologous recombination repair; NE, not estimable; OS, overall survival.
Table 3. Liquid biopsy using microRNAs in the urine of patients with prostate cancer.
Table 3. Liquid biopsy using microRNAs in the urine of patients with prostate cancer.
Number of Samples (Normal/PCa)Altered miRNA Expression in Patients with PCa or a High Risk of PCa
ReferenceSamplesScreeningValidationUpregulated miRNAsDownregulated miRNAs
[193]DRE-urinary sediments10 PCa + 99 (50 high-risk vs. 49 low-risk)40 (11 high-risk vs. 29 low-risk)miR-3195,
let-7b-5p,
miR-144-3p,
miR-451a,
miR 148a-3p,
miR-512- 5p,
miR-431-5p
[194]Urinary exosomes29 (9/20)47 (19/28) miR-196a-5p,
miR-34a-5p,
miR-501-3p,
miR-92a-1-5p,
miR-143-3p
[195] Urinary exosomes208 (20/188)217 (20/197)miR-30c-5p,
miR-30b-5p,
miR-30a-5p,
miR-30a-3p,
miR-125b-5p
miR-375,
miR-21-5p,
miR-141-3p,
let-7c-5p,
miR-26b-5p,
miR-101-3p,
let-7b-5p
[196]DRE-urinary
extracellular vesicles (EVs)
13 (4/9)74 (26/48)miR-10a-5p,
miR-204-5p,
miR-30a-3p
miR-27a,
miR-328,
miR-483,
miR-125a,
miR-146a,
miR-195,
miR-145,
miR-150
DRE, digital rectal examination; PCa, prostate cancer.
Table 4. Clinical trials on PSMA-targeting radionuclide therapy for patients with mCSPC.
Table 4. Clinical trials on PSMA-targeting radionuclide therapy for patients with mCSPC.
DrugConcomitant Drug and/or TreatmentPhaseNumber of PatientsPrior TreatmentPrimary EndpointIdentifier
177Lu-DOTA-TLX591EBRT250RPPSA-PFSNCT05146973
177Lu-PSMA-617SBRT16 NCT05079698
177Lu-PSMA-617ADT31126 rPFSNCT04720157
177Lu-PSMA-617Docetaxel2140 Undetectable PSA rateNCT04343885
ADT, androgen deprivation therapy; EBRT, external beam radiotherapy; mCSPC, metastatic castration-sensitive prostate cancer; PFS, progression-free survival; PSA, prostate-specific antigen; PSMA, prostate-specific membrane antigen; RP, radical prostatectomy; rPFS, radiographic PFS; SBRT, stereotactic body radiotherapy.
Table 5. Clinical trials on PSMA-targeting radionuclide therapy for patients with CRPC.
Table 5. Clinical trials on PSMA-targeting radionuclide therapy for patients with CRPC.
DrugConcomitant Drug and/or TreatmentPhaseNumber of PatientsPrior TreatmentPrimary EndpointIdentifier
225Ac-PSMA-I&T 2100Taxane NCT05219500
225Ac-Labeled Antibody Targeting hk2 170ARSI ± taxane NCT04644770
177Lu-PSMA-617±ADT3450ARSIrPFSNCT04689828
177Lu-PSMA-I&T 3400ARSIrPFSNCT05204927
177Lu-PSMA-PNT2002 3415ARSIrPFSNCT04647526
177Lu-DOTA-rosopatamab 3387TaxanerPFSNCT04876651
177Lu-PSMA-617Abiraterone 58Taxane [274]
177Lu-PSMA-617Cabazitaxel½44ARSI + docetaxel NCT05340374
177Lu-PSMA-617Enzalutamide2160ADTPSA-PFSNCT04419402
177Lu-PSMA-617Olaparib152ARSI + taxane NCT03874884
177Lu-PSMA-617Pembrolizumab½37ARSIPSA responseNCT03658447
177Lu-PSMA-617Pembrolizumab143ARSI NCT03805594
225Ac-J591Pembrolizumab1/276ADT NCT04946370
225Ac-PSMA-617±177Lu-PSMA-617/I&T and/or ARSI130±177Lu-PSMA-617 and/or ARSI NCT04597411
225Ac-J591 + 177Lu-PSMA-I&T ½33ARSI and taxane NCT04886986
ADT, androgen deprivation therapy; ARSI, androgen receptor signaling inhibitor; CRPC, castration-resistant prostate cancer; hk2, human kallikrein-2; I&T, imaging and therapy; PFS, progression-free survival; PSA, prostate specific antigen; PSMA, prostate-specific membrane antigen; rPFS, radiographic PFS.
Table 6. Efficacy of PARPis, as monotherapy or combination therapy, in each clinical trial.
Table 6. Efficacy of PARPis, as monotherapy or combination therapy, in each clinical trial.
ReferenceComparatorControlParticipantsHR for rPFSHR for OS
[290,291]OLAARSIBRCAm or ATMm0.34 (0.25–0.47)0.69 (0.50–0.97)
[292]OLA + ABIABIHRRm0.45 (0.31–0.65)Not shown
[293]OLA + PEMARSIBRCAm0.40 (0.21–0.79)0.52 (0.27–0.99)
[294]NIRA + ABIABIBRCAm0.53 (0.36–0.79)Not shown
[295]RUCADOC or ARSIBRCAm0.50 (0.36–0.69)0.81 (0.58–1.12)
[296]TALA + ENZENZHRRm0.46 (0.30–0.70)Not shown
[288]VELI + ABIABIDDRdNot estimatedNot estimated
ABI, abiraterone; ARSI, androgen receptor signaling inhibitor; ATMm, Participants with tumors harboring aberrations in ATM; BRCAm, Participants with tumors harboring aberrations in BRCA1/2; DDRd, Participants with tumors harboring DNA damage response deficiency; DOC, docetaxel; ENZ, enzalutamide; HR, hazard ratio; HRR, homologous recombination repair genes; HRRm, Participants with tumors harboring aberrations in HRR genes; NIRA, niraparib; OLA, olaparib; OS, overall survival; PEM, pembrolizumab; rPFS, radiographic progression-free survival; RUCA, rucaparib; TALA, talazoparib; VELI, veliparib.
Table 7. Adverse events associated with PARPis, used as monotherapy or combination therapy in each clinical trial.
Table 7. Adverse events associated with PARPis, used as monotherapy or combination therapy in each clinical trial.
TreatmentAnemiaNausea
ComparatorControlActive DrugControlActive DrugControl
[291,298]OLAARSI49.6 (22.7)15.4 (5.4)43.0 (1.6)20.8 (0)
[292]OLA + ABIABI46.0 (15.1)28.1 (0.3)16.4 (3.3)12.6 (0.3)
[293]OLA + PEMARSI46.0 (19.6)35.6 (1.5)2.7 (1.2)5.1 (0)
[294]NIRA + ABIABI46.2 (29.7)20.4 (7.6)23.6 (0.5)13.7 (0)
[295]RUCADOC or ARSI46.7 (23.7)17.7 (0.8)49.6 (2.6)19.2 (0.8)
[296]TALA + ENZENZ65.8 (46.5)20.6 (0.5)17.5 (4.2)12.5 (0.7)
[288]VELI + ABIABI20.3 (2.5)14.9 (1.4)53.2 (1.3)6.8 (0)
(%), All-grade (Grade ≥ 3). ABI, abiraterone; ARSI, androgen receptor signaling inhibitor; DOC, docetaxel; ENZ, enzalutamide; NIRA, niraparib; OLA, olaparib; PEM, pembrolizumab; RUCA, rucaparib; TALA, talazoparib; VELI, veliparib.
Table 8. Clinical trials on ICIs in patients with mCSPC.
Table 8. Clinical trials on ICIs in patients with mCSPC.
DrugsConcomitant Drug and/or TreatmentPhaseNumber of
Patients
Patient CharacteristicsPrimary Endpoint Identifier
Nivolumab
or ipilimumab
Docetaxel2/3135High-volume mCSPCOSNCT03879122 (PROSTRATEGY)
PembrolizumabEnzalutamide31232mCSPCrPFS
OS
NCT04191096 (KEYNOTE-991)
NivolumabDocetaxel260mCSPC with DRD
or inflamed tumors
PSA
response
NCT04126070
NivolumabCabozantinib + abiraterone
acetate
1b22De novo or recurrent mCSPCDose-limiting toxicities NCT04477512 (CABIOS)
CemiplimabDocetaxel220De novo mCSPCPSA responseNCT03951831 (REGN2810)
DurvalumabSBRT296mCSPC with OligometsPFSNCT03795207 (POSTCARD)
CSPC, castration-sensitive prostate cancer; OS, overall survival; rPFS, radiographic progression-free survival; PSA, prostate-specific antigen; DRD, DNA repair deficiency; NCT, ClinicalTrials.gov identifier.
Table 9. Clinical trials on ICIs in patients with mCRPC.
Table 9. Clinical trials on ICIs in patients with mCRPC.
DrugsPhaseNumber of PatientsPrior TreatmentPatient CharacteristicsPrimary EndpointIdentifier
Ipilimumab3988Docetaxel OSNCT00861614 (CA184-043)
Ipilimumab3837Chemotherapy-naïve OSNCT01057810 (CA184-095)
Atezolizumab1b35Sipuleucel-T
or enzalutamide
SafetyNCT01375842 (PCD4989g)
Nivolumab238DocetaxelGermline and somatic DRDPSA responseNCT03040791 (ImmunoProst trial)
Nivolumab215Local therapyMMD, MSIh, TMBh, CDK12 inactivationPSA responseNCT04019964
Pembrolizumab230ARSIMMD, or CDK12 inactivationPSA response
ORR
NCT04104893 (CHOMP trial)
Pembrolizumab250ARSIDRDrPFSNCT03248570
Pembrolizumab2258Docetaxel and one or more ARSIs ORRNCT02787005 (KEYNOTE-199)
Pembrolizumab2100AnyHigh mutational load, DRD, and/or MMDPSA response
ORR
NCT03506997 (PERSEUS1)
Nivolumab + ipilimumab2259Docetaxel ORR
rPFS
NCT02985957 (CheckMate 650)
Nivolumab + ipilimumab2175Systemic therapyDRD or MMD or TMBhPSA response
rPFS
CTC count
NCT03061539 (Neptunes)
Nivolumab + ipilimumab275ADTMMD and/or TMBhDisease control rateNCT04717154 (INSPIRE)
Nivolumab + ipilimumab240Systemic therapy or radiation therapyCDK12 inactivationPSA response
ORR
NCT03570619 (IMPACT)
Nivolumab + ipilimumab215AnyAR-V7PSA responseNCT02601014 (STARVE-PC)
mCRPC, metastatic castration-resistant prostate cancer; OS, overall survival; NCT, ClinicalTrials.gov identifier; ARSI, androgen receptor signaling inhibitor; DRD, DNA repair deficiency; PSA, prostate-specific antigen; MMD, mismatch repair deficiency; MSIh, microsatellite instability-high; TMBh, tumor mutational burden-high; ORR, overall response rate; rPFS, radiographic progression-free survival; CTC, circulating tumor cell; ADT, androgen deprivation therapy; AR-V7, androgen receptor variant 7.
Table 10. Clinical trials on combination therapy with ICIs in patients with mCRPC.
Table 10. Clinical trials on combination therapy with ICIs in patients with mCRPC.
DrugsConcomitant Drugs and/or TreatmentPhaseNumber of PatientsPrior TreatmentPrimary EndpointIdentifier
AtezolizumabSipuleucel-T1b37 SafetyNCT03024216
AtezolizumabRadium-2231b44ARSISafety
ORR
NCT02814669
AtezolizumabCabozantinib3580ARSIPFS
OS
NCT04446117 (CONTACT-02)
AtezolizumabEnzalutamide3759AbirateroneOSNCT03016312 (IMbassador250)
DurvalumabOlaparib217ARSIrPFS
PSA response
NCT02484404
DurvalumabTremelimumab252ARSIORRNCT02788773
NivolumabDocetaxel284ADT + ARSIORR
PSA response
NCT03338790 (CheckMate 9KD)
NivolumabPan-VEGFR/TIE2 tyrosine kinase inhibitor249ARSIPSA response
Adverse events
NCT04159896
PembrolizumabEnzalutamide31200Abiraterone-naïve CRPC or abiraterone for CSPCOS
rPFS
NCT03834493 (KEYNOTE-641)
PembrolizumabDocetaxel31030Chemotherapy-naïveOS
rPFS
NCT03834506 (KEYNOTE-921)
PembrolizumabOlaparib3529Docetaxel + ARSIOS
rPFS
NCT03834519 (KEYLYNK-010)
Pembrolizumab177Lu-PSMA-6171b37 Safety
PSA response
NCT03658447 (PRINCE)
Nivolumab + ipilimumab177Lu-PSMA-6172110DocetaxelPSA responseNCT05150236
mCRPC, metastatic castration-resistant prostate cancer; ARSI, androgen receptor signaling inhibitor; PFS, progression-free survival; OS, overall survival; NCT, ClinicalTrials.gov identifier; ADT, androgen deprivation therapy; ORR, overall response rate; PSA, prostate-specific antigen.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Maekawa, S.; Takata, R.; Obara, W. Molecular Mechanisms of Prostate Cancer Development in the Precision Medicine Era: A Comprehensive Review. Cancers 2024, 16, 523. https://doi.org/10.3390/cancers16030523

AMA Style

Maekawa S, Takata R, Obara W. Molecular Mechanisms of Prostate Cancer Development in the Precision Medicine Era: A Comprehensive Review. Cancers. 2024; 16(3):523. https://doi.org/10.3390/cancers16030523

Chicago/Turabian Style

Maekawa, Shigekatsu, Ryo Takata, and Wataru Obara. 2024. "Molecular Mechanisms of Prostate Cancer Development in the Precision Medicine Era: A Comprehensive Review" Cancers 16, no. 3: 523. https://doi.org/10.3390/cancers16030523

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