Next Article in Journal
Predictive Value of Baseline FDG-PET/CT for the Durable Response to Immune Checkpoint Inhibition in NSCLC Patients Using the Morphological and Metabolic Features of Primary Tumors
Next Article in Special Issue
Beyond Multiparametric MRI and towards Radiomics to Detect Prostate Cancer: A Machine Learning Model to Predict Clinically Significant Lesions
Previous Article in Journal
Interventional Oncolytic Immunotherapy with LTX-315 for Residual Tumor after Incomplete Radiofrequency Ablation of Liver Cancer
Previous Article in Special Issue
Prostate-Specific Antigen Doubling Time Kinetics following Radical Prostatectomy to Guide Need for Treatment Intervention: Validation of Low-Risk Recurrences
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biomarkers for the Detection and Risk Stratification of Aggressive Prostate Cancer

by
Samaneh Eickelschulte
1,2,3,
Anja Lisa Riediger
1,2,3,4,
Arlou Kristina Angeles
3,
Florian Janke
3,
Stefan Duensing
5,
Holger Sültmann
3,6 and
Magdalena Görtz
1,2,*
1
Junior Clinical Cooperation Unit, Multiparametric Methods for Early Detection of Prostate Cancer, German Cancer Research Center (DKFZ), 69120 Heidelberg, Germany
2
Department of Urology, University Hospital Heidelberg, 69120 Heidelberg, Germany
3
Division of Cancer Genome Research, German Cancer Research Center (DKFZ), National Center for Tumor Diseases (NCT), 69120 Heidelberg, Germany
4
Faculty of Biosciences, Heidelberg University, 69120 Heidelberg, Germany
5
Molecular Urooncology, Department of Urology, University Hospital Heidelberg, 69120 Heidelberg, Germany
6
German Cancer Consortium (DKTK), 69120 Heidelberg, Germany
*
Author to whom correspondence should be addressed.
Cancers 2022, 14(24), 6094; https://doi.org/10.3390/cancers14246094
Submission received: 17 November 2022 / Revised: 5 December 2022 / Accepted: 9 December 2022 / Published: 11 December 2022
(This article belongs to the Collection Biomarkers for Detection and Prognosis of Prostate Cancer)

Abstract

:

Simple Summary

Prostate cancer is a heterogeneous disease and a major cause of cancer deaths worldwide. The most widely used prostate cancer biomarker, prostate-specific antigen, lacks sensitivity and specificity in the diagnosis of malignant disease. Hence, novel tissue-based biomarkers have emerged for the detection and risk assessment of prostate cancer. Over the past years, liquid biopsy biomarkers introduced a new diagnostic concept to complement current tissue diagnosis strategies. Liquid biopsies non-invasively provide a characterization of heterogenous tumor profiles. Here, we highlight the most prominent tissue and liquid biopsy biomarkers for the detection and risk assessment of prostate cancer.

Abstract

Current strategies for the clinical management of prostate cancer are inadequate for a precise risk stratification between indolent and aggressive tumors. Recently developed tissue-based molecular biomarkers have refined the risk assessment of the disease. The characterization of tissue biopsy components and subsequent identification of relevant tissue-based molecular alterations have the potential to improve the clinical decision making and patient outcomes. However, tissue biopsies are invasive and spatially restricted due to tumor heterogeneity. Therefore, there is an urgent need for complementary diagnostic and prognostic options. Liquid biopsy approaches are minimally invasive with potential utility for the early detection, risk stratification, and monitoring of tumors. In this review, we focus on tissue and liquid biopsy biomarkers for early diagnosis and risk stratification of prostate cancer, including modifications on the genomic, epigenomic, transcriptomic, and proteomic levels. High-risk molecular alterations combined with orthogonal clinical parameters can improve the identification of aggressive tumors and increase patient survival.

Graphical Abstract

1. Introduction

Prostate cancer (PCa) is the second most frequent malignancy among men, accounting for 27% of all cancer diagnoses in males [1,2]. Although the survival rate of men with early diagnosed localized PCa can be as high as ~99%, men who are diagnosed with late-stage disease have a survival of only 30% over 5 years [3,4,5]. The detection of PCa is feasible using the blood-based biomarker prostate-specific antigen (PSA) and digital rectal examination (DRE) as a screening method. The diagnosis of PCa at an early stage has benefits, including the high possibility of cures, less aggressive treatment options, reduced disease progression to advanced or metastatic stages, and improved quality of life [6]. The widespread use of PSA as a screening method for PCa led to improved PCa diagnoses, with a shift towards earlier tumor stages. Within the European Randomized Study of Screening for Prostate Cancer (ERSPC), PSA screening was shown to reduce PCa death and metastasis in men aged 55–69 years by 29% and 30–42%, respectively [7,8].
However, this reduction in mortality and metastases is associated with considerable overdiagnosis and overtreatment rates of PCa. Moderately elevated PSA serum levels are controversial, as PSA is produced by the normal prostate, as well as PCa cells; therefore, they lack cancer sensitivity and specificity. Accordingly, PSA testing has led to many instances of men with elevated PSA levels with a subsequent negative PCa diagnosis. Considering the psychological burden of false-positive PSA values, only ~25% of men with elevated PSA (>4 ng/mL) have a positive prostate biopsy and show PCa incidences [9,10]. Non-invasive markers with the potential to distinguish between normal prostate and PCa could significantly reduce the number of invasive diagnostics in the form of prostate biopsies. A further challenge in PCa management is the lack of an adequate risk stratification scheme of detected PCa in order to guarantee an optimal, individually adapted therapy. Men with indolent tumor forms unlikely to progress to clinical significance require further diagnostic options [11]. The molecular characterization and analysis of tissue biomarkers are important for the risk stratification of PCa, complementing the histopathological evaluation of prostate biopsy samples [12,13]. Due to intra-tumoral heterogeneity, tumor molecular profiling by tissue biopsy is limited by finite sampling sites. Therefore, there is increasing interest in identifying tumor biomarkers in liquid biopsy samples to improve detection, clinical decision making, and clinical outcomes for patients. Being able to detect and evaluate the aggressiveness of PCa as early as possible can improve patient survival. This review aims to present a comprehensive overview on recent molecular markers for the early detection and risk stratification of aggressive PCa in both tissue and liquid biopsy.

2. Tissue Prostate Cancer Biomarkers

2.1. Genomic and Epigenomic Tissue Biomarkers

Whole genome sequencing (WGS) and characterization of genomic alterations revealed various novel tissue-based biomarkers associated with PCa [13]. On the other hand, epigenomic changes occur early during PCa development and are often acquired prior to somatic mutations or chromosomal aberrations [14]. Consequently, alterations in the epigenome have higher levels of occurrence compared to the most frequent genetic defects in PCa [15,16]. Both genomic and epigenomic alterations emerged as powerful tools to distinguish indolent tumors from those likely to progress.

2.1.1. DNA Mutations and Copy Number Alterations

Mutations in DNA damage repair (DDR) machinery are key mechanisms driving PCa tumorigenesis in a subset of prostate tumors [17,18]. Germline mutations in homologous recombination (HR)-mediated DDR genes, BRCA1 or BRCA2, are relatively rare (<5%) in unselected patients, but important prognostic factors for PCa [19,20,21,22]. Mutations in DDR genes through HR have been shown to be associated with aggressive tumor growth and increase the risk of PCa at a younger age [20,23]. Interestingly, somatic mutations in genes involved in DDR (i.e., BRCA1/2, ATM, and CDK12) can be found at a higher frequency of ~20% [24,25] and, therefore, play a substantial role in the risk stratification of PCa. Mutations leading to a loss of function in the tumor suppressor genes, TP53 and RB1, represent further genomic alterations associated with PCa progression [17,26,27,28,29]. TP53 mutations are frequently observed in primary PCa [30] with high-risk features and have an important prognostic value in the early detection of PCa [31,32,33,34]. Mutations in the androgen receptor (AR) and alteration of AR-regulated signaling pathways are detected in about one-third of metastatic PCa patients, ranking as the most frequently reported genomic alterations in PCa [34,35,36,37]. Mutations in SPOP, CHD1, PTEN, NKX3-1, FOXA1, and APC are other known genomic alterations in primary and advanced PCa (Figure 1) [29,38,39,40,41,42].
The fusion of TMPRSS2 and erythroblast transformation-specific (ETS)-related gene (ERG) is the most frequent (~50%) structural abnormality in PCa. Because of its high frequency and specificity, TMPRSS2-ERG is used as a diagnostic biomarker for PCa [43,44,45]. The PTEN loss often co-occurs with ERG rearrangements and is necessary for its carcinogenesis effect [20,46].
Copy number alterations (CNAs), either alone or in combination with mutations, can be used as a potential biomarker for the risk stratification of aggressive PCa [17,47]. The CNA analysis may have clinical applications in the prognosis of PCa independent of the Gleason score. It has been shown that tumors carrying few CNAs had a more favorable prognosis than did those containing a high number of CNAs [36,48]. Chromosomal gains affecting MYC, NCOA2, and AR are known to be present in high-risk PCa [36,39,49]. The amplification of MYC has been reported as the most frequent copy number gain in primary PCa (up to 30% in advanced disease) [3]. Deletions in GSTP1, CDKN1B, and ARID1A have been shown to be associated with the risk of hereditary PCa and metastasis formation [50,51,52,53]. Figure 2 shows known driver mutations and CNAs detected in tissue that are involved in the initiation, as well as the progression, of PCa. Furthermore, mutations and CNAs in the mitochondrial genome (mtDNA), resulting in the dysregulation of mitochondrial homeostasis, can play a role in tumor initiation, growth, and metastasis. Mutations and CNAs in mitochondrially encoded tRNA, CO1, ATP6, PC3, and ND genes can be potential prognostic biomarkers that may improve the early detection of aggressive PCa [54]. Overall, the heterogeneity of PCa is mirrored by the presence of multiple genomic alterations along PCa’s development. Identifying mutations and CNAs is essential for using these valuable tools for the detection and risk assessment of PCa.

2.1.2. DNA Methylation

DNA methylation, occurring at the 5’-carbon of cytosine residues (5-mC), is the most widely studied epigenomic modification in cancer. 5-mC resides at cytosine–guanine dinucleotides (CpGs), and the hypermethylation of promoter-associated CpG clusters, termed CpG islands, is linked to transcriptional repression [57,58]. Aberrant promoter hypermethylation and concomitant gene silencing have been observed at several genes in PCa, such as GSTP1, RASSF1, APC, CCND2, and PITX2 (Figure 2). Their 5-mC levels, both as an individual biomarker or as a biomarker panel, were demonstrated to discriminate PCa from benign tissue with high sensitivity and specificity [59,60,61,62,63,64,65]. The combined analysis of GSTP1, RASSF1, and APC showed superiority relative to the histological assessment of biopsy specimens, detecting PCa in 62 and 68% of histologically negative biopsies in two independent, large-scale studies (498 and 350 biopsies analyzed, respectively) [60,62]. These findings highlight the applicability of 5-mC biomarkers to large patient cohorts and demonstrate their capability in detecting locally confined tumors. Abnormal DNA methylation is described as emerging progressively in PCa and may be involved in the formation of metastases [66,67]. Hence, various studies aimed to identify 5-mC biomarkers with prognostic significances that could identify aggressive tumors. GSTP1, APC, RARB, and PITX2 are among the most extensively studied prognostic 5-mC biomarkers in Pca, and numerous reports emphasize their value as prognostic markers when analyzed in resected prostate tissues [61,64,65,68,69,70,71,72]. The promoter hypermethylation of those genes has been linked to the increased risk of clinical and biochemical recurrence (BCR) [68,69,70,71], as well as PCa-associated mortality [61,64,72]. Importantly, aberrant DNA hypermethylation events are also detectable in prostate tissue biopsies [63,73,74,75]. This permits risk stratifications prior to clinical interventions and may circumvent the overtreatment of indolent PCa. For instance, combined 5-mC intensities at GSTP1, APC, and RASSF1 in 12-core prostate biopsy specimens increased significantly across Gleason and NCCN (National Comprehensive Cancer Network) risk categories, demonstrating an association between 5-mC signals and the established risk factors. Interestingly, this association was still evident when only cancer-negative biopsy cores were assessed [74]. Hence, 5-mC biomarkers might compensate biopsy sampling limitations and facilitate the identification of high-grade tumors in histologically occult biopsies. Further studies found significant associations between unfavorable outcome and APC hypermethylation, as well as increased PITX2 methylation in high- versus low-risk prostate tumors in 6- and 12-core biopsies, respectively [63,73].
With technological advances and plummeting sequencing costs, genome-wide 5-mC profiling approaches gained popularity for the identification of prognostic biomarkers in PCa. In a population-based cohort of PCa patients following prostatectomy, methylation array analyses identified 42 CpG sites that stratified non-recurrent PCa from patients developing metastatic-lethal disease within 5 years. Eight of forty-two CpG biomarkers were subsequently validated in an independent cohort and improved the accuracy of predicting adverse outcome compared to the mere assessment of the Gleason score [76]. Another study used resected multifocal PCa and categorized the aggressiveness of individual foci based on 5-mC similarities relative to matched lymph node metastases. An aggressiveness classifier, integrating 5-mC signals at 25 CpG sites, separated aggressive (i.e., predicted origin of lymph metastasis) from indolent PCa foci. Classifier validation in the PCa samples of The Cancer Genome Atlas (TCGA) showed significant associations between the predicted aggressiveness and presence of lymph node metastases, as well as tumor stage [77].
These studies demonstrate that the integrative analysis of multiple 5-mC patterns exposes information about a patient´s tumor (e.g., cellular composition or 5-mC signatures of aggressiveness), which is missed by the evaluation of individual markers and may advance 5-mC-based PCa prognosis.

2.1.3. Histone Modification

In eukaryotic cell nuclei, histones are abundant DNA-wrapped structures and the major protein components of chromatin. Histones play an essential role in the regulation of gene expression via a number of post-translational modifications (i.e., acetylation and methylation). The deregulation of tri-methylated histone H3 at lysine 27 (H3K27me3) is known to silence tumor suppressor genes GAS2 and ADRB2, which are both involved in PCa progression [78]. Therefore, H3K27me3 alteration emerged as a new biomarker for aggressive PCa. H3K4me2 methylation contributes to the activation of a ubiquitin-conjugating enzyme complex, UBE2C, and can be used as a biomarker for PCa prognosis [79]. On the other hand, transcription factors, including SOX2, NR2F1, and NANOG, induce the hypomethylation and subsequently inactivation of H3 [80]. The acetylation, methylation, and ubiquitination of the histone-variant, H2A.Z, occur at active promoter sites and contribute to oncogene activation and, subsequently, the progression of PCa [81].

2.1.4. RNA

The transcriptome represents a layer of the molecular landscape of cancer cells that is useful in the development of clinically relevant biomarkers. In a study that evaluated the landscape of fusion genes in 144 localized PCa, the abundance of the read-through fusion transcripts RP11-356O9.1-TTC6 and TBCEL-TECTA was associated with BCR. Tumors harboring both fusion transcripts had a significantly increased risk for relapse [82]. The transcriptome profiling of five PCa specimens discovered eight fusion transcripts enriched in tumors compared to normal prostatic tissue [83]. These read-through transcripts were as follows: TRMT11-GRIK2, SLC45A2-AMACR, MTOR-TP53BP1, LRRC59-FLJ60017, TMEM135-CCDC67, KDM4-AC011523.2, MAN2A1-FER, and CCNH-C5orf30. In a subsequent validation experiment on PCa with clinical outcome data after radical prostatectomy (RP), the increased expression of any of the eight fusion transcripts was associated with BCR. No fusion transcripts were detected in healthy tissue donors, and positive expressions of TRMT11-GRIK2, MTOR-TP53BP1, and LRRC59-FLJ60017 were observed exclusively in recurrent tumors.
Long non-coding RNAs (lncRNAs) are transcripts that are at least 200 nt in length, with minimal or completely without protein-coding potentials [84]. The earliest utilization of lncRNAs for clinical use in PCa began with the discovery of the prostate-specific PCA3 transcript, which promotes cell survival by activating the AR pathway [85]. Unlike PSA, PCA3’s expression remains unaffected in other prostate pathologies, such as chronic prostatitis and benign prostatic hyperplasia (BPH) [86]. The RNA-seq analysis revealed that the lncRNA SChLAP1 is differentially expressed between aggressive and indolent tumors, and significantly associated with clinical risk factors for aggressive diseases, including BCR and metastatic progression [87,88,89]. Other lncRNAs with prognostic potential in PCa are PCAT-1 and PCAT-14. PCAT-1 was highly expressed in high-grade localized tumors and in a subset of metastatic specimens [90]. The low expression of PCAT-14 distinguished malignant from benign tissues, as well as high- and low-grade local tumors [90,91].
MicroRNAs (miRNAs) are small (~22 nt) non-coding RNAs involved in eukaryotic post-transcriptional gene regulation [92]. In localized PCa, differentially expressed miRNAs in tissue have been evaluated as risk stratification biomarkers, based on BCR-free survival and overall survival as common ends [93]. In a systematic review of 215 publications, 120 unique miRNAs were identified as individual prognostic markers in localized PCa [93]. Seven miRNAs—let-7b-5p, miR128a-3p, miR-188-5p, miR-224-5p, miR-23a-3p, miR-23b-3p, and miR-34b/c—were significantly associated with disease progression. Aside from individual miRNAs, miRNA signatures have also been evaluated for prognostic utility in localized PCa. A metric based on the tissue abundance of miR-96-5p, miR-183-5p, miR-145-5p, and miR221-5p was able to distinguish aggressive tumors from non-aggressive Pca in a Swedish cohort [94]. A 3-miRNA prognostic classifier involving miR-185-5p, miR-221-3p, and miR-326 predicted BCR-free survival in independent cohorts [95]. A highly combined RNA score based on the tumor expression of the miR-17-92 cluster (miR-17-3p, miR-17-5p, miR-18a, miR-19a, miR-19b, and miR-92a) was associated with a shorter BCR interval [96]. A 4-miRNA ratio model based on miR-23a-3p, miR-10b-5p, miR-133a, and miR-374b-5p tissue expression was shown to be a significant predictor of BCR-free survival, independent of routine clinicopathologic variables, as well as PCa-specific survival [97].
Circular RNAs (circRNAs) are predominantly noncoding RNAs generated via alterative splicing events that form covalently closed loops [98,99]. A comprehensive characterization of the circRNA landscape in localized PCa has shown significant circRNA dysregulation in tumors [98]. The same study identified 171 circRNAs essential for PCa. Additionally, tumors found to harbor extreme deviations (i.e., abundance or depletion) in circRNA expression were associated with poor BCR-free survival. In a recent study, the biomarker utility of circRNAs in PCa was further investigated using multiple patient cohorts for validation [99]. Ultimately, five circRNAs—circZNF532, circCDYL2, circLPAR3, circELK4, and circMAN1A2—were verified as deregulated between high/low pT stages. Moreover, the expressions of circSLC45A4, circFAT3, and circSEMA3C were independently validated to be negatively correlated with the Gleason score. Finally, circMKLN1 repression was successfully validated to be associated with BCR.

2.2. Gene Expression and Proteomic Biomarkers

Gene expression signatures can provide potential information for PCa prognosis. Importantly, AR acts as a transcription factor that regulates the expression of genes encoding cellular homeostasis and proteases that are important for prostate function. Proteases are excellent biomarkers in PCa as about half of the human proteases are expressed in the prostate gland [100]. KLK3, encoding PSA, is one of the members of the kallikrein-related peptidase family, and it is highly expressed in the prostate [101]. In addition to its widespread role in PCa diagnosis, KLK3 is involved in the progression of tumors towards metastasis by supporting cell proliferation and angiogenesis [102]. In addition to KLK3, the high-level expressions of CD133 and GRHL2 were observed in PCa and increased during disease progression [103]. Studies suggested that the expression levels of another cell surface antigen, prostate-specific membrane antigen (PSMA), directly reflect the tumor stage and BCR [104]. The upregulation of TMPRSS4 is known to be associated with PCa early invasion and metastasis. The expression of four other transcription factors, including SOX2, PROM1, SNAI2, and TWIST1, assists TMPRSS4 functions in early metastasis [105]. Using gene expression profiling tools by comparing normal and PCa samples, DMPK, KCNQ4, and WIF1 were shown to be downregulated [106]. In the same study, it was reported that the expression levels of KCNQ4, WIF1, PLN, and F3 were associated with shorter survival.
Data derived from targeted antibody-independent proteomic assays (PRISM-SRM assays) and mass spectrometry of PCa tissue proteome provide additional support to identify protein signatures for the risk stratification of PCa. In tissue, a few of the differentially expressed protein biomarkers are capable of predicting PCa progression. Comparing metastatic to non-metastatic PCa tissues, three differentially expressed proteins, including PARP, RDH11, and NDRG3, were shown to be involved in PCa progression [107]. A classifier was created using differentially expressed proteins in formalin-fixed parafilm-embedded tissue specimens. This classifier comprised a panel of five proteins (i.e., FOLH1, KLK3, TGFB1, SPARC, and CAMKK2) that are capable of stratifying PCa patients according to the risk of aggressiveness [108]. In a recent study, seven differentially expressed proteins (i.e., HSPA9 and HSPE1, NPM1, VCAN, SERBP1, MRPL3, and UQCRH) were successfully used as indicators of PCa progression and aggressiveness [109,110]. Using gene enrichment and immunohistochemical analyses, POSTN, CALR, and CTSD were found to be associated with PCa progression and short patient survival [111].
Furthermore, protein phosphorylation plays a critical role in cellular signaling pathways and is one of the multiple mechanisms that drives cancer progression. Alterations in the PCa phosphoproteome have been shown to shift metabolic programming and fuel cellular proliferation and metastasis [112]. AKT-mediated phosphorylation events involving leukemia inhibitory factor receptor (LIFR) and its downstream target, ERK2, activate a signaling cascade that promotes tumor growth and metastasis [113]. Therefore, the AKT/LIFR/ERK2 axis can be used as a potential biomarker for predicting the risk of PCa progression.

3. Biomarkers in the Tumor Microenvironment

The tumor microenvironment (TME) is a network consisting of supporting cellular and stromal components: (i) immune cells, including tumor-associated macrophages, and (ii) non-immune cells, such as endothelial cells, inflammatory cells, and cancer associated-fibroblasts (CAFs). Cancer cells remodel the stromal cells and transform the TME into a habitat capable of evading immune cell infiltration to enhance tumor development. The TME releases various factors into blood circulation, including chemokines, cytokines, and enzymes, that can be used as biomarkers for detecting PCa in early stages.
Through accelerating the epithelial-to-mesenchymal transition (EMT) of cancer cells, stromal CAFs enhance the migration, proliferation, and invasiveness of PCa cells. Another regulator of the EMT pathway is transcription factor FOXA1. Mutation and post-translational modifications, such as methylation and acetylation, can alter FOXA1 activity and promote cancer progression (Figure 1) [114]. Studies have shown that miRNAs derived from extracellular vesicles (EVs) play an important role in tumor–stromal interactions [115]. miR-409 is located within the DLK1-DIO3 cluster and released from EVs. The upregulation of miR-409 was found to be associated with higher Gleason scores and PCa aggressiveness by targeting tumor suppressor genes (i.e., Ras suppressor 1, tumor suppressor candidate 1, and stromal antigen 2) [116]. Other DLK1-DIO3 cluster miRNAs (i.e., miR-154* and miR-379) have similar roles and actively take part in tumor-stroma interactions by inhibiting SMAD7 and stromal antigen 2, respectively [117]. SMAD7 is an antagonist of transforming growth factor-beta (TGF-β). TGF-β and growth and differentiation factor 15 (GDF15) play a crucial role in the production of collagen and the deposition of extracellular matrices (ECMs). Stromal cells are sources of TGF15, which can contribute to cancer progression driven by the TME [118,119].
Depending on the tumor stage and genetics, tumoral autophagy in the TME plays a divergent role (i.e., the suppression or promotion) in PCa development. mTORC is an important regulatory kinase involved in cell proliferation and autophagy. Genetic alterations, such as PTEN deletions, lead to the constitutive activation of mTORC, decrease autophagy, and encourage tumor growth and metastasis [120,121]. In addition, the expression levels of four major autophagy genes (LC3A, LC3B, p62, and Beclin1) are associated with higher Gleason scores and tumor aggressiveness [120,122]. The detection of these autophagy biomarkers may be of value for the risk stratification of PCa.
AR and its ligands, testosterone and 5α-dihydrotestosterone, promote cellular signaling events that are essential for prostate growth and optimal function. Stromal cells express high levels of AR during prostate growth. In androgen ablation therapies, a low AR expression during the course of PCa progression is associated with BCR and disease progression [123]. In this scenario, PCa cells use other receptor transcription factors, such as the Glucocorticoid Receptor (GR), independently of AR. The upregulation of GR signaling promotes cell proliferation and differentiation and, therefore, has been associated with aggressive PCa [124].

4. Liquid Biopsy Biomarkers in Prostate Cancer

4.1. Cell-Free DNA

Liquid biopsies emerged as a powerful tool for the early identification of aggressive PCa. Cell-free DNA (cfDNA) isolated from bodily fluids (e.g., blood, urine, saliva, or cerebrospinal fluid) is well known to contain tumor-derived molecular information and has been used for cancer diagnosis, genotyping, prognosis, and disease monitoring in various tumor entities [125,126,127,128]. In contrast to tissue samples, liquid biopsies represent less procedural risk to the patient and are not spatially restricted. A cfDNA sample may contain tumor DNA from multiple (if not all) cancerous lesions and could overcome sampling limitations posed by conventional tissue biopsies [129,130]. Hence, biomarkers for aggressive Pca could be obtained repeatedly and with a lower risk of sampling bias and undersampling. Data derived from a study using 45 plasma samples from Pca patients showed that the circulating tumor DNA (ctDNA)-based analysis in liquid biopsy is capable of identifying all driver DNA modifications present in matched solid biopsy [56]. Promising results were obtained in advanced or late-stage Pca, but the detection of early stage Pca using liquid biopsies is still limited due to the low quantity of ctDNA [131,132].

4.1.1. Total cfDNA Amount and Integrity

The analysis of the total cfDNA content can serve as a first diagnostic tool prior to the analysis of more specific genomic (i.e., mutations and CNAs) and epigenomic (i.e., methylated DNA and histone modifications) alterations. The total amount of detectable cfDNA and, particularly, ctDNA arising from tumor cells can vary widely between different tumor entities. Several studies indicate that the higher plasma cfDNA concentrations in Pca patients are correlated to a positive biopsy, high Gleason score, PSA recurrence, and shorter BCR-free survival [133,134].
In addition to cfDNA quantification, a more detailed characterization of the cfDNA fragments can be applied to further improve the sensitivity of tumor detection. In plasma, cfDNA fragment profiles harbor a characteristic peak at 167 bp. It has been shown that cfDNA fragment lengths are slightly shorter in cancer patients, outlining the potential of fragmentation analyses for tumor diagnostics [135,136,137,138]. CfDNA integrity, determined by the quantification of the ALU gene (i.e., ALU 247 bp ratio to ALU 115 bp), is high in PCa patients and can distinguish PCa patients from BPH patients with elevated PSA [139,140]. Additionally, the ratio of the PTGS2 DNA fragment (small fragment size < 200 bp, associated with apoptosis) to the Reprimo DNA fragment (larger fragment size > 250 bp, associated with other types of cell death) was used to define an apoptosis index [141]. This index distinguishes between PCa patients with localized disease and BPH patients and shows a significant correlation to BCR. Interestingly, seminal fluids from PCa patients contain significantly longer cfDNA fragments (>1000bp) compared to BPH patients and healthy controls [142].

4.1.2. Genomic Analysis of cfDNA

The WGS analysis of cfDNA revealed multiple mutations and CNAs consistent with PCa tissue analyses: BRCA1/2, TP53, SPOP, ATM, APC, and CHEK2 mutations, as well as CNAs in TMPRSS2-ERG and RB1, and PTEN deletions, AR and MYC amplification (Figure 2) [55,56,143,144,145,146,147,148,149,150]. AR and its constitutively active splice variants, AR-V7 and AR-V3, are associated with poor survival and are used as potential biomarkers in PCa [151,152]. In addition to AR, TP53 mutations and TMPRSS2-ERG fusion are among the most frequently altered genes in PCa (Figure 1). Since BRCA1/2 are involved in DDR machinery, existing mutations in BRCA1/2 increase further somatic mutation rates, as well as CNAs, and they are considered high-risk biomarkers for PCa progression [153].
The loss of heterozygosity (LOH) in specific chromosomal regions detected by short tandem repeats (STR; microsatellite DNA) can be used to distinguish PCa and BPH patients at the time of the initial diagnosis [154,155]. D8S360 is a STR marker that discriminates between metastatic and localized disease, as metastatic PCa carries a higher frequency of LOH in D8S360 [154,156]. It has been reported that the LOH frequency at 14 microsatellite markers was linked to risk factors, such as tumor stage, Gleason score, and metastatic disease [156].
Applying tagged-amplicon deep sequencing on the plasma samples of patients with localized PCa, TP53 variants were identified in 22 of 189 cases (12%), and 14 out of 21 detectable TP53 variants were predicted to be pathogenic or likely pathogenic [157]. Patients with the detectable TP53 ctDNA variant harbored a significantly shorter metastasis-free survival [157,158]. Both studies highlight the potential of ctDNA detection for early diagnosis or risk stratification in PCa.

4.1.3. cfDNA Methylation

Various reports demonstrated that hypermethylation events at genes with prognostic implications in PCa tissue (e.g., GSTP1, APC, and PTGS2) are also detectable in plasma, serum, or urine samples [159,160,161,162,163,164,165,166]. Different liquid biopsy-based 5-mC biomarker panels showed positive correlations between 5-mC levels and the increasing Gleason score [159,165,166], suggesting the possibility of minimal invasive risk stratification. For instance, a urinary 4-gene 5-mC panel (i.e., GSTP1, APC, CRIP3, and HOXD8) was found to predict disease progression in patients with Gleason 6 PCa under active surveillance [166]. These results illustrate that aberrant 5-mC signatures in urine may be of value for identifying men that could benefit from initiating active therapy. Another study associated 5-mC levels at six genes (i.e., GSTP1, SFRP2, IGFBP3, IGFBP7, APC, and PTGS2) with increased Gleason scores, as well as clinical risk categories, and demonstrated that urine 5-mC markers can detect high-grade lesions missed by tissue biopsies (Figure 2) [165]. Despite these promising results, the detectability of tumor-derived 5-mC signals in liquid biopsies remains limited, especially in localized PCa [160,161]. This might be overcome by expanding the number of assessed biomarkers using genome-wide approaches. However, only a few studies have employed genome-scale methylome profiling relative to cfDNA samples so far due to the low DNA quantities present in most plasma and urine samples [167,168,169]. Although these studies focus mostly on advanced PCa, their results are encouraging for applications in localized disease. As an example, Moss and colleagues showed that prostate-specific 5-mC signals can be detected in plasma cfDNA via a tissue-of-origin deconvolution approach and linked the prostate-specific 5-mC levels to therapy resistance/sensitivity [168]. This suggests that cellular compositions with potential prognostic values can be inferred from PCa cfDNA samples [170]. Currently, whole-methylome-profiling techniques are adapted to low DNA input quantities (e.g., cfMeDIP-seq [171,172], cfMBD-seq [173,174], cfTAPS [175], and cfMethyl-seq [176]), allowing their usage on cfDNA samples. For example, a recent finding revealed hypermethylation in metastasis, compared to localized PCa using cfMeDIP-seq technology, and provided valuable insights into the 5-mC-dependent mechanisms of tumor aggressiveness [177].

4.2. Cell-Free RNA

An analysis of cell-free RNA (cfRNA) molecules enables the minimally invasive detection of PCa in biofluids including blood, urine, and seminal fluids. Urine-based biomarker tests using RNA analytes are similarly being applied for risk estimation. The ExoDX Prostate IntelliScore (Exosome Diagnostics) assesses tumor grades based on the abundance of the PCA3, ERG, and SPDEF RNA from a first catch, non-DRE urine specimen [178].
In a validation cohort of 519 patients, the 3-gene (i.e., SPDEF, ERG, and PCA3) assay discriminated high-grade tumors (Gleason score ≥ 7) from low-grade and benign tumors [179]. The Mi-Prostate Score integrates TMPRSS2-ERG and PCA3 RNA levels in DRE urine samples with serum PSA abundance. The incorporation of urine TMPRSS2-ERG and PCA3 abundance, independently or in combination, outperformed the risk prediction model for localized diseases based on serum PSA alone [180]. Another disease risk score, utilizing combined HOXC6 and DLX1 mRNA levels in post-DRE urine and traditional clinical risk factors, was successful in detecting high-grade and clinically significant (Gleason score ≥ 7) PCa [181].
The combined circulating transcript levels of GOLM1, NKX3-1, and TRPM8 were shown to be able to stratify low- and high-risk PCa [182]. Aggressive tumors were defined to present extracapsular extension with advanced tumor stage (≥pT3). The 3-mRNA markers outperformed both the Gleason score and PSA in identifying high-risk PCa.
In a study involving a cohort of 100 PCa patients and 50 men with BPH, 14 miRNAs were consistently abundant in the serum samples from patients with low-grade PCa or BPH, while high-grade PCa had significantly reduced levels [183]. A score based on the serum abundance of all 14 miRNAs was predictive of the absence of high-grade PCa in the cohort. Moreover, combining the miRNA signature with BCR data resulted in the accurate classification of low-risk PCa. A 3-miRNA signature (miR-222-3p, miR-24-3p, and miR-30c-5p) detected in urine was identified in cohort 1, and it was validated in cohort 2 that it could distinguish BPH samples from PCa specimens. An additional 3-miRNA signature (miR-125b-5p, let-7a-5p, and miR-151-5p) was simultaneously identified in cohort 2—and validated in cohort 1—which can predict time to BCR independent of clinicopathologic features after RP [184]. The plasma levels of miR-20a and miR-21 were found to be associated with a high Cancer of the Prostate Risk Assessment score [185]. Moreover, the combination of four miRNAs (i.e., miR-20a, miR-21, miR-145, and miR-221) was able to classify low-and high-risk disease, which were initially assessed using the D’Amico risk classification score [186]. In another study, a high-risk disease was identified based on the Gleason score, pathological T stage, surgical margin status, and diagnostic PSA. The plasma levels of miR-17-5p and miR-106a-5p, as well as miR-20a-5p and miR-20b-5p, were shown to be associated with high-risk diseases [187]. These results were also consistent with the miRNA expression and BCR-free survival data from the TCGA dataset.

4.3. Extracellular Vesicles

Nucleic acids and proteins isolated from EVs are potentially useful biomarkers for the diagnosis, detection of tumor origin, and risk stratification of PCa [188]. Four PCa-related miRNA (miR-141, miR-145, miR-221, and miR-451a) and the known PCa mRNA biomarkers, PCA3 and TMPRSS2-ERG, could be quantified in plasma or urinary exosomes from PCa patients [189,190]. Further studies identified GATA-binding protein 2 (GATA2) transcripts in urinary exosomes as additional biomarkers for the detection and risk stratification of PCa. GATA2 mRNA is secreted by PCa cells and known to play a role in PCa development and aggressiveness. The combination of all three biomarkers further improved PCa detection compared to a single biomarker evaluation [191].
In a comprehensive proteomic analysis of EVs from urine and PCa tissue, the majority of proteins identified in EVs originating from PCa tissue were also detectable in urinary EVs (94.69%), indicating that the proteomic analysis of urinary EVs reflects the proteome of EVs in PCa tissues. The comparison of exosomal content in urine samples from PCa patients before and after the initial treatment showed that prostate-specific proteins (e.g. KLK2, KLK3/PSA, FOLH1, MSMB, ACPP, TGM4, NDRG1, and NKX3-1) and androgen-regulated genes (e.g., FKBP5, FAM129A, RAB27A, FASN, and NEFH) were significantly reduced in post-treatment urine samples, which were in turn significantly enriched in EVs containing bladder- and kidney-associated proteins [192]. In a proteomic analysis of urinary exosomes, eleven proteins (SCIN, AMBP, FABP5, CHMP4C, CHMP2B, BAIAP2, GRN, SYTL2, CALR, CHMP4A, and DNPH1) showed significant enrichments in the urine samples of PCa patients versus patients with negative biopsy [193]. Fatty-acid-binding protein 5 showed significant higher levels in PCa samples compared to patients with negative biopsy, and it was significantly associated with the Gleason score for the prediction of PCa.

4.4. Circulating Tumor Cells and Secretome

Circulating tumor cells (CTCs) are secreted from primary and/or metastatic tumor sites; they enter the circulation and are responsible for tumor metastasis. The presence of CTCs in the peripheral blood of cancer patients is associated with decreased patient survival [194]. Due to low numbers in non-metastatic tumors, CTCs are detected and counted using enrichment methods [195]. Carrying the majority of mutations present in PCa tissues, CTCs can provide useful prognostic information for PCa detection at early stages.
The single cell isolation and subsequently amplification of the whole genome, immunostaining, next generation sequencing (NGS), and fluorescence in situ hybridization analyses have been used to characterize genomic aberrations of CTCs. The assessment of a DNA-based signature in CTCs revealed genomic alterations, such as a loss in tumor suppressor genes TP53, PTEN, NKX3-1, CDKN1B, and RB1. Losses in TP53, PTEN, and RB1 were linked to PCa aggressiveness [28,196]. Common chromosomal gains in PCa include the amplification of genes coding the focal adhesion kinase PTK2, NCOA2, and MYC [196,197]. CNAs in NCOA2 offer prognostic values for the identification of aggressive PCa [49]. Additionally, AR and EXT1 are the other affected genes with gains in the chromosomal regions. CTCs in aggressive PCa possess large-scale transition scores, as well as DNA-SCARS as genomic alterations [196].
The proteomic analysis of CTCs showed a number of inflammatory cytokines, proteases, and glycoproteins that can potentially be used as biomarkers. An elevated level of IL-8 has been observed in cell lines with invasive behavior, compared to those less likely to progress [198]. Matrix metallopeptidases (MMPs) are a group of zinc-dependent endopeptidases known as facilitators of tumor invasion and metastasis via degrading connective tissue barriers [55,199]. Analyzing CTC-secreted MMPs revealed higher levels of MMP-2 and MMP-9 expression in aggressive compared to indolent PCa [199,200]. A quantitative mass spectrometry-based approach showed that the urinary glycoprotein, prostatic acid phosphatase (ACPP), has discrimination power for distinguishing aggressive PCa from non-aggressive PCa [201]. The urinary levels of 8-OHdG and 8-iso-PGF2α involved in oxidative DNA damage are also shown to be associated with PCa’s progression [202]. Similarly to tissues, the elevated PSMA expression in CTCs was correlated with tumor grade and PCa progression [104,203]. This is equally the case for CD133, as its overexpression was indicative of progressive late-stage PCa [204]. A systematic review of 43 publications, representing CTCs analyses impact on non-metastatic PCa, reported the utility of CTCs as potential prognostic markers in localized PCa. Nonetheless, validation studies are required to stablish the sensitivity and specificity of CTCs for the characterization of early stage PCa [205].

4.5. Tumor-Educated Platelets

Tumor-Educated Platelets (TEPs) reprogram the behavior of CTCs via a direct interaction accompanied by exchanging RNA, lipids, and proteins (i.e., beta-3 integrin marker CD61) [206]. This trafficking event creates a supportive microenvironment for CTCs and facilitating their immune evasion via protecting them from natural killer cells [207]. The ability of platelets to alter tumor cells’ phenotypes and genetic content promote cell growth and plasticity [206]. During the course of tumor metastases initiation, TEPs facilitate EMT to lead the tumor towards metastases [206]. Hence, TEPs biomarkers can present adjuvant information for the early diagnosis of PCa. For instance, the detections of TEP biomarker transcripts (i.e., KLK2, KLK3, and FOLH1) were shown to be associated with high-level serum PSA and poor survival [208].

5. Importance of Bioinformatic Tools in Prostate Cancer Detection and Risk Stratification

The development of high-throughput technologies, such as NGS, markedly increased our knowledge towards exploring cancer biomarkers, and subsequently facilitated the detection and risk assessment of tumors. Despite having a promising future in development of precision medicine, these technologies generate a massive volume of multiple data streams. Bioinformatic pipelines offer computationally efficient and reliable prospects for analyzing and studying the sequencing data, allowing the development of a sensitive and specific system for PCa management operations.
In addition, PCa intra-tumoral heterogeneity and its following challenges in early tumor detections can be relatively overcome by bioinformatic-assisted workflows. For instance, projects such as TRACERx focus on cancer evolutionary dynamics (i.e., the relationship between prostate tumor heterogeneity and tumor stage) via the characterization of chromosome instability [209,210]. DriverSub, PyClone, and SciClone are examples of novel bioinformatic approaches for the identification of tumor-derived mutations [210,211]. Furthermore, the multiple gene expression profiles of public datasets via the application of bioinformatic analysis identified seven genes (i.e., BCO1, BAIAP2L2, C7, AP000844.2, ASB9, MKI67P1, and TMEM272) involved in the PCa’s early prognosis and risk stratification [212].

6. Multi-Parametric Methods for Identifying Aggressive Prostate Cancer

A major challenge in the early diagnosis of PCa using liquid biopsies is the low-abundance of tumor-derived genetic signatures (i.e., cfDNA and cfRNA). Overcoming this requires high-sequencing coverages to retrieve cancer-informative signals [136]. In addition, only a minor fraction of ctDNA exhibits genomic alterations, particularly at the early stage of cancer development. Therefore, studies sought to integrate different diagnostic/prognostic techniques to establish cancer-informative parameters and to improve the detection of tumor-specific signals. In this regard, there are two possibilities to interpret the data derived from PCa examinations: (i) combining different diagnostic and prognostic approaches (i.e., tissue biopsy, blood, and urine from liquid biopsy, imaging, and histopathology), and (ii) integrating different liquid biopsy profiles (i.e., genomic, transcriptomic, and epigenomic). In support of this, artificial intelligence (AI) can enhance the diagnostic expertise of clinicians for PCa risk stratifications (Figure 3). Further prioritization of biomarker candidates using established data repositories, such as the Cancer Genome Census (CGC), could also be performed before undertaking further validation studies [213].

6.1. Integration of Different Diagnostic/Prognostic Approaches

Although data integration from different diagnostic and prognostic approaches is relatively new, a few attempts have been made to combine genomic signatures and imaging techniques to improve PCa risk assessments. The combination of genomic analysis (i.e., CNAs), multiparametric MRI (mpMRI), and 68Ga-PSMA-PET/CT imaging resulted in a strong correlation between imaging features and genomic index lesions [214]. Using public genomic data, bioinformatic tools can compare mpMRI-visible and mpMRI-invisible lesions at a genomic level to underpin the genomic basis of PCa [215].
On the other hand, liquid biopsy data and prostate imaging enable the integration of multiple data streams into powerful multidisciplinary applications that can improve PCa detection. Recently, this approach advanced beyond pilot studies with a new study integrating CT-based radiomic analysis and liquid biopsy signatures from patients with advanced PCa. Accordingly, CTC counts, plasma cfDNA levels, and specific genomic alterations have shown strong correlations with the radiomic analysis of CT scans [216]. In addition, the implementation of AI and sequencing data derived from both tissue and liquid biopsy resulted in the creation of a sensitive and specific system that non-invasively facilitates PCa management (Figure 3) [217,218]. This system can add prognostic value for the differentiation of indolent and aggressive PCa.

6.2. Multi-Parametric Approach within Liquid Biopsy Profiles

Information derived from multiple liquid biopsy profiles (i.e., genomic, epigenomic, transcriptomic, and fragmentomic) has the potential to introduce additional prominent biomarkers for the detection and risk assessment of PCa. A new study applied genomic, epigenomic, and fragmentation metrics in liquid biopsies to advance the use of liquid biopsy in pediatric cancer [219]. In addition, such multi-omics approaches can facilitate the characterization of heterogenous tumors [220]. For instance, cell-type compositions within tumors are estimated by an in silico method, epigenomic deconvolution (EDec), that combines DNA methylation and gene transcription profiles [221].
Other studies either combined the characterization of CTCs (enumeration and/or RNA analysis) with other analytes (i.e., cfDNA and ctDNA) [222,223,224,225] or performed the profiling of cfDNA and cfRNA to assess genomic alterations, along with transcriptomic aberrations [226]. For instance, an integrative multianalyte approach combined urinary methylation markers from men with a clinical suspicion for PCa with cfRNA quantification from EVs and clinical parameters. Applying a machine learning algorithm, this multivariable prediction model was capable of predicting the presence of PCa [227]. A multimodal genomic and epigenomic approach determined the molecular profile of low-abundance ctDNA via combining somatic mutation, fragmentation, and methylation analyses on cfDNA of patients with head and neck squamous cell carcinomas. Interrogating multiple methods confirmed an association between detectable ctDNA and the risk of having advanced diseases, as well as worse overall survival [228]. As a next step, future studies can employ independent multiparametric approaches to create a comprehensive model that has the potential to non-invasively improve early diagnosis, risk assessment, treatment monitoring, and clinical outcomes for aggressive tumors.

7. Conclusions

Despite its limitations, tissue biopsy remains the preferred tool for the detection and risk assessment of prostate tumors. The multitude of liquid biopsy biomarkers can play a complementary role for the non-invasive detection of PCa and can provide more insights into the differentiation of indolent and aggressive tumors. The identification of high-risk genomic, epigenomic, transcriptomic, and proteomic alterations, combined with orthogonal clinical parameters, can potentially improve the diagnosis of aggressive tumors and can increase patient survival. Integrating tissue, liquid biopsy, and other diagnostic/prognostic modalities, such as imaging, along with the clinical pathological profiles and novel machine learning approaches, could provide the necessary information for optimal patient management.

Author Contributions

Conceptualization, S.E., A.L.R. and M.G.; investigation, S.E., A.L.R., A.K.A., F.J. and M.G.; visualization, S.E.; writing and editing, S.E., A.L.R., A.K.A., F.J., S.D., H.S. and M.G.; supervision, S.D., H.S. and M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was realized through support by the Dieter Morszeck Stiftung.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  2. Gandaglia, G.; Leni, R.; Bray, F.; Fleshner, N.; Freedland, S.J.; Kibel, A.; Stattin, P.; Van Poppel, H.; La Vecchia, C. Epidemiology and Prevention of Prostate Cancer. Eur. Urol. Oncol. 2021, 4, 877–892. [Google Scholar] [CrossRef] [PubMed]
  3. Rebello, R.J.; Oing, C.; Knudsen, K.E.; Loeb, S.; Johnson, D.C.; Reiter, R.E.; Gillessen, S.; Van der Kwast, T.; Bristow, R.G. Prostate cancer. Nat. Rev. Dis. Primers 2021, 7, 9. [Google Scholar] [CrossRef] [PubMed]
  4. Ng, K.; Smith, S.; Shamash, J. Metastatic Hormone-Sensitive Prostate Cancer (mHSPC): Advances and Treatment Strategies in the First-Line Setting. Oncol. Ther. 2020, 8, 209–230. [Google Scholar] [CrossRef] [PubMed]
  5. Sartor, O.; de Bono, J.S. Metastatic Prostate Cancer. N. Engl. J. Med. 2018, 378, 645–657. [Google Scholar] [CrossRef]
  6. Van Poppel, H.; Albreht, T.; Basu, P.; Hogenhout, R.; Collen, S.; Roobol, M. Serum PSA-based early detection of prostate cancer in Europe and globally: Past, present and future. Nat. Rev. Urol. 2022, 19, 562–572. [Google Scholar] [CrossRef]
  7. Schröder, F.H.; Hugosson, J.; Carlsson, S.; Tammela, T.; Määttänen, L.; Auvinen, A.; Kwiatkowski, M.; Recker, F.; Roobol, M.J. Screening for prostate cancer decreases the risk of developing metastatic disease: Findings from the European Randomized Study of Screening for Prostate Cancer (ERSPC). Eur. Urol. 2012, 62, 745–752. [Google Scholar] [CrossRef]
  8. Schröder, F.H.; Hugosson, J.; Roobol, M.J.; Tammela, T.L.; Ciatto, S.; Nelen, V.; Kwiatkowski, M.; Lujan, M.; Lilja, H.; Zappa, M.; et al. Prostate-cancer mortality at 11 years of follow-up. N. Engl. J. Med. 2012, 366, 981–990. [Google Scholar] [CrossRef]
  9. Pierre-Victor, D.; Parnes, H.L.; Andriole, G.L.; Pinsky, P.F. Prostate Cancer Incidence and Mortality Following a Negative Biopsy in a Population Undergoing PSA Screening. Urology 2021, 155, 62–69. [Google Scholar] [CrossRef]
  10. Saini, S. PSA and beyond: Alternative prostate cancer biomarkers. Cell. Oncol. 2016, 39, 97–106. [Google Scholar] [CrossRef]
  11. Loeb, S.; Bjurlin, M.A.; Nicholson, J.; Tammela, T.L.; Penson, D.F.; Carter, H.B.; Carroll, P.; Etzioni, R. Overdiagnosis and overtreatment of prostate cancer. Eur. Urol. 2014, 65, 1046–1055. [Google Scholar] [CrossRef] [Green Version]
  12. Liu, J.; Wang, Z.Q.; Li, M.; Zhou, M.Y.; Yu, Y.F.; Zhan, W.W. Establishment of two new predictive models for prostate cancer to determine whether to require prostate biopsy when the PSA level is in the diagnostic gray zone (4–10 ng mL−1). Asian J. Androl. 2020, 22, 213–216. [Google Scholar] [CrossRef] [PubMed]
  13. Clinton, T.N.; Bagrodia, A.; Lotan, Y.; Margulis, V.; Raj, G.V.; Woldu, S.L. Tissue-based biomarkers in prostate cancer. Expert Rev. Precis. Med. Drug Dev. 2017, 2, 249–260. [Google Scholar] [CrossRef] [Green Version]
  14. Massie, C.E.; Mills, I.G.; Lynch, A.G. The importance of DNA methylation in prostate cancer development. J. Steroid Biochem. Mol. Biol. 2017, 166, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. 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]
  16. Van Neste, L.; Herman, J.G.; Otto, G.; Bigley, J.W.; Epstein, J.I.; Van Criekinge, W. The epigenetic promise for prostate cancer diagnosis. Prostate 2012, 72, 1248–1261. [Google Scholar] [CrossRef]
  17. 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] [Green Version]
  18. Haffner, M.C.; Zwart, W.; Roudier, M.P.; True, L.D.; Nelson, W.G.; Epstein, J.I.; De Marzo, A.M.; Nelson, P.S.; Yegnasubramanian, S. Genomic and phenotypic heterogeneity in prostate cancer. Nat. Rev. Urol. 2021, 18, 79–92. [Google Scholar] [CrossRef]
  19. 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] [Green Version]
  20. Attard, G.; Parker, C.; Eeles, R.A.; Schröder, F.; Tomlins, S.A.; Tannock, I.; Drake, C.G.; de Bono, J.S. Prostate cancer. Lancet 2016, 387, 70–82. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. Oh, M.; Alkhushaym, N.; Fallatah, S.; Althagafi, A.; Aljadeed, R.; Alsowaida, Y.; Jeter, J.; Martin, J.R.; Babiker, H.M.; McBride, A.; et al. The association of BRCA1 and BRCA2 mutations with prostate cancer risk, frequency, and mortality: A meta-analysis. Prostate 2019, 79, 880–895. [Google Scholar] [CrossRef] [PubMed]
  23. Mitra, A.; Fisher, C.; Foster, C.S.; Jameson, C.; Barbachanno, Y.; Bartlett, J.; Bancroft, E.; Doherty, R.; Kote-Jarai, Z.; Peock, S.; et al. Prostate cancer in male BRCA1 and BRCA2 mutation carriers has a more aggressive phenotype. Br. J. Cancer 2008, 98, 502–507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Dall’Era, M.A.; McPherson, J.D.; Gao, A.C.; DeVere White, R.W.; Gregg, J.P.; Lara, P.N., Jr. Germline and somatic DNA repair gene alterations in prostate cancer. Cancer 2020, 126, 2980–2985. [Google Scholar] [CrossRef]
  25. Cancer Genome Atlas Research, N. The Molecular Taxonomy of Primary Prostate Cancer. Cell 2015, 163, 1011–1025. [Google Scholar] [CrossRef] [Green Version]
  26. Chakraborty, G.; Armenia, J.; Mazzu, Y.Z.; Nandakumar, S.; Stopsack, K.H.; Atiq, M.O.; Komura, K.; Jehane, L.; Hirani, R.; Chadalavada, K.; et al. Significance of BRCA2 and RB1 Co-loss in Aggressive Prostate Cancer Progression. Clin. Cancer Res. 2020, 26, 2047–2064. [Google Scholar] [CrossRef]
  27. Beltran, H.; Yelensky, R.; Frampton, G.M.; Park, K.; Downing, S.R.; MacDonald, T.Y.; Jarosz, M.; Lipson, D.; Tagawa, S.T.; Nanus, D.M.; et al. Targeted next-generation sequencing of advanced prostate cancer identifies potential therapeutic targets and disease heterogeneity. Eur. Urol. 2013, 63, 920–926. [Google Scholar] [CrossRef] [Green Version]
  28. 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]
  29. Arora, K.; Barbieri, C.E. Molecular Subtypes of Prostate Cancer. Curr. Oncol. Rep. 2018, 20, 58. [Google Scholar] [CrossRef]
  30. 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]
  31. Teroerde, M.; Nientiedt, C.; Duensing, A.; Hohenfellner, M.; Stenzinger, A.; Duensing, S. Revisiting the Role of p53 in Prostate Cancer. In Prostate Cancer; Bott, S.R.J., Ng, K.L., Eds.; Exon Publications: Brisbane, Australia, 2021. [Google Scholar] [CrossRef]
  32. Nientiedt, C.; Budczies, J.; Endris, V.; Kirchner, M.; Schwab, C.; Jurcic, C.; Behnisch, R.; Hoveida, S.; Lantwin, P.; Kaczorowski, A.; et al. Mutations in TP53 or DNA damage repair genes define poor prognostic subgroups in primary prostate cancer. Urol. Oncol. 2022, 40, 8.e11–18.e18. [Google Scholar] [CrossRef]
  33. Aggarwal, R.R.; Quigley, D.A.; Huang, J.; Zhang, L.; Beer, T.M.; Rettig, M.B.; Reiter, R.E.; Gleave, M.E.; Thomas, G.V.; Foye, A.; et al. Whole-Genome and Transcriptional Analysis of Treatment-Emergent Small-Cell Neuroendocrine Prostate Cancer Demonstrates Intraclass Heterogeneity. Mol. Cancer Res. 2019, 17, 1235–1240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. 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] [Green Version]
  35. Mangolini, A.; Rocca, C.; Bassi, C.; Ippolito, C.; Negrini, M.; Dell’Atti, L.; Lanza, G.; Gafà, R.; Bianchi, N.; Pinton, P.; et al. Detection of disease-causing mutations in prostate cancer by NGS sequencing. Cell Biol. Int. 2022, 46, 1047–1061. [Google Scholar] [CrossRef]
  36. Taylor, B.S.; Schultz, N.; Hieronymus, H.; Gopalan, A.; Xiao, Y.; Carver, B.S.; Arora, V.K.; Kaushik, P.; Cerami, E.; Reva, B.; et al. Integrative genomic profiling of human prostate cancer. Cancer Cell 2010, 18, 11–22. [Google Scholar] [CrossRef] [Green Version]
  37. Quigley, D.A.; Dang, H.X.; Zhao, S.G.; Lloyd, P.; Aggarwal, R.; Alumkal, J.J.; Foye, A.; Kothari, V.; Perry, M.D.; Bailey, A.M.; et al. Genomic Hallmarks and Structural Variation in Metastatic Prostate Cancer. Cell 2018, 174, 758–769.e759. [Google Scholar] [CrossRef] [Green Version]
  38. 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]
  39. Cotter, K.; Rubin, M.A. The evolving landscape of prostate cancer somatic mutations. Prostate 2022, 82 (Suppl. 1), S13–S24. [Google Scholar] [CrossRef]
  40. Sooreshjani, M.A.; Nikhil, K.; Kamra, M.; Nguyen, D.N.; Kumar, D.; Shah, K. LIMK2-NKX3.1 Engagement Promotes Castration-Resistant Prostate Cancer. Cancers 2021, 13, 2324. [Google Scholar] [CrossRef]
  41. Asatiani, E.; Huang, W.X.; Wang, A.; Rodriguez Ortner, E.; Cavalli, L.R.; Haddad, B.R.; Gelmann, E.P. Deletion, methylation, and expression of the NKX3.1 suppressor gene in primary human prostate cancer. Cancer Res. 2005, 65, 1164–1173. [Google Scholar] [CrossRef] [Green Version]
  42. Dong, H.Y.; Ding, L.; Zhou, T.R.; Yan, T.; Li, J.; Liang, C. FOXA1 in prostate cancer. Asian J. Androl. 2022. [Google Scholar] [CrossRef]
  43. Wang, Z.; Wang, Y.; Zhang, J.; Hu, Q.; Zhi, F.; Zhang, S.; Mao, D.; Zhang, Y.; Liang, H. Significance of the TMPRSS2:ERG gene fusion in prostate cancer. Mol. Med. Rep. 2017, 16, 5450–5458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Burdova, A.; Bouchal, J.; Tavandzis, S.; Kolar, Z. TMPRSS2-ERG gene fusion in prostate cancer. Biomed. Pap. Med. Fac. Univ. Palacky Olomouc Czech Repub. 2014, 158, 502–510. [Google Scholar] [CrossRef] [Green Version]
  45. Rosen, P.; Sesterhenn, I.A.; Brassell, S.A.; McLeod, D.G.; Srivastava, S.; Dobi, A. Clinical potential of the ERG oncoprotein in prostate cancer. Nat. Rev. Urol. 2012, 9, 131–137. [Google Scholar] [CrossRef] [PubMed]
  46. Shah, R.B.; Bentley, J.; Jeffery, Z.; DeMarzo, A.M. Heterogeneity of PTEN and ERG expression in prostate cancer on core needle biopsies: Implications for cancer risk stratification and biomarker sampling. Hum. Pathol. 2015, 46, 698–706. [Google Scholar] [CrossRef]
  47. Liu, W.; Xie, C.C.; Thomas, C.Y.; Kim, S.T.; Lindberg, J.; Egevad, L.; Wang, Z.; Zhang, Z.; Sun, J.; Sun, J.; et al. Genetic markers associated with early cancer-specific mortality following prostatectomy. Cancer 2013, 119, 2405–2412. [Google Scholar] [CrossRef] [Green Version]
  48. Tapia-Laliena, M.A.; Korzeniewski, N.; Hohenfellner, M.; Duensing, S. High-risk prostate cancer: A disease of genomic instability. Urol. Oncol. 2014, 32, 1101–1107. [Google Scholar] [CrossRef]
  49. Silva, M.P.; Barros-Silva, J.D.; Vieira, J.; Lisboa, S.; Torres, L.; Correia, C.; Vieira-Coimbra, M.; Martins, A.T.; Jerónimo, C.; Henrique, R.; et al. NCOA2 is a candidate target gene of 8q gain associated with clinically aggressive prostate cancer. Genes Chromosomes Cancer 2016, 55, 365–374. [Google Scholar] [CrossRef]
  50. Faisal, F.A.; Murali, S.; Kaur, H.; Vidotto, T.; Guedes, L.B.; Salles, D.C.; Kothari, V.; Tosoian, J.J.; Han, S.; Hovelson, D.H.; et al. CDKN1B Deletions are Associated with Metastasis in African American Men with Clinically Localized, Surgically Treated Prostate Cancer. Clin. Cancer Res. 2020, 26, 2595–2602. [Google Scholar] [CrossRef]
  51. Chang, B.L.; Zheng, S.L.; Isaacs, S.D.; Wiley, K.E.; Turner, A.; Li, G.; Walsh, P.C.; Meyers, D.A.; Isaacs, W.B.; Xu, J. A polymorphism in the CDKN1B gene is associated with increased risk of hereditary prostate cancer. Cancer Res. 2004, 64, 1997–1999. [Google Scholar] [CrossRef] [Green Version]
  52. Kim, M.S.; Je, E.M.; Yoo, N.J.; Lee, S.H. Loss of ARID1A expression is uncommon in gastric, colorectal, and prostate cancers. Apmis 2012, 120, 1020–1022. [Google Scholar] [CrossRef] [PubMed]
  53. Santric, V.; Djokic, M.; Suvakov, S.; Pljesa-Ercegovac, M.; Nikitovic, M.; Radic, T.; Acimovic, M.; Stankovic, V.; Bumbasirevic, U.; Milojevic, B.; et al. GSTP1 rs1138272 Polymorphism Affects Prostate Cancer Risk. Medicina 2020, 56, 128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Hopkins, J.F.; Sabelnykova, V.Y.; Weischenfeldt, J.; Simon, R.; Aguiar, J.A.; Alkallas, R.; Heisler, L.E.; Zhang, J.; Watson, J.D.; Chua, M.L.K.; et al. Mitochondrial mutations drive prostate cancer aggression. Nat. Commun. 2017, 8, 656. [Google Scholar] [CrossRef] [PubMed]
  55. Dathathri, E.; Isebia, K.T.; Abali, F.; Lolkema, M.P.; Martens, J.W.M.; Terstappen, L.; Bansal, R. Liquid Biopsy Based Circulating Biomarkers in Metastatic Prostate Cancer. Front. Oncol. 2022, 12, 863472. [Google Scholar] [CrossRef]
  56. Wyatt, A.W.; Annala, M.; Aggarwal, R.; Beja, K.; Feng, F.; Youngren, J.; Foye, A.; Lloyd, P.; Nykter, M.; Beer, T.M.; et al. Concordance of Circulating Tumor DNA and Matched Metastatic Tissue Biopsy in Prostate Cancer. J. Natl. Cancer Inst. 2017, 109. [Google Scholar] [CrossRef] [Green Version]
  57. Dor, Y.; Cedar, H. Principles of DNA methylation and their implications for biology and medicine. Lancet 2018, 392, 777–786. [Google Scholar] [CrossRef]
  58. Jones, P.A.; Baylin, S.B. The epigenomics of cancer. Cell 2007, 128, 683–692. [Google Scholar] [CrossRef] [Green Version]
  59. Fiano, V.; Zugna, D.; Grasso, C.; Trevisan, M.; Delsedime, L.; Molinaro, L.; Cassoni, P.; Papotti, M.; Merletti, F.; Akre, O.; et al. DNA methylation in repeat negative prostate biopsies as a marker of missed prostate cancer. Clin. Epigenetics 2019, 11, 152. [Google Scholar] [CrossRef]
  60. Partin, A.W.; Van Neste, L.; Klein, E.A.; Marks, L.S.; Gee, J.R.; Troyer, D.A.; Rieger-Christ, K.; Jones, J.S.; Magi-Galluzzi, C.; Mangold, L.A.; et al. Clinical validation of an epigenetic assay to predict negative histopathological results in repeat prostate biopsies. J. Urol. 2014, 192, 1081–1087. [Google Scholar] [CrossRef] [Green Version]
  61. Rosenbaum, E.; Hoque, M.O.; Cohen, Y.; Zahurak, M.; Eisenberger, M.A.; Epstein, J.I.; Partin, A.W.; Sidransky, D. Promoter Hypermethylation as an Independent Prognostic Factor for Relapse in Patients with Prostate Cancer Following Radical Prostatectomy. Clin. Cancer Res. 2005, 11, 8321–8325. [Google Scholar] [CrossRef] [Green Version]
  62. Stewart, G.D.; Van Neste, L.; Delvenne, P.; Delrée, P.; Delga, A.; McNeill, S.A.; O’Donnell, M.; Clark, J.; Van Criekinge, W.; Bigley, J.; et al. Clinical utility of an epigenetic assay to detect occult prostate cancer in histopathologically negative biopsies: Results of the MATLOC study. J. Urol. 2013, 189, 1110–1116. [Google Scholar] [CrossRef] [PubMed]
  63. Uhl, B.; Gevensleben, H.; Tolkach, Y.; Sailer, V.; Majores, M.; Jung, M.; Meller, S.; Stein, J.; Ellinger, J.; Dietrich, D.; et al. PITX2 DNA Methylation as Biomarker for Individualized Risk Assessment of Prostate Cancer in Core Biopsies. J. Mol. Diagn. 2017, 19, 107–114. [Google Scholar] [PubMed] [Green Version]
  64. Vasiljević, N.; Ahmad, A.S.; Carter, P.D.; Fisher, G.; Berney, D.M.; Foster, C.S.; Cuzick, J.; Lorincz, A.T. DNA methylation of PITX2 predicts poor survival in men with prostate cancer. Biomark. Med. 2014, 8, 1143–1150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Weiss, G.; Cottrell, S.; Distler, J.; Schatz, P.; Kristiansen, G.; Ittmann, M.; Haefliger, C.; Lesche, R.; Hartmann, A.; Corman, J.; et al. DNA Methylation of the PITX2 Gene Promoter Region is a Strong Independent Prognostic Marker of Biochemical Recurrence in Patients With Prostate Cancer After Radical Prostatectomy. J. Urol. 2009, 181, 1678–1685. [Google Scholar] [CrossRef]
  66. Majumdar, S.; Buckles, E.; Estrada, J.; Koochekpour, S. Aberrant DNA methylation and prostate cancer. Curr. Genom. 2011, 12, 486–505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. McMahon, K.W.; Karunasena, E.; Ahuja, N. The Roles of DNA Methylation in the Stages of Cancer. Cancer J. 2017, 23, 257–261. [Google Scholar] [CrossRef] [PubMed]
  68. Bañez, L.L.; Sun, L.; van Leenders, G.J.; Wheeler, T.M.; Bangma, C.H.; Freedland, S.J.; Ittmann, M.M.; Lark, A.L.; Madden, J.F.; Hartman, A.; et al. Multicenter clinical validation of PITX2 methylation as a prostate specific antigen recurrence predictor in patients with post-radical prostatectomy prostate cancer. J. Urol. 2010, 184, 149–156. [Google Scholar] [CrossRef]
  69. Daniunaite, K.; Jarmalaite, S.; Kalinauskaite, N.; Petroska, D.; Laurinavicius, A.; Lazutka, J.R.; Jankevicius, F. Prognostic value of RASSF1 promoter methylation in prostate cancer. J. Urol. 2014, 192, 1849–1855. [Google Scholar] [CrossRef]
  70. Dietrich, D.; Hasinger, O.; Bañez, L.L.; Sun, L.; van Leenders, G.J.; Wheeler, T.M.; Bangma, C.H.; Wernert, N.; Perner, S.; Freedland, S.J.; et al. Development and clinical validation of a real-time PCR assay for PITX2 DNA methylation to predict prostate-specific antigen recurrence in prostate cancer patients following radical prostatectomy. J. Mol. Diagn. 2013, 15, 270–279. [Google Scholar] [CrossRef]
  71. Maldonado, L.; Brait, M.; Loyo, M.; Sullenberger, L.; Wang, K.; Peskoe, S.B.; Rosenbaum, E.; Howard, R.; Toubaji, A.; Albadine, R.; et al. GSTP1 promoter methylation is associated with recurrence in early stage prostate cancer. J. Urol. 2014, 192, 1542–1548. [Google Scholar] [CrossRef] [Green Version]
  72. Richiardi, L.; Fiano, V.; Vizzini, L.; De Marco, L.; Delsedime, L.; Akre, O.; Tos, A.G.; Merletti, F. Promoter methylation in APC, RUNX3, and GSTP1 and mortality in prostate cancer patients. J. Clin. Oncol. 2009, 27, 3161–3168. [Google Scholar] [CrossRef] [PubMed]
  73. Henrique, R.; Ribeiro, F.R.; Fonseca, D.; Hoque, M.O.; Carvalho, A.L.; Costa, V.L.; Pinto, M.; Oliveira, J.; Teixeira, M.R.; Sidransky, D.; et al. High promoter methylation levels of APC predict poor prognosis in sextant biopsies from prostate cancer patients. Clin. Cancer Res. 2007, 13, 6122–6129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Van Neste, L.; Groskopf, J.; Grizzle, W.E.; Adams, G.W.; DeGuenther, M.S.; Kolettis, P.N.; Bryant, J.E.; Kearney, G.P.; Kearney, M.C.; Van Criekinge, W.; et al. Epigenetic risk score improves prostate cancer risk assessment. Prostate 2017, 77, 1259–1264. [Google Scholar] [CrossRef] [PubMed]
  75. Van Neste, L.; Partin, A.W.; Stewart, G.D.; Epstein, J.I.; Harrison, D.J.; Van Criekinge, W. Risk score predicts high-grade prostate cancer in DNA-methylation positive, histopathologically negative biopsies. Prostate 2016, 76, 1078–1087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Zhao, S.; Geybels, M.S.; Leonardson, A.; Rubicz, R.; Kolb, S.; Yan, Q.; Klotzle, B.; Bibikova, M.; Hurtado-Coll, A.; Troyer, D.; et al. Epigenome-Wide Tumor DNA Methylation Profiling Identifies Novel Prognostic Biomarkers of Metastatic-Lethal Progression in Men Diagnosed with Clinically Localized Prostate Cancer. Clin. Cancer Res. 2017, 23, 311–319. [Google Scholar] [CrossRef] [Green Version]
  77. Mundbjerg, K.; Chopra, S.; Alemozaffar, M.; Duymich, C.; Lakshminarasimhan, R.; Nichols, P.W.; Aron, M.; Siegmund, K.D.; Ukimura, O.; Aron, M.; et al. Identifying aggressive prostate cancer foci using a DNA methylation classifier. Genome Biol. 2017, 18, 3. [Google Scholar] [CrossRef] [Green Version]
  78. Chen, Z.; Wang, L.; Wang, Q.; Li, W. Histone modifications and chromatin organization in prostate cancer. Epigenomics 2010, 2, 551–560. [Google Scholar] [CrossRef] [Green Version]
  79. Pang, J.; Yang, Y.W.; Huang, Y.; Yang, J.; Zhang, H.; Chen, R.; Dong, L.; Huang, Y.; Wang, D.; Liu, J.; et al. P110β Inhibition Reduces Histone H3K4 Di-Methylation in Prostate Cancer. Prostate 2017, 77, 299–308. [Google Scholar] [CrossRef]
  80. Cackowski, F.C.; Heath, E.I. Prostate cancer dormancy and recurrence. Cancer Lett. 2022, 524, 103–108. [Google Scholar] [CrossRef]
  81. Dryhurst, D.; Ausió, J. Histone H2A.Z deregulation in prostate cancer. Cause or effect? Cancer Metastasis Rev. 2014, 33, 429–439. [Google Scholar] [CrossRef] [Green Version]
  82. Chen, S.; Huang, V.; Xu, X.; Livingstone, J.; Soares, F.; Jeon, J.; Zeng, Y.; Hua, J.T.; Petricca, J.; Guo, H.; et al. Widespread and Functional RNA Circularization in Localized Prostate Cancer. Cell 2019, 176, 831–843.e22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Yu, Y.P.; Ding, Y.; Chen, Z.; Liu, S.; Michalopoulos, A.; Chen, R.; Gulzar, Z.G.; Yang, B.; Cieply, K.M.; Luvison, A.; et al. Novel fusion transcripts associate with progressive prostate cancer. Am. J. Pathol. 2014, 184, 2840–2849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ulitsky, I.; Bartel, D.P. lincRNAs: Genomics, evolution, and mechanisms. Cell 2013, 154, 26–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ferreira, L.B.; Palumbo, A.; de Mello, K.D.; Sternberg, C.; Caetano, M.S.; de Oliveira, F.L.; Neves, A.F.; Nasciutti, L.E.; Goulart, L.R.; Gimba, E.R. PCA3 noncoding RNA is involved in the control of prostate-cancer cell survival and modulates androgen receptor signaling. BMC Cancer 2012, 12, 507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Deras, I.L.; Aubin, S.M.; Blase, A.; Day, J.R.; Koo, S.; Partin, A.W.; Ellis, W.J.; Marks, L.S.; Fradet, Y.; Rittenhouse, H.; et al. PCA3: A molecular urine assay for predicting prostate biopsy outcome. J. Urol. 2008, 179, 1587–1592. [Google Scholar] [CrossRef]
  87. Prensner, J.R.; Zhao, S.; Erho, N.; Schipper, M.; Iyer, M.K.; Dhanasekaran, S.M.; Magi-Galluzzi, C.; Mehra, R.; Sahu, A.; Siddiqui, J.; et al. RNA biomarkers associated with metastatic progression in prostate cancer: A multi-institutional high-throughput analysis of SChLAP1. Lancet Oncol. 2014, 15, 1469–1480. [Google Scholar] [CrossRef] [Green Version]
  88. Mehra, R.; Shi, Y.; Udager, A.M.; Prensner, J.R.; Sahu, A.; Iyer, M.K.; Siddiqui, J.; Cao, X.; Wei, J.; Jiang, H.; et al. A novel RNA in situ hybridization assay for the long noncoding RNA SChLAP1 predicts poor clinical outcome after radical prostatectomy in clinically localized prostate cancer. Neoplasia 2014, 16, 1121–1127. [Google Scholar] [CrossRef] [Green Version]
  89. Cesnik, A.J.; Yang, B.; Truong, A.; Etheridge, T.; Spiniello, M.; Steinbrink, M.I.; Shortreed, M.R.; Frey, B.L.; Jarrard, D.F.; Smith, L.M. Long Noncoding RNAs AC009014.3 and Newly Discovered XPLAID Differentiate Aggressive and Indolent Prostate Cancers. Transl. Oncol. 2018, 11, 808–814. [Google Scholar] [CrossRef]
  90. Prensner, J.R.; Iyer, M.K.; Balbin, O.A.; Dhanasekaran, S.M.; Cao, Q.; Brenner, J.C.; Laxman, B.; Asangani, I.A.; Grasso, C.S.; Kominsky, H.D.; et al. Transcriptome sequencing across a prostate cancer cohort identifies PCAT-1, an unannotated lincRNA implicated in disease progression. Nat. Biotechnol. 2011, 29, 742–749. [Google Scholar] [CrossRef]
  91. Shukla, S.; Zhang, X.; Niknafs, Y.S.; Xiao, L.; Mehra, R.; Cieślik, M.; Ross, A.; Schaeffer, E.; Malik, B.; Guo, S.; et al. Identification and Validation of PCAT14 as Prognostic Biomarker in Prostate Cancer. Neoplasia 2016, 18, 489–499. [Google Scholar] [CrossRef] [Green Version]
  92. Bartel, D.P. Metazoan MicroRNAs. Cell 2018, 173, 20–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Rana, S.; Valbuena, G.N.; Curry, E.; Bevan, C.L.; Keun, H.C. MicroRNAs as biomarkers for prostate cancer prognosis: A systematic review and a systematic reanalysis of public data. Br. J. Cancer 2022, 126, 502–513. [Google Scholar] [CrossRef] [PubMed]
  94. Larne, O.; Martens-Uzunova, E.; Hagman, Z.; Edsjö, A.; Lippolis, G.; den Berg, M.S.; Bjartell, A.; Jenster, G.; Ceder, Y. miQ--a novel microRNA based diagnostic and prognostic tool for prostate cancer. Int. J. Cancer 2013, 132, 2867–2875. [Google Scholar] [CrossRef]
  95. Kristensen, H.; Thomsen, A.R.; Haldrup, C.; Dyrskjøt, L.; Høyer, S.; Borre, M.; Mouritzen, P.; Ørntoft, T.F.; Sørensen, K.D. Novel diagnostic and prognostic classifiers for prostate cancer identified by genome-wide microRNA profiling. Oncotarget 2016, 7, 30760–30771. [Google Scholar] [CrossRef]
  96. Feng, S.; Qian, X.; Li, H.; Zhang, X. Combinations of elevated tissue miRNA-17-92 cluster expression and serum prostate-specific antigen as potential diagnostic biomarkers for prostate cancer. Oncol. Lett. 2017, 14, 6943–6949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Schmidt, L.; Fredsøe, J.; Kristensen, H.; Strand, S.H.; Rasmussen, A.; Høyer, S.; Borre, M.; Mouritzen, P.; Ørntoft, T.; Sørensen, K.D. Training and validation of a novel 4-miRNA ratio model (MiCaP) for prediction of postoperative outcome in prostate cancer patients. Ann. Oncol. 2018, 29, 2003–2009. [Google Scholar] [CrossRef] [PubMed]
  98. Chen, K.; Zhao, H.; Shi, Y.; Yang, F.; Wang, L.T.; Kang, G.; Nie, Y.; Wang, J. Perioperative Dynamic Changes in Circulating Tumor DNA in Patients with Lung Cancer (DYNAMIC). Clin. Cancer Res. 2019, 25, 7058–7067. [Google Scholar] [CrossRef] [Green Version]
  99. Hansen, E.B.; Fredsøe, J.; Okholm, T.L.H.; Ulhøi, B.P.; Klingenberg, S.; Jensen, J.B.; Kjems, J.; Bouchelouche, K.; Borre, M.; Damgaard, C.K.; et al. The transcriptional landscape and biomarker potential of circular RNAs in prostate cancer. Genome Med. 2022, 14, 8. [Google Scholar] [CrossRef]
  100. Kappelhoff, R.; Puente, X.S.; Wilson, C.H.; Seth, A.; López-Otín, C.; Overall, C.M. Overview of transcriptomic analysis of all human proteases, non-proteolytic homologs and inhibitors: Organ, tissue and ovarian cancer cell line expression profiling of the human protease degradome by the CLIP-CHIP™ DNA microarray. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 2210–2219. [Google Scholar] [CrossRef]
  101. Koistinen, H.; Künnapuu, J.; Jeltsch, M. KLK3 in the Regulation of Angiogenesis-Tumorigenic or Not? Int. J. Mol. Sci. 2021, 22, 13545. [Google Scholar] [CrossRef]
  102. Fuhrman-Luck, R.A.; Loessner, D.; Clements, J.A. Kallikrein-Related Peptidases in Prostate Cancer: From Molecular Function to Clinical Application. Ejifcc 2014, 25, 269–281. [Google Scholar] [PubMed]
  103. Wu, J.; Guo, S.; Wang, L.; Liao, Z. Correlation analysis between CD133, Klk3 and grhl2 expression and tumor characteristics in prostate cancer. Cell. Mol. Biol. 2022, 67, 68–73. [Google Scholar] [CrossRef] [PubMed]
  104. Ristau, B.T.; O’Keefe, D.S.; Bacich, D.J. The prostate-specific membrane antigen: Lessons and current clinical implications from 20 years of research. Urol. Oncol. 2014, 32, 272–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lee, Y.; Yoon, J.; Ko, D.; Yu, M.; Lee, S.; Kim, S. TMPRSS4 promotes cancer stem–like properties in prostate cancer cells through upregulation of SOX2 by SLUG and TWIST1. J. Exp. Clin. Cancer Res. 2021, 40, 372. [Google Scholar] [CrossRef]
  106. Santos, N.J.; Camargo, A.C.L.; Carvalho, H.F.; Justulin, L.A.; Felisbino, S.L. Prostate Cancer Secretome and Membrane Proteome from Pten Conditional Knockout Mice Identify Potential Biomarkers for Disease Progression. Int. J. Mol. Sci. 2022, 23, 9224. [Google Scholar] [CrossRef]
  107. Muller, A.K.; Foll, M.; Heckelmann, B.; Kiefer, S.; Werner, M.; Schilling, O.; Biniossek, M.L.; Jilg, C.A.; Drendel, V. Proteomic Characterization of Prostate Cancer to Distinguish Nonmetastasizing and Metastasizing Primary Tumors and Lymph Node Metastases. Neoplasia 2018, 20, 140–151. [Google Scholar] [CrossRef]
  108. Gao, Y.; Wang, Y.T.; Chen, Y.; Wang, H.; Young, D.; Shi, T.; Song, Y.; Schepmoes, A.A.; Kuo, C.; Fillmore, T.L.; et al. Proteomic Tissue-Based Classifier for Early Prediction of Prostate Cancer Progression. Cancers 2020, 12, 1268. [Google Scholar] [CrossRef]
  109. Lygirou, V.; Fasoulakis, K.; Stroggilos, R.; Makridakis, M.; Latosinska, A.; Frantzi, M.; Katafigiotis, I.; Alamanis, C.; Stravodimos, K.G.; Constantinides, C.A.; et al. Proteomic Analysis of Prostate Cancer FFPE Samples Reveals Markers of Disease Progression and Aggressiveness. Cancers 2022, 14, 3765. [Google Scholar] [CrossRef]
  110. Shipitsin, M.; Small, C.; Choudhury, S.; Giladi, E.; Friedlander, S.; Nardone, J.; Hussain, S.; Hurley, A.D.; Ernst, C.; Huang, Y.E.; et al. Identification of proteomic biomarkers predicting prostate cancer aggressiveness and lethality despite biopsy-sampling error. Br. J. Cancer 2014, 111, 1201–1212. [Google Scholar] [CrossRef]
  111. Garcia-Marques, F.; Liu, S.; Totten, S.M.; Bermudez, A.; Tanimoto, C.; Hsu, E.C.; Nolley, R.; Hembree, A.; Stoyanova, T.; Brooks, J.D.; et al. Protein signatures to distinguish aggressive from indolent prostate cancer. Prostate 2022, 82, 605–616. [Google Scholar] [CrossRef]
  112. Ahmad, F.; Cherukuri, M.K.; Choyke, P.L. Metabolic reprogramming in prostate cancer. Br. J. Cancer 2021, 125, 1185–1196. [Google Scholar] [CrossRef] [PubMed]
  113. Shao, J.; Zhu, W.; Ding, Y.; Zhu, H.; Jing, X.; Yu, H.; Lu, M.; Qiao, Y.; Wang, X.; Ai, H.; et al. Phosphorylation of LIFR promotes prostate cancer progression by activating the AKT pathway. Cancer Lett. 2019, 451, 110–121. [Google Scholar] [CrossRef] [PubMed]
  114. Teng, M.; Zhou, S.; Cai, C.; Lupien, M.; He, H.H. Pioneer of prostate cancer: Past, present and the future of FOXA1. Protein Cell 2021, 12, 29–38. [Google Scholar] [CrossRef]
  115. Josson, S.; Chung, L.W.; Gururajan, M. microRNAs and Prostate Cancer. Adv. Exp. Med. Biol. 2015, 889, 105–118. [Google Scholar] [CrossRef] [Green Version]
  116. Josson, S.; Gururajan, M.; Sung, S.Y.; Hu, P.; Shao, C.; Zhau, H.E.; Liu, C.; Lichterman, J.; Duan, P.; Li, Q.; et al. Stromal fibroblast-derived miR-409 promotes epithelial-to-mesenchymal transition and prostate tumorigenesis. Oncogene 2015, 34, 2690–2699. [Google Scholar] [CrossRef]
  117. Gururajan, M.; Josson, S.; Chu, G.C.; Lu, C.L.; Lu, Y.T.; Haga, C.L.; Zhau, H.E.; Liu, C.; Lichterman, J.; Duan, P.; et al. miR-154* and miR-379 in the DLK1-DIO3 microRNA mega-cluster regulate epithelial to mesenchymal transition and bone metastasis of prostate cancer. Clin. Cancer Res. 2014, 20, 6559–6569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Bruzzese, F.; Hägglöf, C.; Leone, A.; Sjöberg, E.; Roca, M.S.; Kiflemariam, S.; Sjöblom, T.; Hammarsten, P.; Egevad, L.; Bergh, A.; et al. Local and systemic protumorigenic effects of cancer-associated fibroblast-derived GDF15. Cancer Res. 2014, 74, 3408–3417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Shiao, S.L.; Chu, G.C.; Chung, L.W. Regulation of prostate cancer progression by the tumor microenvironment. Cancer Lett. 2016, 380, 340–348. [Google Scholar] [CrossRef] [Green Version]
  120. Loizzo, D.; Pandolfo, S.D.; Rogers, D.; Cerrato, C.; di Meo, N.A.; Autorino, R.; Mirone, V.; Ferro, M.; Porta, C.; Stella, A.; et al. Novel Insights into Autophagy and Prostate Cancer: A Comprehensive Review. Int. J. Mol. Sci. 2022, 23, 3826. [Google Scholar] [CrossRef]
  121. Zou, Z.; Tao, T.; Li, H.; Zhu, X. mTOR signaling pathway and mTOR inhibitors in cancer: Progress and challenges. Cell Biosci. 2020, 10, 31. [Google Scholar] [CrossRef] [Green Version]
  122. Giatromanolaki, A.; Sivridis, E.; Mendrinos, S.; Koutsopoulos, A.V.; Koukourakis, M.I. Autophagy proteins in prostate cancer: Relation with anaerobic metabolism and Gleason score. Urol. Oncol. 2014, 32, 39.e11–39.e18. [Google Scholar] [CrossRef] [PubMed]
  123. Aurilio, G.; Cimadamore, A.; Mazzucchelli, R.; Lopez-Beltran, A.; Verri, E.; Scarpelli, M.; Massari, F.; Cheng, L.; Santoni, M.; Montironi, R. Androgen Receptor Signaling Pathway in Prostate Cancer: From Genetics to Clinical Applications. Cells 2020, 9, 2653. [Google Scholar] [PubMed]
  124. Hu, J.; Chen, Q. The role of glucocorticoid receptor in prostate cancer progression: From bench to bedside. Int. Urol. Nephrol. 2017, 49, 369–380. [Google Scholar] [CrossRef] [PubMed]
  125. Abbosh, C.; Birkbak, N.J.; Wilson, G.A.; Jamal-Hanjani, M.; Constantin, T.; Salari, R.; Le Quesne, J.; Moore, D.A.; Veeriah, S.; Rosenthal, R.; et al. Phylogenetic ctDNA analysis depicts early-stage lung cancer evolution. Nature 2017, 545, 446–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Cohen, J.D.; Li, L.; Wang, Y.; Thoburn, C.; Afsari, B.; Danilova, L.; Douville, C.; Javed, A.A.; Wong, F.; Mattox, A.; et al. Detection and localization of surgically resectable cancers with a multi-analyte blood test. Science 2018, 359, 926–930. [Google Scholar] [CrossRef] [Green Version]
  127. Kurtz, D.M.; Soo, J.; Co Ting Keh, L.; Alig, S.; Chabon, J.J.; Sworder, B.J.; Schultz, A.; Jin, M.C.; Scherer, F.; Garofalo, A.; et al. Enhanced detection of minimal residual disease by targeted sequencing of phased variants in circulating tumor DNA. Nat. Biotechnol. 2021, 39, 1537–1547. [Google Scholar] [CrossRef]
  128. Newman, A.M.; Bratman, S.V.; To, J.; Wynne, J.F.; Eclov, N.C.; Modlin, L.A.; Liu, C.L.; Neal, J.W.; Wakelee, H.A.; Merritt, R.E.; et al. An ultrasensitive method for quantitating circulating tumor DNA with broad patient coverage. Nat. Med. 2014, 20, 548–554. [Google Scholar] [CrossRef]
  129. Murtaza, M.; Dawson, S.-J.; Pogrebniak, K.; Rueda, O.M.; Provenzano, E.; Grant, J.; Chin, S.-F.; Tsui, D.W.Y.; Marass, F.; Gale, D.; et al. Multifocal clonal evolution characterized using circulating tumour DNA in a case of metastatic breast cancer. Nat. Commun. 2015, 6, 8760. [Google Scholar] [CrossRef] [Green Version]
  130. Wan, J.C.M.; Massie, C.; Garcia-Corbacho, J.; Mouliere, F.; Brenton, J.D.; Caldas, C.; Pacey, S.; Baird, R.; Rosenfeld, N. Liquid biopsies come of age: Towards implementation of circulating tumour DNA. Nat. Rev. Cancer 2017, 17, 223–238. [Google Scholar] [CrossRef]
  131. Campos-Fernández, E.; Barcelos, L.S.; de Souza, A.G.; Goulart, L.R.; Alonso-Goulart, V. Research landscape of liquid biopsies in prostate cancer. Am. J. Cancer Res. 2019, 9, 1309–1328. [Google Scholar]
  132. Wang, Y.; Wang, Z.; Gang, X.; Wang, G. Liquid biopsy in prostate cancer: Current status and future challenges of clinical application. Aging Male 2021, 24, 58–71. [Google Scholar] [CrossRef] [PubMed]
  133. Wroclawski, M.L.; Serpa-Neto, A.; Fonseca, F.L.; Castro-Neves-Neto, O.; Pompeo, A.S.; Machado, M.T.; Pompeo, A.C.; del Giglio, A. Cell-free plasma DNA as biochemical biomarker for the diagnosis and follow-up of prostate cancer patients. Tumor Biol. 2013, 34, 2921–2927. [Google Scholar] [CrossRef] [PubMed]
  134. Bastian, P.J.; Palapattu, G.S.; Yegnasubramanian, S.; Lin, X.; Rogers, C.G.; Mangold, L.A.; Trock, B.; Eisenberger, M.; Partin, A.W.; Nelson, W.G. Prognostic Value of Preoperative Serum Cell-Free Circulating DNA in Men with Prostate Cancer Undergoing Radical Prostatectomy. Clin. Cancer Res. 2007, 13, 5361–5367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Mouliere, F.; Smith, C.G.; Heider, K.; Su, J.; van der Pol, Y.; Thompson, M.; Morris, J.; Wan, J.C.M.; Chandrananda, D.; Hadfield, J.; et al. Fragmentation patterns and personalized sequencing of cell-free DNA in urine and plasma of glioma patients. EMBO Mol. Med. 2021, 13, e12881. [Google Scholar] [CrossRef]
  136. Van der Pol, Y.; Mouliere, F. Toward the Early Detection of Cancer by Decoding the Epigenetic and Environmental Fingerprints of Cell-Free DNA. Cancer Cell 2019, 36, 350–368. [Google Scholar] [CrossRef]
  137. Mouliere, F.; Chandrananda, D.; Piskorz, A.M.; Moore, E.K.; Morris, J.; Ahlborn, L.B.; Mair, R.; Goranova, T.; Marass, F.; Heider, K.; et al. Enhanced detection of circulating tumor DNA by fragment size analysis. Sci. Transl. Med. 2018, 10, eaat4921. [Google Scholar] [CrossRef] [Green Version]
  138. Underhill, H.R.; Kitzman, J.O.; Hellwig, S.; Welker, N.C.; Daza, R.; Baker, D.N.; Gligorich, K.M.; Rostomily, R.C.; Bronner, M.P.; Shendure, J. Fragment Length of Circulating Tumor DNA. PLoS Genet. 2016, 12, e1006162. [Google Scholar] [CrossRef] [Green Version]
  139. Feng, J.; Gang, F.; Li, X.; Jin, T.; Houbao, H.; Yu, C.; Guorong, L. Plasma cell-free DNA and its DNA integrity as biomarker to distinguish prostate cancer from benign prostatic hyperplasia in patients with increased serum prostate-specific antigen. Int. Urol. Nephrol. 2013, 45, 1023–1028. [Google Scholar] [CrossRef]
  140. Khani, M.; Hosseini, J.; Mirfakhraie, R.; Habibi, M.; Azargashb, E.; Pouresmaeili, F. The value of the plasma circulating cell-free DNA concentration and integrity index as a clinical tool for prostate cancer diagnosis: A prospective case-control cohort study in an Iranian population. Cancer Manag. Res. 2019, 11, 4549–4556. [Google Scholar] [CrossRef] [Green Version]
  141. Ellinger, J.; Bastian, P.J.; Haan, K.I.; Heukamp, L.C.; Buettner, R.; Fimmers, R.; Mueller, S.C.; von Ruecker, A. Noncancerous PTGS2 DNA fragments of apoptotic origin in sera of prostate cancer patients qualify as diagnostic and prognostic indicators. Int. J. Cancer 2008, 122, 138–143. [Google Scholar] [CrossRef]
  142. Ponti, G.; Maccaferri, M.; Micali, S.; Manfredini, M.; Milandri, R.; Bianchi, G.; Pellacani, G.; Kaleci, S.; Chester, J.; Conti, A.; et al. Seminal Cell Free DNA Concentration Levels Discriminate between Prostate Cancer and Benign Prostatic Hyperplasia. Anticancer Res. 2018, 38, 5121–5125. [Google Scholar] [CrossRef] [PubMed]
  143. Romanel, A.; Gasi Tandefelt, D.; Conteduca, V.; Jayaram, A.; Casiraghi, N.; Wetterskog, D.; Salvi, S.; Amadori, D.; Zafeiriou, Z.; Rescigno, P.; et al. Plasma AR and abiraterone-resistant prostate cancer. Sci. Transl. Med. 2015, 7, 312re310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Salvi, S.; Casadio, V.; Conteduca, V.; Burgio, S.L.; Menna, C.; Bianchi, E.; Rossi, L.; Carretta, E.; Masini, C.; Amadori, D.; et al. Circulating cell-free AR and CYP17A1 copy number variations may associate with outcome of metastatic castration-resistant prostate cancer patients treated with abiraterone. Br. J. Cancer 2015, 112, 1717–1724. [Google Scholar] [CrossRef] [Green Version]
  145. 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] [PubMed]
  146. Del Re, M.; Conteduca, V.; Crucitta, S.; Gurioli, G.; Casadei, C.; Restante, G.; Schepisi, G.; Lolli, C.; Cucchiara, F.; Danesi, R.; et al. Androgen receptor gain in circulating free DNA and splicing variant 7 in exosomes predict clinical outcome in CRPC patients treated with abiraterone and enzalutamide. Prostate Cancer Prostatic Dis. 2021, 24, 524–531. [Google Scholar] [CrossRef]
  147. Heitzer, E.; Ulz, P.; Belic, J.; Gutschi, S.; Quehenberger, F.; Fischereder, K.; Benezeder, T.; Auer, M.; Pischler, C.; Mannweiler, S.; et al. Tumor-associated copy number changes in the circulation of patients with prostate cancer identified through whole-genome sequencing. Genome Med. 2013, 5, 30. [Google Scholar] [CrossRef] [Green Version]
  148. 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]
  149. 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] [Green Version]
  150. Warner, E.; Herberts, C.; Fu, S.; Yip, S.; Wong, A.; Wang, G.; Ritch, E.; Murtha, A.J.; Vandekerkhove, G.; Fonseca, N.M.; et al. BRCA2, ATM, and CDK12 Defects Differentially Shape Prostate Tumor Driver Genomics and Clinical Aggression. Clin. Cancer Res. 2021, 27, 1650–1662. [Google Scholar] [CrossRef]
  151. Zhang, T.; Karsh, L.I.; Nissenblatt, M.J.; Canfield, S.E. Androgen Receptor Splice Variant, AR-V7, as a Biomarker of Resistance to Androgen Axis-Targeted Therapies in Advanced Prostate Cancer. Clin. Genitourin. Cancer 2020, 18, 1–10. [Google Scholar] [CrossRef] [Green Version]
  152. Duffy, M.J. Biomarkers for prostate cancer: Prostate-specific antigen and beyond. Clin. Chem. Lab. Med. 2020, 58, 326–339. [Google Scholar] [CrossRef] [PubMed]
  153. Castro, E.; Jugurnauth-Little, S.; Karlsson, Q.; Al-Shahrour, F.; Piñeiro-Yañez, E.; Van de Poll, F.; Leongamornlert, D.; Dadaev, T.; Govindasami, K.; Guy, M.; et al. High burden of copy number alterations and c-MYC amplification in prostate cancer from BRCA2 germline mutation carriers. Ann. Oncol. 2015, 26, 2293–2300. [Google Scholar] [CrossRef] [PubMed]
  154. Seyedolmohadessin, S.M.; Akbari, M.T.; Nourmohammadi, Z.; Basiri, A.; Pourmand, G. Detection of Loss of Heterozygosity (LOH) Using Circulating Cell-free DNA (cfDNA) by Fluorescence-based Multiplex PCR for Identification of Patients With Prostate Cancer. Appl. Immunohistochem. Mol. Morphol. 2018, 26, 749–759. [Google Scholar] [CrossRef]
  155. Müller, I.; Urban, K.; Pantel, K.; Schwarzenbach, H. Comparison of Genetic Alterations Detected in Circulating Microsatellite DNA in Blood Plasma Samples of Patients with Prostate Cancer and Benign Prostatic Hyperplasia. Ann. N. Y. Acad. Sci. 2006, 1075, 222–229. [Google Scholar] [CrossRef] [PubMed]
  156. Schwarzenbach, H.; Alix-Panabières, C.; Müller, I.; Letang, N.; Vendrell, J.-P.; Rebillard, X.; Pantel, K. Cell-free Tumor DNA in Blood Plasma As a Marker for Circulating Tumor Cells in Prostate Cancer. Clin. Cancer Res. 2009, 15, 1032–1038. [Google Scholar] [CrossRef] [Green Version]
  157. Lau, E.; McCoy, P.; Reeves, F.; Chow, K.; Clarkson, M.; Kwan, E.M.; Packwood, K.; Northen, H.; He, M.; Kingsbury, Z.; et al. Detection of ctDNA in plasma of patients with clinically localised prostate cancer is associated with rapid disease progression. Genome Med. 2020, 12, 72. [Google Scholar] [CrossRef]
  158. Hennigan, S.T.; Trostel, S.Y.; Terrigino, N.T.; Voznesensky, O.S.; Schaefer, R.J.; Whitlock, N.C.; Wilkinson, S.; Carrabba, N.V.; Atway, R.; Shema, S.; et al. Low Abundance of Circulating Tumor DNA in Localized Prostate Cancer. JCO Precis. Oncol. 2019, 3. [Google Scholar] [CrossRef]
  159. Baden, J.; Adams, S.; Astacio, T.; Jones, J.; Markiewicz, J.; Painter, J.; Trust, C.; Wang, Y.; Green, G. Predicting prostate biopsy result in men with prostate specific antigen 2.0 to 10.0 ng/ml using an investigational prostate cancer methylation assay. J. Urol. 2011, 186, 2101–2106. [Google Scholar] [CrossRef]
  160. Bastian, P.J.; Palapattu, G.S.; Lin, X.; Yegnasubramanian, S.; Mangold, L.A.; Trock, B.; Eisenberger, M.A.; Partin, A.W.; Nelson, W.G. Preoperative serum DNA GSTP1 CpG island hypermethylation and the risk of early prostate-specific antigen recurrence following radical prostatectomy. Clin. Cancer Res. 2005, 11, 4037–4043. [Google Scholar] [CrossRef] [Green Version]
  161. Cairns, P.; Esteller, M.; Herman, J.G.; Schoenberg, M.; Jeronimo, C.; Sanchez-Cespedes, M.; Chow, N.H.; Grasso, M.; Wu, L.; Westra, W.B.; et al. Molecular detection of prostate cancer in urine by GSTP1 hypermethylation. Clin. Cancer Res. 2001, 7, 2727–2730. [Google Scholar]
  162. Hendriks, R.J.; Dijkstra, S.; Smit, F.P.; Vandersmissen, J.; Van de Voorde, H.; Mulders, P.F.A.; van Oort, I.M.; Van Criekinge, W.; Schalken, J.A. Epigenetic markers in circulating cell-free DNA as prognostic markers for survival of castration-resistant prostate cancer patients. Prostate 2018, 78, 336–342. [Google Scholar] [CrossRef] [PubMed]
  163. 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]
  164. Moreira-Barbosa, C.; Barros-Silva, D.; Costa-Pinheiro, P.; Torres-Ferreira, J.; Constâncio, V.; Freitas, R.; Oliveira, J.; Antunes, L.; Henrique, R.; Jerónimo, C. Comparing diagnostic and prognostic performance of two-gene promoter methylation panels in tissue biopsies and urines of prostate cancer patients. Clin. Epigenetics 2018, 10, 132. [Google Scholar] [CrossRef] [PubMed]
  165. O’Reilly, E.; Tuzova, A.V.; Walsh, A.L.; Russell, N.M.; O’Brien, O.; Kelly, S.; Dhomhnallain, O.N.; DeBarra, L.; Dale, C.M.; Brugman, R.; et al. epiCaPture: A Urine DNA Methylation Test for Early Detection of Aggressive Prostate Cancer. JCO Precis Oncol. 2019, 2019. [Google Scholar] [CrossRef] [PubMed]
  166. Zhao, F.; Olkhov-Mitsel, E.; van der Kwast, T.; Sykes, J.; Zdravic, D.; Venkateswaran, V.; Zlotta, A.R.; Loblaw, A.; Fleshner, N.E.; Klotz, L.; et al. Urinary DNA Methylation Biomarkers for Noninvasive Prediction of Aggressive Disease in Patients with Prostate Cancer on Active Surveillance. J. Urol. 2017, 197, 335–341. [Google Scholar] [CrossRef] [PubMed]
  167. 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] [Green Version]
  168. Moss, J.; Magenheim, J.; Neiman, D.; Zemmour, H.; Loyfer, N.; Korach, A.; Samet, Y.; Maoz, M.; Druid, H.; Arner, P.; et al. Comprehensive human cell-type methylation atlas reveals origins of circulating cell-free DNA in health and disease. Nat. Commun. 2018, 9, 5068. [Google Scholar] [CrossRef] [Green Version]
  169. Silva, R.; Moran, B.; Russell, N.M.; Fahey, C.; Vlajnic, T.; Manecksha, R.P.; Finn, S.P.; Brennan, D.J.; Gallagher, W.M.; Perry, A.S. Evaluating liquid biopsies for methylomic profiling of prostate cancer. Epigenetics 2020, 15, 715–727. [Google Scholar] [CrossRef] [Green Version]
  170. Gerhauser, C.; Favero, F.; Risch, T.; Simon, R.; Feuerbach, L.; Assenov, Y.; Heckmann, D.; Sidiropoulos, N.; Waszak, S.M.; Hübschmann, D.; et al. Molecular Evolution of Early-Onset Prostate Cancer Identifies Molecular Risk Markers and Clinical Trajectories. Cancer Cell 2018, 34, 996–1011.e1018. [Google Scholar] [CrossRef] [Green Version]
  171. Shen, S.Y.; Burgener, J.M.; Bratman, S.V.; De Carvalho, D.D. Preparation of cfMeDIP-seq libraries for methylome profiling of plasma cell-free DNA. Nat. Protoc. 2019, 14, 2749–2780. [Google Scholar] [CrossRef]
  172. Shen, S.Y.; Singhania, R.; Fehringer, G.; Chakravarthy, A.; Roehrl, M.H.A.; Chadwick, D.; Zuzarte, P.C.; Borgida, A.; Wang, T.T.; Li, T.; et al. Sensitive tumour detection and classification using plasma cell-free DNA methylomes. Nature 2018, 563, 579–583. [Google Scholar] [CrossRef] [PubMed]
  173. Chemi, F.; Pearce, S.P.; Clipson, A.; Hill, S.M.; Conway, A.M.; Richardson, S.A.; Kamieniecka, K.; Caeser, R.; White, D.J.; Mohan, S.; et al. cfDNA methylome profiling for detection and subtyping of small cell lung cancers. Nat Cancer 2022, 3, 1260–1270. [Google Scholar] [CrossRef] [PubMed]
  174. Huang, J.; Soupir, A.C.; Wang, L. Cell-free DNA methylome profiling by MBD-seq with ultra-low input. Epigenetics 2022, 17, 239–252. [Google Scholar] [CrossRef] [PubMed]
  175. Siejka-Zielińska, P.; Cheng, J.; Jackson, F.; Liu, Y.; Soonawalla, Z.; Reddy, S.; Silva, M.; Puta, L.; McCain, M.V.; Culver, E.L.; et al. Cell-free DNA TAPS provides multimodal information for early cancer detection. Sci. Adv. 2021, 7, eabh0534. [Google Scholar] [CrossRef]
  176. Stackpole, M.L.; Zeng, W.; Li, S.; Liu, C.-C.; Zhou, Y.; He, S.; Yeh, A.; Wang, Z.; Sun, F.; Li, Q.; et al. Cost-effective methylome sequencing of cell-free DNA for accurately detecting and locating cancer. Nat. Commun. 2022, 13, 5566. [Google Scholar] [CrossRef]
  177. Chen, S.; Petricca, J.; Ye, W.; Guan, J.; Zeng, Y.; Cheng, N.; Gong, L.; Shen, S.Y.; Hua, J.T.; Crumbaker, M.; et al. The cell-free DNA methylome captures distinctions between localized and metastatic prostate tumors. Nat. Commun. 2022, 13, 6467. [Google Scholar] [CrossRef] [PubMed]
  178. Roest, H.P.; JNM, I.J.; van der Laan, L.J.W. Evaluation of RNA isolation methods for microRNA quantification in a range of clinical biofluids. BMC Biotechnol. 2021, 21, 48. [Google Scholar] [CrossRef]
  179. McKiernan, J.; Donovan, M.J.; O’Neill, V.; Bentink, S.; Noerholm, M.; Belzer, S.; Skog, J.; Kattan, M.W.; Partin, A.; Andriole, G.; et al. A Novel Urine Exosome Gene Expression Assay to Predict High-grade Prostate Cancer at Initial Biopsy. JAMA Oncol. 2016, 2, 882–889. [Google Scholar] [CrossRef] [Green Version]
  180. Tomlins, S.A.; Day, J.R.; Lonigro, R.J.; Hovelson, D.H.; Siddiqui, J.; Kunju, L.P.; Dunn, R.L.; Meyer, S.; Hodge, P.; Groskopf, J.; et al. Urine TMPRSS2:ERG Plus PCA3 for Individualized Prostate Cancer Risk Assessment. Eur. Urol. 2016, 70, 45–53. [Google Scholar] [CrossRef] [Green Version]
  181. Van Neste, L.; Hendriks, R.J.; Dijkstra, S.; Trooskens, G.; Cornel, E.B.; Jannink, S.A.; de Jong, H.; Hessels, D.; Smit, F.P.; Melchers, W.J.; et al. Detection of High-grade Prostate Cancer Using a Urinary Molecular Biomarker-Based Risk Score. Eur. Urol. 2016, 70, 740–748. [Google Scholar] [CrossRef]
  182. Souza, M.F.; Kuasne, H.; Barros-Filho, M.C.; Cilião, H.L.; Marchi, F.A.; Fuganti, P.E.; Rogatto, S.R.; Cólus, I.M.S. Circulating mRNA signature as a marker for high-risk prostate cancer. Carcinogenesis 2020, 41, 139–145. [Google Scholar] [CrossRef] [PubMed]
  183. Mihelich, B.L.; Maranville, J.C.; Nolley, R.; Peehl, D.M.; Nonn, L. Elevated serum microRNA levels associate with absence of high-grade prostate cancer in a retrospective cohort. PLoS ONE 2015, 10, e0124245. [Google Scholar] [CrossRef]
  184. 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] [PubMed] [Green Version]
  185. Cooperberg, M.R.; Pasta, D.J.; Elkin, E.P.; Litwin, M.S.; Latini, D.M.; Du Chane, J.; Carroll, P.R. The University of California, San Francisco Cancer of the Prostate Risk Assessment score: A straightforward and reliable preoperative predictor of disease recurrence after radical prostatectomy. J. Urol. 2005, 173, 1938–1942. [Google Scholar] [CrossRef] [Green Version]
  186. Shen, J.; Hruby, G.W.; McKiernan, J.M.; Gurvich, I.; Lipsky, M.J.; Benson, M.C.; Santella, R.M. Dysregulation of circulating microRNAs and prediction of aggressive prostate cancer. Prostate 2012, 72, 1469–1477. [Google Scholar] [CrossRef] [Green Version]
  187. Hoey, C.; Ahmed, M.; Fotouhi Ghiam, A.; Vesprini, D.; Huang, X.; Commisso, K.; Commisso, A.; Ray, J.; Fokas, E.; Loblaw, D.A.; et al. Circulating miRNAs as non-invasive biomarkers to predict aggressive prostate cancer after radical prostatectomy. J. Transl. Med. 2019, 17, 173. [Google Scholar] [CrossRef] [Green Version]
  188. Wang, Y.T.; Shi, T.; Srivastava, S.; Kagan, J.; Liu, T.; Rodland, K.D. Proteomic Analysis of Exosomes for Discovery of Protein Biomarkers for Prostate and Bladder Cancer. Cancers 2020, 12, 2335. [Google Scholar] [CrossRef]
  189. Nilsson, J.; Skog, J.; Nordstrand, A.; Baranov, V.; Mincheva-Nilsson, L.; Breakefield, X.O.; Widmark, A. Prostate cancer-derived urine exosomes: A novel approach to biomarkers for prostate cancer. Br. J. Cancer 2009, 100, 1603–1607. [Google Scholar] [CrossRef]
  190. Zabegina, L.; Nazarova, I.; Nikiforova, N.; Slyusarenko, M.; Sidina, E.; Knyazeva, M.; Tsyrlina, E.; Novikov, S.; Reva, S.; Malek, A. A New Approach for Prostate Cancer Diagnosis by miRNA Profiling of Prostate-Derived Plasma Small Extracellular Vesicles. Cells 2021, 10, 2372. [Google Scholar]
  191. Woo, J.; Santasusagna, S.; Banks, J.; Pastor-Lopez, S.; Yadav, K.; Carceles-Cordon, M.; Dominguez-Andres, A.; Den, R.B.; Languino, L.R.; Pippa, R.; et al. Urine Extracellular Vesicle GATA2 mRNA Discriminates Biopsy Result in Men with Suspicion of Prostate Cancer. J. Urol. 2020, 204, 691–700. [Google Scholar] [CrossRef]
  192. Dhondt, B.; Geeurickx, E.; Tulkens, J.; Van Deun, J.; Vergauwen, G.; Lippens, L.; Miinalainen, I.; Rappu, P.; Heino, J.; Ost, P.; et al. Unravelling the proteomic landscape of extracellular vesicles in prostate cancer by density-based fractionation of urine. J. Extracell. Vesicles 2020, 9, 1736935. [Google Scholar] [CrossRef] [PubMed]
  193. Fujita, K.; Nonomura, N. Urinary biomarkers of prostate cancer. Int. J. Urol. 2018, 25, 770–779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Vasseur, A.; Kiavue, N.; Bidard, F.C.; Pierga, J.Y.; Cabel, L. Clinical utility of circulating tumor cells: An update. Mol. Oncol. 2021, 15, 1647–1666. [Google Scholar] [CrossRef] [PubMed]
  195. Boya, M.; Chu, C.H.; Liu, R.; Ozkaya-Ahmadov, T.; Sarioglu, A.F. Circulating Tumor Cell Enrichment Technologies. In Tumor Liquid Biopsies; Recent Results in Cancer Research Book Series; Springer: Berlin/Heidelberg, Germany, 2020; Volume 215, pp. 25–55. [Google Scholar] [CrossRef]
  196. Malihi, P.D.; Graf, R.P.; Rodriguez, A.; Ramesh, N.; Lee, J.; Sutton, R.; Jiles, R.; Ruiz Velasco, C.; Sei, E.; Kolatkar, A.; et al. Single-Cell Circulating Tumor Cell Analysis Reveals Genomic Instability as a Distinctive Feature of Aggressive Prostate Cancer. Clin. Cancer Res. 2020, 26, 4143–4153. [Google Scholar] [CrossRef]
  197. Wang, X.; Grasso, C.S.; Jordahl, K.M.; Kolb, S.; Nyame, Y.A.; Wright, J.L.; Ostrander, E.A.; Troyer, D.A.; Lance, R.; Feng, Z.; et al. Copy number alterations are associated with metastatic-lethal progression in prostate cancer. Prostate Cancer Prostatic Dis. 2020, 23, 494–506. [Google Scholar] [CrossRef]
  198. Harshman, L.C.; Wang, V.X.; Hamid, A.A.; Santone, G.; Drake, C.G.; Carducci, M.A.; DiPaola, R.S.; Fichorova, R.N.; Sweeney, C.J. Impact of baseline serum IL-8 on metastatic hormone-sensitive prostate cancer outcomes in the Phase 3 CHAARTED trial (E3805). Prostate 2020, 80, 1429–1437. [Google Scholar] [CrossRef]
  199. Gong, Y.; Chippada-Venkata, U.D.; Oh, W.K. Roles of matrix metalloproteinases and their natural inhibitors in prostate cancer progression. Cancers 2014, 6, 1298–1327. [Google Scholar] [CrossRef] [Green Version]
  200. Shajarehpoor Salavati, L.; Tafvizi, F.; Manjili, H.K. The association between MMP2 -1306 C > T (rs243865) polymorphism and risk of prostate cancer. Ir. J. Med. Sci. 2017, 186, 103–111. [Google Scholar] [CrossRef]
  201. Dong, M.; Lih, T.M.; Chen, S.Y.; Cho, K.C.; Eguez, R.V.; Höti, N.; Zhou, Y.; Yang, W.; Mangold, L.; Chan, D.W.; et al. Urinary glycoproteins associated with aggressive prostate cancer. Theranostics 2020, 10, 11892–11907. [Google Scholar] [CrossRef]
  202. Di Minno, A.; Aveta, A.; Gelzo, M.; Tripodi, L.; Pandolfo, S.D.; Crocetto, F.; Imbimbo, C.; Castaldo, G. 8-Hydroxy-2-Deoxyguanosine and 8-Iso-Prostaglandin F2α: Putative Biomarkers to assess Oxidative Stress Damage Following Robot-Assisted Radical Prostatectomy (RARP). J. Clin. Med. 2022, 11, 6102. [Google Scholar] [CrossRef]
  203. Pantel, K.; Alix-Panabières, C. Liquid biopsy and minimal residual disease—Latest advances and implications for cure. Nat. Rev. Clin. Oncol. 2019, 16, 409–424. [Google Scholar] [CrossRef] [PubMed]
  204. Yang, Y.; Liu, Z.; Wang, Q.; Chang, K.; Zhang, J.; Ye, D.; Kong, Y.; Dai, B. Presence of CD133-positive circulating tumor cells predicts worse progression-free survival in patients with metastatic castration-sensitive prostate cancer. Int. J. Urol. 2022, 29, 383–389. [Google Scholar] [CrossRef] [PubMed]
  205. Broncy, L.; Paterlini-Bréchot, P. Clinical Impact of Circulating Tumor Cells in Patients with Localized Prostate Cancer. Cells 2019, 8, 676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Rodriguez-Martinez, A.; Simon-Saez, I.; Perales, S.; Garrido-Navas, C.; Russo, A.; de Miguel-Perez, D.; Puche-Sanz, I.; Alaminos, C.; Ceron, J.; Lorente, J.A.; et al. Exchange of cellular components between platelets and tumor cells: Impact on tumor cells behavior. Theranostics 2022, 12, 2150–2161. [Google Scholar] [CrossRef]
  207. Clar, K.L.; Hinterleitner, C.; Schneider, P.; Salih, H.R.; Maurer, S. Inhibition of NK Reactivity against Solid Tumors by Platelet-Derived RANKL. Cancers 2019, 11, 277. [Google Scholar] [CrossRef] [Green Version]
  208. Tjon-Kon-Fat, L.A.; Lundholm, M.; Schröder, M.; Wurdinger, T.; Thellenberg-Karlsson, C.; Widmark, A.; Wikström, P.; Nilsson, R.J.A. Platelets harbor prostate cancer biomarkers and the ability to predict therapeutic response to abiraterone in castration resistant patients. Prostate 2018, 78, 48–53. [Google Scholar] [CrossRef]
  209. TracerX. TRAcking Cancer Evolution through Therapy (Rx). Available online: http://tracerx.co.uk/studies/prostate/ (accessed on 26 November 2022).
  210. Turajlic, S.; Sottoriva, A.; Graham, T.; Swanton, C. Resolving genetic heterogeneity in cancer. Nat. Rev. Genet 2019, 20, 404–416. [Google Scholar] [CrossRef]
  211. Xi, J.; Yuan, X.; Wang, M.; Li, A.; Li, X.; Huang, Q. Inferring subgroup-specific driver genes from heterogeneous cancer samples via subspace learning with subgroup indication. Bioinformatics 2020, 36, 1855–1863. [Google Scholar] [CrossRef]
  212. Liu, S.; Wang, W.; Zhao, Y.; Liang, K.; Huang, Y. Identification of Potential Key Genes for Pathogenesis and Prognosis in Prostate Cancer by Integrated Analysis of Gene Expression Profiles and the Cancer Genome Atlas. Front. Oncol. 2020, 10, 809. [Google Scholar] [CrossRef]
  213. The Cancer Gene Census (CGC). Available online: https://cancer.sanger.ac.uk/census (accessed on 26 November 2022).
  214. Kesch, C.; Radtke, J.P.; Wintsche, A.; Wiesenfarth, M.; Luttje, M.; Gasch, C.; Dieffenbacher, S.; Pecqueux, C.; Teber, D.; Hatiboglu, G.; et al. Correlation between genomic index lesions and mpMRI and (68)Ga-PSMA-PET/CT imaging features in primary prostate cancer. Sci. Rep. 2018, 8, 16708. [Google Scholar] [CrossRef] [Green Version]
  215. Norris, J.M.; Simpson, B.S.; Parry, M.A.; Allen, C.; Ball, R.; Freeman, A.; Kelly, D.; Kirkham, A.; Kasivisvanathan, V.; Whitaker, H.C.; et al. Genetic landscape of prostate cancer conspicuity on multiparametric MRI: A protocol for a systematic review and bioinformatic analysis. BMJ Open 2020, 10, e034611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Morrison, G.; Buckley, J.; Ostrow, D.; Varghese, B.; Cen, S.Y.; Werbin, J.; Ericson, N.; Cunha, A.; Lu, Y.T.; George, T.; et al. Non-Invasive Profiling of Advanced Prostate Cancer via Multi-Parametric Liquid Biopsy and Radiomic Analysis. Int. J. Mol. Sci. 2022, 23, 2571. [Google Scholar] [CrossRef] [PubMed]
  217. Alarcón-Zendejas, A.P.; Scavuzzo, A.; Jiménez-Ríos, M.A.; Álvarez-Gómez, R.M.; Montiel-Manríquez, R.; Castro-Hernández, C.; Jiménez-Dávila, M.A.; Pérez-Montiel, D.; González-Barrios, R.; Jiménez-Trejo, F.; et al. The promising role of new molecular biomarkers in prostate cancer: From coding and non-coding genes to artificial intelligence approaches. Prostate Cancer Prostatic Dis. 2022, 25, 431–443. [Google Scholar] [CrossRef] [PubMed]
  218. Schwartzberg, L.S.; Li, G.; Tolba, K.; Bourla, A.B.; Schulze, K.; Gadgil, R.; Fine, A.; Lofgren, K.T.; Graf, R.P.; Oxnard, G.R.; et al. Complementary Roles for Tissue- and Blood-Based Comprehensive Genomic Profiling for Detection of Actionable Driver Alterations in Advanced NSCLC. JTO Clin. Res. Rep. 2022, 3, 100386. [Google Scholar] [CrossRef] [PubMed]
  219. Peneder, P.; Stutz, A.M.; Surdez, D.; Krumbholz, M.; Semper, S.; Chicard, M.; Sheffield, N.C.; Pierron, G.; Lapouble, E.; Totzl, M.; et al. Multimodal analysis of cell-free DNA whole-genome sequencing for pediatric cancers with low mutational burden. Nat. Commun. 2021, 12, 3230. [Google Scholar] [CrossRef] [PubMed]
  220. Lee, D.; Park, Y.; Kim, S. Towards multi-omics characterization of tumor heterogeneity: A comprehensive review of statistical and machine learning approaches. Brief. Bioinform. 2021, 22, bbaa188. [Google Scholar] [CrossRef]
  221. Onuchic, V.; Hartmaier, R.J.; Boone, D.N.; Samuels, M.L.; Patel, R.Y.; White, W.M.; Garovic, V.D.; Oesterreich, S.; Roth, M.E.; Lee, A.V.; et al. Epigenomic Deconvolution of Breast Tumors Reveals Metabolic Coupling between Constituent Cell Types. Cell Rep. 2016, 17, 2075–2086. [Google Scholar] [CrossRef] [Green Version]
  222. De Laere, B.; Oeyen, S.; Mayrhofer, M.; Whitington, T.; van Dam, P.J.; Van Oyen, P.; Ghysel, C.; Ampe, J.; Ost, P.; Demey, W.; et al. TP53 Outperforms Other Androgen Receptor Biomarkers to Predict Abiraterone or Enzalutamide Outcome in Metastatic Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2019, 25, 1766–1773. [Google Scholar] [CrossRef] [Green Version]
  223. De Laere, B.; van Dam, P.J.; Whitington, T.; Mayrhofer, M.; Diaz, E.H.; Van den Eynden, G.; Vandebroek, J.; Del-Favero, J.; Van Laere, S.; Dirix, L.; et al. Comprehensive Profiling of the Androgen Receptor in Liquid Biopsies from Castration-resistant Prostate Cancer Reveals Novel Intra-AR Structural Variation and Splice Variant Expression Patterns. Eur. Urol. 2017, 72, 192–200. [Google Scholar] [CrossRef]
  224. Kohli, M.; Li, J.; Du, M.; Hillman, D.W.; Dehm, S.M.; Tan, W.; Carlson, R.; Campion, M.B.; Wang, L.; Wang, L.; et al. Prognostic association of plasma cell-free DNA-based androgen receptor amplification and circulating tumor cells in pre-chemotherapy metastatic castration-resistant prostate cancer patients. Prostate Cancer Prostatic Dis. 2018, 21, 411–418. [Google Scholar] [CrossRef]
  225. Hofmann, L.; Sallinger, K.; Haudum, C.; Smolle, M.; Heitzer, E.; Moser, T.; Novy, M.; Gesson, K.; Kroneis, T.; Bauernhofer, T.; et al. A Multi-Analyte Approach for Improved Sensitivity of Liquid Biopsies in Prostate Cancer. Cancers 2020, 12, 2247. [Google Scholar] [CrossRef] [PubMed]
  226. Fettke, H.; Kwan, E.M.; Docanto, M.M.; Bukczynska, P.; Ng, N.; Graham, L.K.; Mahon, K.; Hauser, C.; Tan, W.; Wang, X.H.; et al. Combined Cell-free DNA and RNA Profiling of the Androgen Receptor: Clinical Utility of a Novel Multianalyte Liquid Biopsy Assay for Metastatic Prostate Cancer. Eur. Urol. 2020, 78, 173–180. [Google Scholar] [CrossRef] [PubMed]
  227. Connell, S.P.; O’Reilly, E.; Tuzova, A.; Webb, M.; Hurst, R.; Mills, R.; Zhao, F.; Bapat, B.; Cooper, C.S.; Perry, A.S.; et al. Development of a multivariable risk model integrating urinary cell DNA methylation and cell-free RNA data for the detection of significant prostate cancer. Prostate 2020, 80, 547–558. [Google Scholar] [CrossRef] [PubMed]
  228. Burgener, J.M.; Zou, J.; Zhao, Z.; Zheng, Y.; Shen, S.Y.; Huang, S.H.; Keshavarzi, S.; Xu, W.; Liu, F.F.; Liu, G.; et al. Tumor-Naive Multimodal Profiling of Circulating Tumor DNA in Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2021, 27, 4230–4244. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Common mutations detected in PCa are shown according to their occurrence along the disease progression. The bubble size indicates the frequency of mutation. Created with BioRender.com (Accessed on 18 October 2022).
Figure 1. Common mutations detected in PCa are shown according to their occurrence along the disease progression. The bubble size indicates the frequency of mutation. Created with BioRender.com (Accessed on 18 October 2022).
Cancers 14 06094 g001
Figure 2. Common genomic and epigenomic instabilities detected in tissue, blood, and urine of PCa patients [17,29,35,39,55,56]. Created with BioRender.com.
Figure 2. Common genomic and epigenomic instabilities detected in tissue, blood, and urine of PCa patients [17,29,35,39,55,56]. Created with BioRender.com.
Cancers 14 06094 g002
Figure 3. Multi-parametric approach converges the data derived from several sources (i.e., tissue and liquid biopsy, as well as MRI and histology) and uses them as an input for machine learning algorithms. Subsequent to this validation, this approach offers great potential to create a comprehensive patient profile and improves clinical outcomes. Created with BioRender.com (Accessed on 9 November 2022).
Figure 3. Multi-parametric approach converges the data derived from several sources (i.e., tissue and liquid biopsy, as well as MRI and histology) and uses them as an input for machine learning algorithms. Subsequent to this validation, this approach offers great potential to create a comprehensive patient profile and improves clinical outcomes. Created with BioRender.com (Accessed on 9 November 2022).
Cancers 14 06094 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eickelschulte, S.; Riediger, A.L.; Angeles, A.K.; Janke, F.; Duensing, S.; Sültmann, H.; Görtz, M. Biomarkers for the Detection and Risk Stratification of Aggressive Prostate Cancer. Cancers 2022, 14, 6094. https://doi.org/10.3390/cancers14246094

AMA Style

Eickelschulte S, Riediger AL, Angeles AK, Janke F, Duensing S, Sültmann H, Görtz M. Biomarkers for the Detection and Risk Stratification of Aggressive Prostate Cancer. Cancers. 2022; 14(24):6094. https://doi.org/10.3390/cancers14246094

Chicago/Turabian Style

Eickelschulte, Samaneh, Anja Lisa Riediger, Arlou Kristina Angeles, Florian Janke, Stefan Duensing, Holger Sültmann, and Magdalena Görtz. 2022. "Biomarkers for the Detection and Risk Stratification of Aggressive Prostate Cancer" Cancers 14, no. 24: 6094. https://doi.org/10.3390/cancers14246094

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