Next Article in Journal
Trypsin Genes Are Regulated through the miRNA Bantam and Associated with Drug Sensitivity in the Sea Louse Caligus rogercresseyi
Next Article in Special Issue
Role of MicroRNAs in Neuroendocrine Prostate Cancer
Previous Article in Journal
Co-Regulation of Protein Coding Genes by Transcription Factor and Long Non-Coding RNA in SARS-CoV-2 Infected Cells: An In Silico Analysis
Previous Article in Special Issue
MicroRNAs as Potential Biomarkers in Pituitary Adenomas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of Neuroendocrine-like Differentiation in Prostate Cancer by Non-Coding RNAs

by
Eva Slabáková
,
Zuzana Kahounová
,
Jiřina Procházková
and
Karel Souček
*
Department of Cytokinetics, Institute of Biophysics of the Czech Academy of Sciences, 61265 Brno, Czech Republic
*
Author to whom correspondence should be addressed.
Authors contributed equally to this study.
Non-Coding RNA 2021, 7(4), 75; https://doi.org/10.3390/ncrna7040075
Submission received: 7 September 2021 / Revised: 18 November 2021 / Accepted: 29 November 2021 / Published: 2 December 2021
(This article belongs to the Collection Role of microRNA in Neuroendocrine Neoplasms)

Abstract

:
Neuroendocrine prostate cancer (NEPC) represents a variant of prostate cancer that occurs in response to treatment resistance or, to a much lesser extent, de novo. Unravelling the molecular mechanisms behind transdifferentiation of cancer cells to neuroendocrine-like cancer cells is essential for development of new treatment opportunities. This review focuses on summarizing the role of small molecules, predominantly microRNAs, in this phenomenon. A published literature search was performed to identify microRNAs, which are reported and experimentally validated to modulate neuroendocrine markers and/or regulators and to affect the complex neuroendocrine phenotype. Next, available patients’ expression datasets were surveyed to identify deregulated microRNAs, and their effect on NEPC and prostate cancer progression is summarized. Finally, possibilities of miRNA detection and quantification in body fluids of prostate cancer patients and their possible use as liquid biopsy in prostate cancer monitoring are discussed. All the addressed clinical and experimental contexts point to an association of NEPC with upregulation of miR-375 and downregulation of miR-34a and miR-19b-3p. Together, this review provides an overview of different roles of non-coding RNAs in the emergence of neuroendocrine prostate cancer.

Graphical Abstract

1. Introduction

Neuroendocrine prostate cancer (NEPC) is a highly aggressive treatment-resistant stage of prostate cancer (PCa) with poor patient outcome. It often occurs after long-term androgen-deprivation therapy; however, de novo pure neuroendocrine tumours are also described. Prostate cancer cells undergo robust phenotypic changes (so called transdifferentiation), resulting in the neuroendocrine-like phenotype. This is associated with low or absent signalling of the androgen receptor (AR), neuron-like morphology, expression and secretion of neuropeptides and biologically active factors, and deregulation of expression of several molecular drivers (e.g., Rb, TP53, MYCN) [1]. The molecular machinery behind the development of NEPC is still being investigated, as well as the elucidation of the role of microRNAs in the induction and regulation of NEPC. So far, deregulation of several oncogenic miRNAs (e.g., miR-21, miR-141, miR-32, miR-650, miR-106b/miR-25 cluster, and others) as well as tumour suppressor miRNAs (e.g., miR-34, miR-145, miR-200b, let-7 miRNAs, and others) was described in the context of PCa (summarized in [2]). Moreover, miRNAs are also investigated as potential biomarkers for PCa diagnosis. Interestingly, one molecule can exert the opposite action in different contexts. Namely, miR-204 acts as a tumour suppressor in prostate cancer cells and as an oncomiR in neuroendocrine cancer cells [3]. Similarly, LINC00261 was described as a tumour suppressor in multiple cancers [4], while promoting tumorigenesis in prostate neuroendocrine cells [5]. Therefore, a deep understanding of the involvement of particular miRNAs in NEPC emergence and pathology is desirable. This review article summarizes the findings concerning potential miRNA and other non-coding RNA regulators in the context of neuroendocrine transdifferentiation of advanced prostate cancer.

Neuroendocrine Prostate Cancer

Prostate cancer (PCa) is the second most often diagnosed and fifth leading cause of cancer death among men worldwide [6]. Primary therapy of clinically localized non-metastatic disease consists of radical prostatectomy and/or radiotherapy [7]. Since the growth and function of the prostate gland as well as cancer cells are dependent on the action of steroid hormones androgens, targeting androgen production or signalling of AR is used for treatment of recurrent and metastatic PCa. Although initially responding to androgen deprivation therapy (ADT), cancer cells can adapt to ADT and restore AR signalling under low levels of androgens, and consequently, the disease progresses to more aggressive castration-resistant prostate cancer (CRPC) [8]. Since CRPC is still dependent on AR signalling, next-generation AR pathway inhibitors (ARPIs) are used to inhibit intratumoral androgen biosynthesis (abiraterone acetate) or block AR function (enzalutamide) with significant clinical benefits [9]. However, in about 20–25% of patients, resistance may also develop to these newer agents following ADT. Several mechanisms of resistance were described as a consequence of the restoration of AR signalling—gain-of-function mutations of AR, upregulation of the constitutive active AR splice variant, increased intratumoral androgen biosynthesis, or bypassing AR signalling by signalling of glucocorticoid receptor (GR) (summarized in [8,10]). Prolonged androgen inhibition treatment leads to progression to the advanced stage of CRPC associated with reversible transdifferentiation of cancer cells, which lose prostate cell characteristics and acquire neuroendocrine characteristics. This highly aggressive stage with rapid tumour dissemination and therapy resistance is called treatment-emergent neuroendocrine prostate cancer (t-NEPC) (summarized in [1,11]).
The prostate epithelium is composed of basal, luminal, and neuroendocrine cells (NE cells). NE cells are the least abundant cell type in prostate epithelium and are scattered among basal and luminal cells. The origin of NE cells is still ambiguous; NE cells are either derived from common prostate stem cell (summarized in [12]) or they migrate from the neural crest to the glandular prostate epithelium [13]. It is assumed that NE cells are involved in the growth and differentiation of the prostate as well as in the regulation of secretion of the prostatic gland [14]. NE cells do not express AR or the prostate specific antigen (PSA) and are postmitotic [15]. These cells contain and secrete a variety of factors and neuropeptides, which can act in an endocrine, autocrine, and paracrine manner on target cells, as discussed subsequently (summarized in [16,17]). NE cells are found scattered also in prostate adenocarcinoma with a similar frequency to normal prostate epithelium (no more than 1%) (reviewed in [17]). However, in advanced stages (metastatic CRPC, mCRPC) an increasing number of foci of cells with NE characteristics are found.
NE cells and NE-like cancer cells express and secrete a broad spectrum of biological active factors and neuropeptides, denominated together as NED markers. Some of these markers are also detected in patients’ blood/serum/plasma, which makes them useful biomarkers of cancer progression. Importantly, both NE cells and surrounding cancer cells express several receptors for these factors; therefore, these factors can act in an autocrine or paracrine manner and can support the growth of both cell types (summarized in [18,19]). For detailed characterisation and role of NED markers in PCa and/or NEPC, see Table 1.
De novo pure neuroendocrine tumours, so called small cell carcinomas (SCC), are a very rare (0.5–2.0%) and very aggressive subtype of NEPC. They are characterized by low PSA levels, short or no response to conventional ADT, and the presence of lytic bone metastasis and intracranial metastasis (summarized in [93]). A total of 94% of SCCs were positive for at least one NE marker (CgA, NSE, Syp, CD56) [94]. Up to 50% of men diagnosed with SSC have a history of conventional prostatic carcinoma [93]. More often, cancer cells with similar characteristics to SCC mixed with adenocarcinoma are found. This is frequently found in patients progressing after ADT [95]. NEPC is characterized by low or absent AR signalling, loss of RB1 and TP53, amplification of MYCN, ERG rearrangement, upregulation of BRN2, down-regulation of DNA methyltransferases and altered DNA methylation, and upregulation of EZH2 and Polycomb-mediated gene silencing. These tumours are positive for NED markers CgA, NSE, SYP, or CD56 and negative for luminal markers (PSA and PAP). Patients with NEPC are treated with platinum-based chemotherapy, and the survival ranges from 7 months to 2 years (summarized in [1,10]).
The origin of NEPC is still ambiguous. The possible mechanisms leading to NED induction are followed: AR-targeted therapies, other therapies (cyclooxygenase-2 inhibitors, genistein, ionizing radiation), various cells from tumour microenvironment (cancer-associated fibroblasts, mast cells, macrophages, bone marrow-derived cells), Ca2+ ion channels and Ca2+ ion homeostasis, or exosomes [1]. Recent studies show that the lineage plasticity (transition from one developmental pathway to another) also plays a role in the context of NEPC development and therapy resistance. Specifically, the lineage plasticity is associated with the acquisition of independence on AR signalling and treatment resistance in about 20% of advanced PCa patients. This progressive state of CRPC is associated with the loss of AR-regulated lineage characteristics (luminal epithelial phenotype) and, in some situations, the acquisition of new phenotypes (e.g., NE features, NEPC), with involvement of metabolic, genetic, and epigenetic changes (summarized in [96]). There is evidence that NE-like cancer cells in CRPC arise through transdifferentiation from luminal epithelial cells in the mouse CRPC model [97]. Recently, Dong and colleagues proposed a model of PCa development, where NE-like cancer cells arise through transdifferentiation of luminal cancer cells, and these NE-like cancer cells are responsible for at first focal NED, which evolves in pure NEPC [98]. Nouri and colleagues showed that, in response to ARPIs, androgen-sensitive PCa cells are reprogrammed to cancer stem-like cells with characteristics of metastable neural/neural crest stem cells, which can transdifferentiate in neuroendocrine-like PCa cells [99]. Importantly, since the neural crest-derived origin of major fraction of normal NE cells in both human and mouse prostate was experimentally described [13], the possible origin of NE-like cancer cells from these neural crest-derived NE cells should be also taken into account and examined. Further research is needed to understand the evolution of CRPC and NEPC and the molecular machinery behind this to be able to develop potent treatment strategies.

2. Regulatory Circuits Driving Neuroendocrine Differentiation in Prostate Cancer

During the acquisition of neuroendocrine phenotype, prostate cancer cells have been reported to undergo complex remodelling of their transcriptional and phenotypical landscapes (reviewed in, e.g., [100,101,102]). The list of underlying molecular mechanisms growths year by year [103,104] and also involves processes such as enhanced infiltration of the primary tumour with early neural progenitors or direct interaction of cancer cells with nerves present in the reshaping tumour microenvironment [105,106,107]. Since the complexity of such regulatory circuits is multispectral and still emerges, here, we briefly preferentially summarize those driving events that were experimentally validated in in vivo and/or in vitro studies and are represented by deregulated functions of specific tumour suppressors, oncogenes, and transcription factors (TFs) in prostate cancer cells.

2.1. Signalling and Genetic Hallmarks of mCRPC Samples

Integrative genomic analysis of mCRPC samples revealed an accumulation of multiple somatic aberrations in genes encoding AR (amplification) and tumour suppressors p53, PTEN, and RB1 (deletions and/or mutations) [108]. These genetic perturbations support genomic instability, cancer cell survival, dedifferentiation, and pro-neuronal differentiation [109,110,111]. The complexity and type of changes that lead to the deregulation of AR, PTEN, p53, and pRB signalling during PCa progression is a matter of intensive investigation and vivid discussions [112,113,114,115,116]. They may encompass processes such as expression of constitutively active AR variants [117], ligand-independent activation of AR [118], or expression of gain-of-function p53 mutants, which have been recently demonstrated to induce conversion of fibroblasts to a cancer-associated phenotype that supports increased tumour growth and metastasis [108,119,120]. Acquisition of the neuroendocrine program is accompanied by genetic alterations and rewiring of other important signalling networks such as those mediated by Aurora kinase A (AURKA) or PI3K/AKT [121,122,123,124].

2.2. ‘Lost and Found’ Protein Keys Unlocking Neuroendocrine Trans-Differentiation of Prostate Cancer

EHF (also known as ESE3) is an epithelial-specific ETS transcription factor previously shown to be highly expressed in normal prostate tissue, where it prevents prostate pathogenesis and contributes to the maintenance of homeostasis and differentiation status of epithelial cells. EHF expression is reduced in PCa samples, and its re-expression inhibits the clonogenic survival of PCa cells and promotes their apoptosis [125]. EHF deficiency or loss induces an epithelial-to-mesenchymal transition (EMT) and endows epithelial prostate cells with stem-like features and tumour-initiating and metastatic properties [126]. Recently, EHF loss has been demonstrated to facilitate the development of treatment-induced NEPC via transcriptional de-repression of EZH2 and LIN28B and consequential deregulation of let-7 miRNAs expression and its maturation [127,128,129].
A similar impact on NEPC development is caused by the RE1 silencing transcription factor (REST), which is known as a transcriptional repressor of neuronal genes in neural progenitors and in non-neuronal tissue including prostate [111,130,131]. Loss of REST activity mediated by a splicing regulator serine/arginine repetitive matrix 4 (SRRM4) has been suggested to promote the emergence of the NE phenotype in CRPC [111] and endows cancer cells with stemness and neuroendocrine features most likely by de-repressing expression of REST targets such as CD44, Twist1, and secretagogin (SCGN) [131,132].
Finally, a transcription factor forkhead box A1 (FOXA1), previously reported due to its pioneering chromatin remodelling activity, which mediates an access of various nuclear receptors to their target regulatory regions, is an internal component of transcriptional program controlling AR signalling status [101,133]. FOXA1 co-regulates an AR-mediated transcriptional program in healthy prostate and in primary tumours via making response elements (ARE) accessible to AR and/or by direct interaction with AR itself. Additionally, in cooperation with other TFs, FOXA1 facilitates the oncogenic switch of AR signalling, yet its expression has a tumour-suppressive impact on the progression of primary PCa to NEPC [134,135,136,137].
Generally, TFs involved in nervous system development, especially those engaged in the transcriptional control of early neurogenesis, play essential roles in the process of prostate cancer neuroendocrine differentiation [138,139,140,141,142,143,144,145]. Sex-determining region Y2 (SOX2) [146], Achaete-scute family BHLH transcription factor 1 (ASCL1; also known as hASH1 or Mash1) [147,148,149], Doublecortin (DCX) [106,150], Forkhead box A2 (FOXA2) [151,152], POU class 3 homeobox 2 and 4 (POU3F2 and 4 also known as BRN2 and 4) [153,154], Neuroblastoma MYC oncogene (MYCN) [121,122,155,156], and One cut homeobox 2 (ONECUT2) [157] have been all demonstrated as active components of NEPC development and represent functional molecular tools essential for the reshaping of prostate epithelial tumours into neuroendocrine [98]. Similarly, remodelling of epigenetic and transcriptional landscapes mediated by transcriptional repressor Enhancer of Zeste homolog 2 (EZH2) [122,158,159], pro-neuronal splicing regulator SRRM4 [160,161], or transposable element Paternally expressed 10 (PEG10) [162,163] has been reported to promote neuroendocrine differentiation of PCa as well.

2.3. Rewiring, Remodelling, and Reshaping—A Vicious Program Turned on

The rewiring of signalling networks together with the remodelling of transcriptional program goes hand in hand with the reshaping of phenotypic landscapes displayed by prostate cancer cells. Mechanistic studies have not yet fully revealed the network of mutual crosstalks between particular driving events and phenotypic parameters (Figure 1). ADT represents dramatic selective pressure, which, together with the loss of REST, an AR co-repressor and master negative regulator of neurogenesis, represents an important prerequisite for the triggering of lineage plasticity [164,165,166]. N-MYC was demonstrated to be stabilized by AURKA through a kinase-independent process and, together with constitutively active AKT1 (a consequence of N-MYC-mediated AR signalling abrogation and/or PTEN inactivation) and via direct interaction with EZH2, acts as a master driver of NEPC initiated from prostate epithelial cells [122,155]. In non-small cell lung cancer (NSCLC), N-MYC has been also reported as a downstream target of SOX2 [167]. Once AR activity is intervened during ADT, the master neural transcription factor BRN2 is shown together with SOX2 to release from AR-mediated suppression and drive expression of e.g., ASCL1 and other members of pro-neural gene battery [153]. Interestingly, BRN2 and BRN4 have been both demonstrated as internal components of extracellular vesicles released by prostate cancer cells and to promote NED [154]. SOX2 itself stands as critical promoter of lineage plasticity and androgen resistance in TP53- and RB1-deficient prostate cancer; it reprograms transcriptional circuits in favour of a pro-neuronal specific gene expression pattern, which includes accumulation of ASCL1-positive neural progenitors and DCX-positive neuroblasts, both events known from SOX2-mediated adult neurogenesis in the brain [146,168,169]. Nuclear ASCL1 expression seems to persist in neuroendocrine prostate cancer cells and, moreover, DCX-positive neural progenitors from the central nervous system have been demonstrated to infiltrate prostate tumours and metastases and initiate tumour neurogenesis, thus contributing to the stabilization and promotion of neuroendocrine phenotype [106,149]. Both FOXA1 and FOXA2 factors act as pioneering chromatin remodelling factors for AR signalling and NE-specific transcription, with FOXA2 being strongly expressed in association with SYP-positive neuroendocrine prostate carcinoma samples, high-grade adenocarcinomas, and castration-resistant prostate cancer [151,152,170,171]. In mCRPC, a direct negative regulation of AR signalling (including its downstream targets such as EHF) and FOXA1 by ONECUT2 has been reported by Rotinen et al., together with increased expression of PEG10, a putative target gene of ONECUT2 being in primary PCa repressed by AR and REST [172]. ONECUT2 regulates HIF1α binding to its response elements and in synergism with SMAD3 and hypoxic conditions it activates a transcription program specific for mCRPC as well as it drives tumour aggressiveness and plasticity in NEPC [157].
Direct identification of neuroendocrine cells in mCRPC samples via their surface fingerprint may provide essential clues on how to target aggressive neuroendocrine prostate cancer cells. Interestingly, Mucin 1 (MUC1) is known as a transmembrane surface protein with an altered glycosylation pattern in prostate cancer cells [173]. Additionally, its cleavage may lead to the nuclear localization of the C-terminal part (MUC1-C), formation of chromatin-based protein interactions, and cell type context-dependent alteration of the transcription program [174]. Indeed, such a role that oncoprotein MUC1-C plays in lineage plasticity driving NEPC has been recently reported [175]. Moreover, MUC1-C is highly expressed in advanced PCa; it suppresses AR signalling, activates MYCN and BRN2 pathways, and drives expression of stemness-specific master regulators such as SOX2, NANOG, and NOTCH1 signalling [176]. Another example of surface molecule involved in NEPC development is represented by Trop2, a tumour-associated calcium signal transducer 2 (TACSTD2). Luminal epithelial cells highly positive for surface Trop2 (TACSTD2) that express high levels of SOX2 are more predisposed to NED [177]; they drive the NE phenotype together with PARP1 and are predictive of recurrence of localised PCa [178].

3. miRNAs as Multifaceted Crossroads Driving Neuroendocrine Prostate Cancer Development

3.1. miRNA Biogenesis and Mechanisms of Action

MicroRNAs (miRNAs) are evolutionarily highly conserved non-coding RNA molecules, exerting both pro- and anti-tumorigenic effects in prostate cancer [2]. Although miRNAs accomplish important regulatory functions at all stages of cancer progression, their clinical relevance in cancer diagnosis, outcome prediction, and targeted therapy is still a matter of debate and investigation [179].
More than 2000 miRNA genes in the human genome [180] are located either in protein-coding or non-coding regions of transcription units. Expression of miRNA transcripts is driven from promoter regions regulated by canonical transcription factors and epigenetic mechanisms. Several miRNAs, which are transcribed from physically adjacent miRNA genes, form a miRNA cluster [181].
miRNAs are transcribed as long hairpin molecules (pri-miRNAs) that are subsequently processed by canonical or non-canonical pathways of miRNA biogenesis. Most frequently, pri-miRNAs are cleaved to approximately 70-nt long stem-loop precursors (pre-miRNAs). Following nuclear export, the pre-miRNA is further processed into ~22 nt long mature strands, which assemble with Argonaute family proteins to form RNA-induced silencing complex (RISC). Both cleaved single strands from the 5′ and 3′ arm of the precursor miRNA can form a RISC complex in varied proportions, creating mature miRNA complexes with -5p or -3p RNA strands [182].
The RISC complex mediates downregulation of target proteins through inhibition of translation, or mRNA degradation following its deadenylation and decapping [182]. Under specific conditions, the miRNA-target interaction can induce translational activation [183]. Activation of TLR receptors by double-stranded RNA was recently identified as another mechanism of miRNA action [184].
miRNAs interact with their target mRNA molecules via imperfect base pairing. Computational miRNA-target prediction algorithms search UTRs and CDS of putative target genes for sequences complementary to the miRNA seed region, an 8-nt stretch with near-perfect base pairing. Short miRNA sequences and imperfect base-pairing imply that every miRNA can regulate a plethora of different targets and vice versa [185]. Only a fraction of miRNA:target interactions have been experimentally validated [186].
Mechanisms of post-transcriptional regulation of miRNAs by lncRNAs and circular endogenous RNAs are gaining increased attention for their potential to regulate miRNA and/or target availability [187,188]. Lo et al. recently discovered a novel mechanism of how the miRNA turnover can be further regulated by the interferon response pathway, through IFIT5-XRN1-mediated degradation [189].

3.2. miRNAs in the Regulation of NED and Prostate Cancer Progression

miRNAs can be implicated at multiple levels of NED control. The phenotypic shift towards NED can be promoted either by downregulation of miRNAs directly targeting the transcripts associated with the neuroendocrine phenotype or more frequently by modulation of expression of upstream molecules regulating the NE transformation (positive or negative). Table 2 and Supplementary Table S2 summarize the effects of selected miRNAs that were described in the context of NED phenotype or in the control of NED regulators. The purpose of these tables is to assemble current knowledge concerning miRNA expression in prostate cancer patients, potential correlation with their prognosis, validated miRNA targets relevant for the NED phenotype, and clinical utility of specific miRNA expression for cancer diagnosis or treatment prediction. Less common mechanisms of expression control are also mentioned.
As NEPC is most often associated with advanced disease, resection of tumours at this stage is rare, as it does not bring therapeutic benefit. Therefore, sources of patient samples are very limited, and very few studies actually exploit clinical samples of prostate cancer with NEPC traits for next generation sequencing analysis of transcripts. Still, findings resulting from analysis of clinical specimens can be considered biologically more relevant than observations from tissue cultures, in which the combination of artificial culture conditions and cellular plasticity can result in NE-like transformation [291]. The implication of miRNAs in the context of treatment-induced NED was recently reviewed in [292].
Bhagirath et al. have assembled a cohort of eight NEPC tumour tissues and performed sequencing of small RNAs [194], while Beltran et al. profiled protein-coding transcripts of seven NEPC samples by NGS [121]. miRNAs enriched [293] for the genes reported in the latter dataset, which overlapped with miRNAs deregulated in NEPC samples [194], highlighted the overexpression of miR-375 and downregulation of miR-34a, miR-30c-5p, miR-363-3p, and miR-19b-3p (Supplementary Figure S1) (detailed information about the generation of miRNAs overlay is described in Supplementary Material). Notably, increased expression of miR-375 was detected also in the serum extracellular vesicles (EVs) of patients with NEPC [217]. miR-375, miR-34a and miR-19b-3p are validated regulators of molecules associated with the NED phenotype (Table 2).
Mechanistically, miRNAs can play dual roles in prostate carcinogenesis. Opposite findings of miRNA function in biological processes implicated in cancer progression (proliferation, migration, invasion, and apoptosis) are reported for most miRNAs (Table 2 and Supplementary Table S2). These discrepancies may result from non-physiological concentrations of experimentally introduced miRNAs or miRNA antagonists, as concentrations that are typically used in transfection experiments far exceed the total cellular concentration of the most highly expressed miRNAs [188]. Most importantly, certain miRNAs such as miR-204 may act as tumour suppressors in prostate adenocarcinomas but promote cancer in neuroendocrine tumours [3].
The following subchapters summarize the implication of miRNAs in the NED phenotype, which were described in distinct contexts with different degrees of physiological relevance (Figure 2).

3.3. miRNAs Associated with Neuroendocrine Prostate Cancer

The recent availability of a next-generation sequencing methodology has enabled in-depth studies of less common phenotypes in cancer such as NEPC. Transcriptome profiling of clinical specimens identified deregulated miRNAs in prostate tumours with a neuroendocrine phenotype compared to prostate adenocarcinomas. The relevance of these findings was confirmed by validation in available patient cohorts and functional in vitro experiments [194]. An alternative approach exploiting Ago-HITS-CLIP based identification of miRNA binding sites revealed a correlation between miR-194 and the NE phenotype, with subsequent validation in patient samples and in vitro models [190]. On the other hand, the miR-106b~25 cluster and let-7 were implicated in regulatory mechanisms associated with NED induction, and their expression correlates with clinical observations [128,260]. The following information about the role of individual miRNAs and miRNA clusters associated with validated NE phenotype in PCa is summarized in Table 2.

3.3.1. hsa-miR-194

miR-194 was found to be associated with plasticity of prostate cancer cells [191], and its elevated expression and activity was recently detected in NEPC [190]. Increased miR-194 inversely correlated with AR activity. Frequent gains and amplifications of miR-194, whose two copies are located on human chromosomes 1 and 11, were observed in NEPC datasets. Of the 160 putative miR-194 targets identified in the study, FOXA1 was identified as a target gene by which miR-194 influences the emergence of NEPC [190].
Upregulated miR-194 was discovered in prostatectomy specimens as well as in the circulation of relapsing patients and was associated with a higher Gleason score and poor prognosis [192,193]. By targeting SOCS2 and associated STAT3 and ERK signalling pathways, miR-194 was identified as a promoter of metastases in prostate cancer [191], although inhibition of cell motility and a negative effect on viability of cancer cells were also described [294,295].

3.3.2. hsa-miR-375

miR-375 was predominantly enriched in patient tissues with NEPC features, and its experimental overexpression induced expression of NE markers SYP, ENO2, and CHGA [194]. Of NED-associated targets, hsa-miR-375 targets TP53 and NCAM1 in gastric cancer cells and neurons, respectively [195,196]. miR-375 has lately been intensively studied as a prognostic factor, and its diagnostic potential was evaluated in combination with other miRNAs elevated in advanced prostate disease [208]. In urinary exosomes, miR-375 was decreasing with disease progression [296], while in serum EVs, miR-375 expression was enriched in patients with NEPC [217]. miR-375-based non-invasive screening of circulating miRNAs could distinguish benign and aggressive disease and predict treatment response [297].

3.3.3. hsa-miR-301

Similarly to miR-375, enrichment of miR-301 was detected in tumours with NEPC characteristics, and its experimental manipulation affected the NED phenotype [194]. Of NED-associated targets, miR-301 was validated to target PTEN in breast cancer [218]. Besides elevated expression in NEPC tissues, miR-301 was also increased in prostate cancer compared to adjacent tissue [220] and manifested various pro-tumorigenic effects [221,222,223,224]. In prostatectomy specimens, high levels of miR-301a were associated with higher risk of biochemical recurrence [223] and were proposed as a predictor of metastatic disease [225].

3.3.4. hsa-miR-106a~363 Cluster

Six miRNAs expressed from the cluster encoded on human chromosome X (miR-106a, miR-18b, miR-19b, miR-20b, miR-92a, and miR-363) were concomitantly downregulated in NEPC [194]. Experimental downregulation of these miRNAs induced expression of NED markers CHGA, ENO2, and SYP. STAT3, MYCN, and E2F1 were identified as direct targets of the miR-106a~363 cluster [194]. miRNAs of the hsa-miR-106a~363 cluster target TP53, RB1, and PTEN tumour suppressors and negative regulators of NED [229,230,231,241,249,250,254]. In experimental settings, miR-106a~363 cluster miRNAs exhibit oncogenic properties (Table 2) and are associated with poor patient prognosis [232,244].

3.3.5. hsa-miR-106b, miR-93, and miR-25 Cluster

Hsa-miR-106b was upregulated in NED arising in hypoxic conditions [260] and was significantly upregulated in multiple cohorts of patients with NEPC [121,194]. Experimentally validated targets of the miR-106b~25 cluster implicated in NED control comprise PTEN [265,268], TP53 [278], RB1 [261], and EZH2 [277]. Furthermore, the miR-106b~25 cluster downregulates the transcriptional repressor REST, which represses neuron-specific protein-coding and miRNA-coding genes. miRNAs encoded by this cluster were deregulated by hypoxia and in high grade PCa patients, and their experimental overexpression induced proneural genes in model cell lines [260].
In the rare prostatic small cell neuroendocrine carcinoma, the absence or mutation of p53 deregulated expression of miR-25 and the E3 ligase FBXW7 results in elevated levels of AURKA and enhancement of cancer cell proliferation and aggressive behaviour [276]. Amplifications of AURKA were strongly associated with the emergence of treatment-associated NED phenotype [298]. Other members of the miR-25 family, miR-92a and miR-92b, also interacted with AURKA based on NGS results, although additional evidence by a low throughput method is still lacking [299,300].

3.3.6. hsa-let-7

Members of the let-7 miRNA family were implicated in the development of NEPC in the context of LIN28B signalling as its negative downstream effectors [128]. Changes in expression of multiple let-7 family members in NE cells correspond with the general function of LIN28 proteins in miRNA biogenesis [301], while the downstream regulation of AR and Myc by the LIN28B-let-7 axis influences prostate cancer progression [290]. Specifically, let-7 family members regulate multiple molecules associated with the NED phenotype: ASCL1 [284], EZH2 [281,282], MYCN [283], and HMGA-2 [128].

3.4. miRNAs Associated with Neuroendocrine-like Changes in Prostate Cancer Models

Experimental modulation of miRNA expression was capable of inducing an NE phenotype in cell cultures derived from prostate cancer cells, as manifested by changes in cell morphology and the induction of NE markers (Table 1). With respect to potential off-target effects of non-physiological concentrations of exogenous miRNAs [188], deregulation of miRNAs described below was frequently observed in prostate cancer, but direct association with the NE phenotype was not investigated. Detailed information about expression, function, and diagnostic utility of the following miRNAs is summarized in Supplementary Table S2.

3.4.1. hsa-miR-17/92 Cluster

A well-studied miRNA cluster with oncogenic properties, encoded on human chromosome 13, encompasses six miRNAs: miR-17, miR-18a, miR-19a, miR-20a, miR-19b-1, and miR-92a-1 [302]. Upregulation of miRNAs from the miR-17/92 cluster, associated with SOX4 overexpression, was confirmed in patient sample and induced an NED phenotype in prostate cancer cells [303]. RB1 was identified as a target of miR-17/92, which is frequently absent in small cell neuroendocrine carcinoma of the prostate [304]. From other negative NED regulators, PTEN and TP53 were experimentally validated as targets of miR-17, miR-18a, miR-19a/b, miR-20a, and miR-92a [229,241,250,305,306,307,308,309,310]. On the other hand, positive regulators of NED, MYCN, and AKT1 are targeted by miR-19a and miR-19b [248,311,312].
Contrarily, downregulation of all six members of the miR-17 family, including hsa-miR-106b, was observed in experimental NED of LNCaP cells, which was associated with induction of Cyclin D1 [228]. In accordance with the Akt signalling pathway being an important regulator of NED [123], miR-17, -20b, and -106b negatively regulated AKT3, whose expression accompanied the NED phenotype in clinical samples of advanced prostate cancer [313].

3.4.2. hsa-miR-663

hsa-miR-663 was characterized as an oncogenic miRNA, capable of modulating expression of the NED marker NSE in LNCaP cells. miR-663 expression correlated with Gleason score and TNM stage and was suggested as an independent prognostic predictor of clinical recurrence [314]. miR-663 was shown to target the tumour suppressor genes encoding p53 and p21 [315]. miR-663 was also identified as one of 5 miRNAs overexpressed in metastatic PCa and responsible for STAT3 deregulation [316].

3.4.3. hsa-miR-32

The function of hsa-miR-32 in NED was discovered in the context of inhibited AR signalling, resulting from the interaction of prostate cancer cells with mast cells, or from anti-androgen therapy [317]. miR-32 expression was found to be regulated by AR transcriptional activity and increased in CRPC [318]. Elevated miR-32 expression was associated with pro-tumorigenic effects such as increased proliferation and transformation in vivo, chemoresistance, and radioresistance [319,320,321]. Of NED regulators, miR-32 was validated to target PTEN [322].

3.4.4. hsa-miR-204 and hsa-miR-34a

A negative feedback loop, encompassing AR-driven downregulation of miR-204, inhibition or XRN-1, and downregulation of miR-34a, which enables re-expression of AR, was described in prostate cancer cells [3]. Interestingly, the described signalling loop exhibited opposite effects in PCa and NED cells, with a pro-tumorigenic effect of upregulated miR-204 on cells with the NED phenotype. In accordance with the general tumour-suppressive role of miR-34a, the expression of both miR-204 and miR-34a was found to be downregulated in advanced PCa [3,323,324,325,326]. While the expression of miR-204 is downregulated by androgens [3], miR-34a itself was validated to target AR [327] as well as other regulators of the NED phenotype MYCN [328], SOX2 [329], and TP53 [330].

3.4.5. hsa-miR-221

Elevated expression of miR-221 was described in an androgen-independent subline of prostate cancer cell line LNCaP, and its experimental modulation affected the neuroendocrine phenotype of the cells, with concomitant effect on the Wnt signalling pathway through the regulation of DVL2 [331]. In prostate cancer and advanced disease, miR-221 expression was downregulated [199,332,333,334], but a significant upregulation was detected in clinical NEPC samples [194]. miR-221 was shown to affect the expression of NED regulators RB1 [335] and PTEN [336,337].

3.5. Additional miRNAs Implicated in the Modulation of Key Positive and Negative Regulators of NEPC

The following subchapter points out several miRNA candidates whose direct association with NED has not been experimentally demonstrated but which can regulate several positive or negative regulators of the NED phenotype: AR, MYCN, and AKT, alone or in combination with other NED-related targets.

3.5.1. miRNAs Implicated in the Regulation of AR

With regard to the crucial role of AR signalling in PCa pathogenesis, a high throughput miRNA screen was performed to identify potential miRNA regulators of AR expression and transcriptional activity [338]. Besides miR-301a and miR-34a described in Section 3.3.3 and Section 3.4.4, miR-30 family members were identified as AR regulators with binding sites in both UTR and coding regions of AR; loss of miR-30c-5p and miR-30d-5p expression correlated with advanced disease [338]. Importantly, miR-30d-5p in serum EVs was deregulated in both treatment-induced and de novo NEPC [217]. miR-31, whose binding side in the AR coding region was frequently mutated in cancer, suppressed tumour formation in experimental models [339]. miR-31 negatively correlated with AR expression in a transcriptome analysis of prostate cancer tissues [340]. In general, miRNAs that were suppressed in metastatic prostate cancer, including miR-31, strongly affected AR expression and transcriptional activity and their decrease was associated with worse biochemical recurrence-free survival [341]. Notably, miR-346, miR-361-3p, and miR-197 increased AR activity through a novel and anti-dogmatic mechanism of direct association with AR 3′UTR and transcript stabilisation [342].

3.5.2. miRNAs Implicated in the Regulation of AKT and MYCN

Out of 20 miRNAs experimentally validated to control the expression of AKT1 whose constitutional activation drives the emergence of the NEPC phenotype, several miRNAs were reported to be specifically implicated in prostate cancer. miR-644 was described to control several regulators of the NED phenotype in the context of CRPC, including AKT, MYC, and AR coregulators [343]. However, information about the expression or function of miR-644 in prostate cancer pathology is lacking. miR-373-3p and miR-409-3p were described as promoters of prostate tumorigenesis, migration, and invasion [344,345,346,347]. Moreover, AKT activity can be controlled indirectly such as by the mechanism involving miR-197 regulation of the VDAC1/AKT/beta-catenin pathway [348] or miR-101 control of the AKT pathway through RLIP76 [349].
AKT is a validated target of miR-27a [350], and reciprocally, upstream regulation of miR-27a expression by AR and MYC suggest a potential indirect involvement of this miRNA in NEPC [351,352,353,354]. Downregulation of miR-27a caused by aberrant AR signalling and PI3K/Akt signalling after ADT was proposed to promote the progression of castration-resistant prostate cancer [352]. Although reports of miR-27a expression in tumour samples are conflicting [352,354,355,356,357], high serum levels of miR-27a correlate with poor survival [356] and indicate presence of metastases [358].

3.5.3. miRNAs Implicated in the Regulation of MYCN

From NED-associated miRNAs, MYCN is targeted by the above-described hsa-miR-19b, hsa-miR-let-7, and hsa-miR-34a. Hsa-miR-101 directly targets two positive regulators of NED phenotype, MYCN and EZH2 [248,359,360]. In general, miR-101 expression is often downregulated in primary and metastatic tumours [355,361,362] and, when restored, it exerts anti-cancer effects [363,364]. Its diagnostic utility for metastatic disease was proposed by several studies [253,365,366]. The expression of miR-101 can be modulated by several alternative mechanisms such as lncRNA CRNDE [367], biogenesis control by IFIT5 [189], or androgen stimulation [368].

3.6. LncRNAs Implicated in NEPC

Recently, next-generation sequencing data reveal novel potential RNA regulators of NED from the group of long non-coding RNAs [369]. Upregulation of lncRNA-p21 was detected in NEPC and enhanced by Enzalutamide treatment via EZH2/STAT3 signalling [370]. Differential expression of lncRNA-p21 distinguished PCa patients from BPH [371]. A similar NEPC-promoting effect was also described for lncRNA-PCAT6 by sponging miR-326 [372]. lncRNA-PCAT6 was associated with prostate cancer metastases [373]. Upregulation of LINC00261 in NEPC was discovered in patient-derived xenografts and confirmed by analysis of multiple patient cohorts [5]. Opposite trend in expression between prostate cancer and healthy tissue suggests that upregulation of LINC00261 can be specific for NEPC [374].
LncRNAs HOTAIR and MALAT1 were increased in PCa samples with neuroendocrine characteristics [340]. HOTAIR was found upregulated in CRPC and its experimental modulation regulated the NED phenotype in prostate cancer cells [375], although its function in the clinical setting was questioned [376]. Nevertheless, HOTAIR may be implicated in NED based on its ability to regulate AR degradation [377]. MALAT1 was identified as one of the most abundant transcripts in CRPC biopsies [378], and high expression of MALAT1 was proposed to stratify patients with advanced PCa who would benefit from Enzalutamide treatment [379]. Altogether, lncRNAs emerge as a novel class of RNA regulators of NEPC interacting with already known transcription factors and miRNAs, with possible diagnostic and clinical utility.

4. Clinical Significance of Non-Coding RNAs as Biomarkers and Therapeutic Targets in NEPC

With increasing incidence of prostate cancer, efforts to improve diagnostic and prognostic methods for patient benefit include the investigation of miRNA expression in cancer tissues and body fluids, with the scope of clinical application. The advantage of miRNAs for cancer screening is their relative abundance and stability in body fluids and straightforward quantification by PCR-based methods or alternative techniques [380]. Multiple studies detect differences in miRNA expression in biopsies or body fluids of cancer patients to stratify healthy individuals, benign prostate hyperplasia, and prostate cancer. Non-invasive collection of liquid biomarkers from blood (plasma or serum), urine, or seminal plasma can serve for prediction of disease prognosis, metastatic dissemination, and treatment outcome and to stratify patients who would benefit from therapy of advanced prostate cancer. Computational algorithms and panels of several miRNAs were designed to increase the prediction power [211,236]. Several promising candidate biomarkers of prostate cancer were proposed based on analysis of samples from circulating blood (miR-25-3p and miR-18b-5p), urine (miR-95, miR-21, miR-19a, and miR-19b), and prostatic secretions (miR-203) [381].
miRNAs can be detected both in cell-free preparations and in exosomes/extracellular vesicles (EVs), along with lncRNAs harbouring seed regions of miRNAs implicated in NEPC regulation such as let-7 family members as well as miR-17, miR-18a, miR-20a, miR-93, and miR-106b [382]. Selective excretion of miRNAs in EVs underlies the observed negative correlation between miRNA content in EVs and tumour cells. For example, the cluster comprising miR-92a was downregulated in NEPC tumours, while miR-92a iso-miRs were significantly enriched in EVs obtained from plasma of NEPC patients [154,217].
Table 2 and Supplementary Table S2 include information on possible biomarker properties of NEPC-associated miRNAs. For most miRNAs associated with the NED phenotype, a certain correlation with prostate cancer was identified, and, in most datasets, miRNAs and their combinations were characterized as better biomarkers than PSA alone. Nevertheless, only very limited information is available about miRNA implication in the diagnosis of NEPC. A recent study by Bhagirath et al. identified an EV-microRNA classifier in serum of six CRPC patients with NEPC features comprising miRs-9-3p, -28-5p, -378d, -592, and -155-5p [217]. These molecules can bring increased specificity to the diagnosis of NEPC along with the detection of MYCN and AURKA transcripts [298].

5. Conclusions and Future Perspectives

Despite many tools for experimental manipulation of miRNA expression, therapeutic application of miRNAs or miRNA antagonists remains challenging. Current clinical trials in prostate cancer involving miRNAs focus on diagnosis, disease monitoring, and prediction of treatment outcome, but therapeutic application of miRNA in advanced prostate cancer treatment is still far from implementation. Questions of drug dosage, pharmacokinetics, and delivery need to be properly addressed [383]. Besides lipidic or polymeric nanoparticle delivery of selected miRNA(s), an alternative mechanism of silencing three members of the oncogenic miR-17~92 cluster was discovered, whereby a small molecule interfering with the Dicer processing site leads to impaired miRNA biogenesis [384], and conjugation of a chemotherapeutic drug with this inhibitor represents a powerful targeting strategy of the entire oncogenic pri-miR-17-92 cluster [385].
Altogether, the current clinical utility of NEPC-associated non-coding RNAs focuses on disease diagnosis, monitoring, and prediction of treatment outcome. Nevertheless, with improved methods of delivery of miRNA-based therapeutics leading to their increased tolerability, NEPC-associated miRNAs may serve as good candidates to slow down the progression to advanced disease in prostate cancer patients.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ncrna7040075/s1, Figure S1: Candidate miRNAs involved in prostate cancer neuroendocrine differentiation, Table S1: Regulators and markers of neuroendocrine differentiation in prostate cancer, Table S2: Cancer-related effects of NED-associated non-coding RNAs. References [386,387,388,389,390,391,392,393,394,395,396,397,398,399,400,401,402,403,404,405,406,407,408,409,410,411,412,413,414,415,416,417,418,419,420,421,422,423,424,425,426,427,428,429,430,431,432,433,434,435,436,437,438,439,440,441,442,443,444,445,446,447,448,449,450,451,452,453,454,455,456,457,458,459,460,461,462,463,464,465,466,467,468,469,470,471,472,473,474,475,476,477,478,479,480,481,482,483] are refered to in Supplementary Materials.

Author Contributions

conceptualization and writing—original draft preparation, E.S., Z.K., J.P. and K.S.; writing—review and editing, K.S.; visualization, J.P.; supervision, K.S.; project administration and funding acquisition, K.S. All authors have read and agreed to the published version of the manuscript. All authors contributed substantially to the work reported.

Funding

This work was supported by the Ministry of Health of the Czech Republic, grant nr. 17-28518A, all rights reserved, and European Structural and Investment Funds, Operational Programme Research, Development and Education—Preclinical Progression of New Organic Compounds with Targeted Biological Activity” (Preclinprogress)—CZ.02.1.01/0.0/0.0/16_025/0007381.

Data Availability Statement

Openly available datasets have been analyzed and cited in this study—Supplementary Table S3 in [121] at 10.1158/2159-8290.CD-11-0130. Supplementary Table S2 in [194] at/10.1038/s41388-020-01493-8.

Acknowledgments

The authors would like to thank Iva Lišková, Martina Urbánková, and Kateřina Svobodová for technical assistance.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

ADTandrogen-deprivation therapy
Ago-HITS-CLIPhigh-throughput sequencing of RNA isolated by cross-linking immunoprecipitation of Argonaute protein
AIandrogen independence
AKT1
AR
AKT Serine/threonine Kinase 1
androgen receptor
ARPIs
ASCL1
AURKA
AR pathway inhibitors
Achaete-Scute Family BHLH Transcription Factor 1
Aurora kinase A
BCRbiochemical recurrence
BPH
BRN2
BRN4
CgA (CHGA)
benign prostatic hyperplasia
POU Class 3 Homeobox 2
POU Class 3 Homeobox 4
chromogranin A
CRPCcastration-resistant prostate cancer
CTCcirculating tumour cell
DCX
E2F1
EHF
EMT
Doublecortin
E2F Transcription Factor 1
ETS Homologous Factor
epithelial-to-mesenchymal transition
EVsextracellular vesicles
EZH2
FOXA1
FOXA2
HOTAIR
HRPCa
Enhancer Of Zeste 2 Polycomb Repressive Complex 2 Subunit
Forkhead box A1
Forkhead box A2
HOX transcript antisense RNA
hormone refractory prostate cancer
LINC00261
lncRNA
Long Intergenic Non-Protein Coding RNA 261
long non-coding RNA
lncRNA-PCAT6
m-CRPC
long non-coding RNA prostate cancer-associated transcript 6
metastatic castration-resistant prostate cancer
MALAT1
miRNA
Metastasis Associated Lung Adenocarcinoma Transcript 1
microRNA
MUC1-C
N-MYC (MYCN)
NE cells
Mucin 1 C-terminal part
Neuroblastoma MYC oncogene
neuroendocrine cells
NE-like cancer cellsneuroendocrine-like cancer cells
NEDneuroendocrine differentiation
NEPCneuroendocrine prostate cancer
NSE (ENO2)neuron-specific enolase
ONECUT2
PARP1
PCa
One cut homeobox 2
Poly (ADP-ribose) polymerase 1
prostate cancer
PEG10
PSA
Paternally Expressed 10
prostate specific antigen
PTEN
RB1
REST
SCC
Phosphatase and tensin homolog
RB Transcriptional Corepressor 1
RE1 Silencing Transcription Factor
small cell carcinoma
SOX2
SRRM4
STAT3
Syp
Sex-determining region Y2
Serine/Arginine Repetitive Matrix 4
Signal Transducer And Activator Of Transcription 3
synaptophysin
t-NEPC
TF
Trop-2
treatment-emergent neuroendocrine prostate cancer
transcription factor
Tumor Associated Calcium Signal Transducer 2

References

  1. Patel, G.K.; Chugh, N.; Tripathi, M. Neuroendocrine Differentiation of Prostate Cancer—An Intriguing Example of Tumor Evolution at Play. Cancers 2019, 11, 1405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Vanacore, D.; Boccellino, M.; Rossetti, S.; Cavaliere, C.; D’Aniello, C.; Di Franco, R.; Romano, F.J.; Montanari, M.; La Mantia, E.; Piscitelli, R.; et al. Micrornas in prostate cancer: An overview. Oncotarget 2017, 8, 50240–50251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ding, M.; Lin, B.; Li, T.; Liu, Y.; Li, Y.; Zhou, X.; Miao, M.; Gu, J.; Pan, H.; Yang, F.; et al. A dual yet opposite growth-regulating function of miR-204 and its target XRN1 in prostate adenocarcinoma cells and neuroendocrine-like prostate cancer cells. Oncotarget 2015, 6, 7686–7700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhang, M.; Gao, F.; Yu, X.; Zhang, Q.; Sun, Z.; He, Y.; Guo, W. LINC00261: A burgeoning long noncoding RNA related to cancer. Cancer Cell Int. 2021, 21, 274. [Google Scholar] [CrossRef]
  5. Mather, R.L.; Parolia, A.; Carson, S.E.; Venalainen, E.; Roig-Carles, D.; Jaber, M.; Chu, S.C.; Alborelli, I.; Wu, R.; Lin, D.; et al. The evolutionarily conserved long non-coding RNA LINC00261 drives neuroendocrine prostate cancer proliferation and metastasis via distinct nuclear and cytoplasmic mechanisms. Mol. Oncol. 2021, 15, 1921–1941. [Google Scholar] [CrossRef]
  6. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  7. Parker, C.; Castro, E.; Fizazi, K.; Heidenreich, A.; Ost, P.; Procopio, G.; Tombal, B.; Gillessen, S. Prostate cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2020, 31, 1119–1134. [Google Scholar] [CrossRef]
  8. Watson, P.A.; Arora, V.K.; Sawyers, C.L. Emerging mechanisms of resistance to androgen receptor inhibitors in prostate cancer. Nat. Rev. Cancer 2015, 15, 701–711. [Google Scholar] [CrossRef] [Green Version]
  9. Nevedomskaya, E.; Baumgart, S.J.; Haendler, B. Recent Advances in Prostate Cancer Treatment and Drug Discovery. Int. J. Mol. Sci. 2018, 19, 1359. [Google Scholar] [CrossRef] [Green Version]
  10. Davies, A.H.; Beltran, H.; Zoubeidi, A. Cellular plasticity and the neuroendocrine phenotype in prostate cancer. Nat. Rev. Urol. 2018, 15, 271–286. [Google Scholar] [CrossRef]
  11. Clermont, P.-L.; Ci, X.; Pandha, H.; Wang, Y.; Crea, F. Treatment-emergent neuroendocrine prostate cancer: Molecularly driven clinical guidelines. Int. J. Endocr. Oncol. 2019, 6, IJE20. [Google Scholar] [CrossRef] [Green Version]
  12. Taylor, R.A.; Toivanen, R.; Risbridger, G.P. Stem cells in prostate cancer: Treating the root of the problem. Endocr.-Relat. Cancer 2010, 17, R273–R285. [Google Scholar] [CrossRef] [Green Version]
  13. Szczyrba, J.; Niesen, A.; Wagner, M.; Wandernoth, P.M.; Aumuller, G.; Wennemuth, G. Neuroendocrine Cells of the Prostate Derive from the Neural Crest. J. Biol. Chem. 2017, 292, 2021–2031. [Google Scholar] [CrossRef] [Green Version]
  14. Abrahamsson, P.A.; di Sant’Agnese, P.A. Neuroendocrine cells in the human prostate gland. J. Androl. 1993, 14, 307–309. [Google Scholar]
  15. Huang, J.; Yao, J.L.; di Sant’Agnese, P.A.; Yang, Q.; Bourne, P.A.; Na, Y. Immunohistochemical characterization of neuroendocrine cells in prostate cancer. Prostate 2006, 66, 1399–1406. [Google Scholar] [CrossRef]
  16. Abrahamsson, P.A. Neuroendocrine cells in tumour growth of the prostate. Endocr.-Relat. Cancer 1999, 6, 503–519. [Google Scholar] [CrossRef] [Green Version]
  17. Huang, Y.H.; Zhang, Y.Q.; Huang, J.T. Neuroendocrine cells of prostate cancer: Biologic functions and molecular mechanisms. Asian J. Androl. 2019, 21, 291–295. [Google Scholar] [CrossRef]
  18. Cindolo, L.; Cantile, M.; Vacherot, F.; Terry, S.; de la Taille, A. Neuroendocrine differentiation in prostate cancer: From lab to bedside. Urol. Int. 2007, 79, 287–296. [Google Scholar] [CrossRef]
  19. Li, Z.; Chen, C.J.; Wang, J.K.; Hsia, E.; Li, W.; Squires, J.; Sun, Y.; Huang, J. Neuroendocrine differentiation of prostate cancer. Asian J. Androl. 2013, 15, 328–332. [Google Scholar] [CrossRef] [Green Version]
  20. Ozawa, H.; Takata, K. The granin family—Its role in sorting and secretory granule formation. Cell Struct. Funct. 1995, 20, 415–420. [Google Scholar] [CrossRef] [Green Version]
  21. Bartolomucci, A.; Possenti, R.; Mahata, S.K.; Fischer-Colbrie, R.; Loh, Y.P.; Salton, S.R. The extended granin family: Structure, function, and biomedical implications. Endocr. Rev. 2011, 32, 755–797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Abrahamsson, P.A.; Falkmer, S.; Falt, K.; Grimelius, L. The course of neuroendocrine differentiation in prostatic carcinomas. An immunohistochemical study testing chromogranin A as an “endocrine marker”. Pathol. Res. Pr. 1989, 185, 373–380. [Google Scholar] [CrossRef]
  23. Berruti, A.; Mosca, A.; Tucci, M.; Terrone, C.; Torta, M.; Tarabuzzi, R.; Russo, L.; Cracco, C.; Bollito, E.; Scarpa, R.M.; et al. Independent prognostic role of circulating chromogranin A in prostate cancer patients with hormone-refractory disease. Endocr.-Relat. Cancer 2005, 12, 109–117. [Google Scholar] [CrossRef] [PubMed]
  24. Ather, M.H.; Abbas, F.; Faruqui, N.; Israr, M.; Pervez, S. Correlation of three immunohistochemically detected markers of neuroendocrine differentiation with clinical predictors of disease progression in prostate cancer. BMC Urol. 2008, 8, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Hong, P.; Guo, R.Q.; Song, G.; Yang, K.W.; Zhang, L.; Li, X.S.; Zhang, K.; Zhou, L.Q. Prognostic role of chromogranin A in castration-resistant prostate cancer: A meta-analysis. Asian J. Androl. 2018, 20, 561–566. [Google Scholar] [CrossRef] [PubMed]
  26. Pruneri, G.; Galli, S.; Rossi, R.S.; Roncalli, M.; Coggi, G.; Ferrari, A.; Simonato, A.; Siccardi, A.G.; Carboni, N.; Buffa, R. Chromogranin A and B and secretogranin II in prostatic adenocarcinomas: Neuroendocrine expression in patients untreated and treated with androgen deprivation therapy. Prostate 1998, 34, 113–120. [Google Scholar] [CrossRef]
  27. Yuan, T.C.; Veeramani, S.; Lin, F.F.; Kondrikou, D.; Zelivianski, S.; Igawa, T.; Karan, D.; Batra, S.K.; Lin, M.F. Androgen deprivation induces human prostate epithelial neuroendocrine differentiation of androgen-sensitive LNCaP cells. Endocr.-Relat. Cancer 2006, 13, 151–167. [Google Scholar] [CrossRef] [Green Version]
  28. Xu, C.M.; Luo, Y.L.; Li, S.; Li, Z.X.; Jiang, L.; Zhang, G.X.; Owusu, L.; Chen, H.L. Multifunctional neuron-specific enolase: Its role in lung diseases. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef] [Green Version]
  29. Muoio, B.; Pascale, M.; Roggero, E. The role of serum neuron-specific enolase in patients with prostate cancer: A systematic review of the recent literature. Int. J. Biol. Markers 2018, 33, 10–21. [Google Scholar] [CrossRef]
  30. Kamiya, N.; Akakura, K.; Suzuki, H.; Isshiki, S.; Komiya, A.; Ueda, T.; Ito, H. Pretreatment serum level of neuron specific enolase (NSE) as a prognostic factor in metastatic prostate cancer patients treated with endocrine therapy. Eur. Urol. 2003, 44, 309–314, discussion 314. [Google Scholar] [CrossRef]
  31. Szarvas, T.; Csizmarik, A.; Fazekas, T.; Huttl, A.; Nyirady, P.; Hadaschik, B.; Grunwald, V.; Pullen, L.; Juranyi, Z.; Kocsis, Z.; et al. Comprehensive analysis of serum chromogranin A and neuron-specific enolase levels in localized and castration-resistant prostate cancer. BJU Int. 2021, 127, 44–55. [Google Scholar] [CrossRef]
  32. El Far, O.; Betz, H. Synaptophysins: Vesicular cation channels? J. Physiol. 2002, 539, 332. [Google Scholar] [CrossRef]
  33. Sainio, M.; Visakorpi, T.; Tolonen, T.; Ilvesaro, J.; Bova, G.S. Expression of neuroendocrine differentiation markers in lethal metastatic castration-resistant prostate cancer. Pathol. Res. Pr. 2018, 214, 848–856. [Google Scholar] [CrossRef] [Green Version]
  34. Pal, S.K.; He, M.; Chen, L.; Yang, L.; Pillai, R.; Twardowski, P.; Hsu, J.; Kortylewski, M.; Jones, J.O. Synaptophysin expression on circulating tumor cells in patients with castration resistant prostate cancer undergoing treatment with abiraterone acetate or enzalutamide. Urol. Oncol. 2018, 36, 162.e161–162.e166. [Google Scholar] [CrossRef]
  35. Van Acker, H.H.; Capsomidis, A.; Smits, E.L.; Van Tendeloo, V.F. CD56 in the Immune System: More Than a Marker for Cytotoxicity? Front. Immunol. 2017, 8, 892. [Google Scholar] [CrossRef]
  36. Mlika, M.; Zendah, I.; Braham, E.; El Mezni, F. CD56 antibody: Old-fashioned or still trendy in endocrine lung tumors. J. Immunoass. Immunochem. 2015, 36, 414–419. [Google Scholar] [CrossRef]
  37. Lee, J.K.; Bangayan, N.J.; Chai, T.; Smith, B.A.; Pariva, T.E.; Yun, S.; Vashisht, A.; Zhang, Q.; Park, J.W.; Corey, E.; et al. Systemic surfaceome profiling identifies target antigens for immune-based therapy in subtypes of advanced prostate cancer. Proc. Natl. Acad. Sci. USA 2018, 115, E4473–E4482. [Google Scholar] [CrossRef] [Green Version]
  38. Bertoldi, M. Mammalian Dopa decarboxylase: Structure, catalytic activity and inhibition. Arch. Biochem. Biophys 2014, 546, 1–7. [Google Scholar] [CrossRef]
  39. Wafa, L.A.; Cheng, H.; Rao, M.A.; Nelson, C.C.; Cox, M.; Hirst, M.; Sadowski, I.; Rennie, P.S. Isolation and identification of L-dopa decarboxylase as a protein that binds to and enhances transcriptional activity of the androgen receptor using the repressed transactivator yeast two-hybrid system. Biochem. J. 2003, 375, 373–383. [Google Scholar] [CrossRef]
  40. Wafa, L.A.; Palmer, J.; Fazli, L.; Hurtado-Coll, A.; Bell, R.H.; Nelson, C.C.; Gleave, M.E.; Cox, M.E.; Rennie, P.S. Comprehensive expression analysis of L-dopa decarboxylase and established neuroendocrine markers in neoadjuvant hormone-treated versus varying Gleason grade prostate tumors. Hum. Pathol. 2007, 38, 161–170. [Google Scholar] [CrossRef]
  41. Margiotti, K.; Wafa, L.A.; Cheng, H.; Novelli, G.; Nelson, C.C.; Rennie, P.S. Androgen-regulated genes differentially modulated by the androgen receptor coactivator L-dopa decarboxylase in human prostate cancer cells. Mol. Cancer 2007, 6, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Mariani, M.; Karki, R.; Spennato, M.; Pandya, D.; He, S.; Andreoli, M.; Fiedler, P.; Ferlini, C. Class III beta-tubulin in normal and cancer tissues. Gene 2015, 563, 109–114. [Google Scholar] [CrossRef] [PubMed]
  43. Terry, S.; Ploussard, G.; Allory, Y.; Nicolaiew, N.; Boissiere-Michot, F.; Maille, P.; Kheuang, L.; Coppolani, E.; Ali, A.; Bibeau, F.; et al. Increased expression of class III beta-tubulin in castration-resistant human prostate cancer. Br. J. Cancer 2009, 101, 951–956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ploussard, G.; Terry, S.; Maille, P.; Allory, Y.; Sirab, N.; Kheuang, L.; Soyeux, P.; Nicolaiew, N.; Coppolani, E.; Paule, B.; et al. Class III beta-tubulin expression predicts prostate tumor aggressiveness and patient response to docetaxel-based chemotherapy. Cancer Res. 2010, 70, 9253–9264. [Google Scholar] [CrossRef] [Green Version]
  45. Thomas, R.P.; Hellmich, M.R.; Townsend, C.M., Jr.; Evers, B.M. Role of gastrointestinal hormones in the proliferation of normal and neoplastic tissues. Endocr. Rev. 2003, 24, 571–599. [Google Scholar] [CrossRef]
  46. Ischia, J.; Patel, O.; Bolton, D.; Shulkes, A.; Baldwin, G.S. Expression and function of gastrin-releasing peptide (GRP) in normal and cancerous urological tissues. BJU Int. 2014, 113 (Suppl. S2), 40–47. [Google Scholar] [CrossRef]
  47. Qiao, J.; Grabowska, M.M.; Forestier-Roman, I.S.; Mirosevich, J.; Case, T.C.; Chung, D.H.; Cates, J.M.; Matusik, R.J.; Manning, H.C.; Jin, R. Activation of GRP/GRP-R signaling contributes to castration-resistant prostate cancer progression. Oncotarget 2016, 7, 61955–61969. [Google Scholar] [CrossRef]
  48. Solorzano, S.R.; Imaz-Rosshandler, I.; Camacho-Arroyo, I.; Garcia-Tobilla, P.; Morales-Montor, G.; Salazar, P.; Arena-Ortiz, M.L.; Rodriguez-Dorantes, M. GABA promotes gastrin-releasing peptide secretion in NE/NE-like cells: Contribution to prostate cancer progression. Sci. Rep. 2018, 8, 10272. [Google Scholar] [CrossRef] [Green Version]
  49. Li, X.; Cai, H.; Wu, X.; Li, L.; Wu, H.; Tian, R. New Frontiers in Molecular Imaging Using Peptide-Based Radiopharmaceuticals for Prostate Cancer. Front. Chem. 2020, 8, 583309. [Google Scholar] [CrossRef]
  50. Russell, F.A.; King, R.; Smillie, S.J.; Kodji, X.; Brain, S.D. Calcitonin gene-related peptide: Physiology and pathophysiology. Physiol. Rev. 2014, 94, 1099–1142. [Google Scholar] [CrossRef] [Green Version]
  51. Warrington, J.I.; Richards, G.O.; Wang, N. The Role of the Calcitonin Peptide Family in Prostate Cancer and Bone Metastasis. Curr. Mol. Biol. Rep. 2017, 3, 197–203. [Google Scholar] [CrossRef] [Green Version]
  52. Suzuki, K.; Kobayashi, Y.; Morita, T. Significance of serum calcitonin gene-related peptide levels in prostate cancer patients receiving hormonal therapy. Urol. Int. 2009, 82, 291–295. [Google Scholar] [CrossRef]
  53. Nagakawa, O.; Ogasawara, M.; Fujii, H.; Murakami, K.; Murata, J.; Fuse, H.; Saiki, I. Effect of prostatic neuropeptides on invasion and migration of PC-3 prostate cancer cells. Cancer Lett. 1998, 133, 27–33. [Google Scholar] [CrossRef]
  54. Martinez, A.; Zudaire, E.; Portal-Nunez, S.; Guedez, L.; Libutti, S.K.; Stetler-Stevenson, W.G.; Cuttitta, F. Proadrenomedullin NH2-terminal 20 peptide is a potent angiogenic factor, and its inhibition results in reduction of tumor growth. Cancer Res. 2004, 64, 6489–6494. [Google Scholar] [CrossRef] [Green Version]
  55. Jimenez, N.; Calvo, A.; Martinez, A.; Rosell, D.; Cuttitta, F.; Montuenga, L.M. Expression of adrenomedullin and proadrenomedullin N-terminal 20 peptide in human and rat prostate. J. Histochem. Cytochem. 1999, 47, 1167–1178. [Google Scholar] [CrossRef] [Green Version]
  56. Calvo, A.; Abasolo, I.; Jimenez, N.; Wang, Z.; Montuenga, L. Adrenomedullin and proadrenomedullin N-terminal 20 peptide in the normal prostate and in prostate carcinoma. Microsc. Res. Tech. 2002, 57, 98–104. [Google Scholar] [CrossRef]
  57. Larrayoz, I.M.; Martinez-Herrero, S.; Garcia-Sanmartin, J.; Ochoa-Callejero, L.; Martinez, A. Adrenomedullin and tumour microenvironment. J. Transl. Med. 2014, 12, 339. [Google Scholar] [CrossRef] [Green Version]
  58. Rocchi, P.; Boudouresque, F.; Zamora, A.J.; Muracciole, X.; Lechevallier, E.; Martin, P.M.; Ouafik, L. Expression of adrenomedullin and peptide amidation activity in human prostate cancer and in human prostate cancer cell lines. Cancer Res. 2001, 61, 1196–1206. [Google Scholar]
  59. Berenguer, C.; Boudouresque, F.; Dussert, C.; Daniel, L.; Muracciole, X.; Grino, M.; Rossi, D.; Mabrouk, K.; Figarella-Branger, D.; Martin, P.M.; et al. Adrenomedullin, an autocrine/paracrine factor induced by androgen withdrawal, stimulates ‘neuroendocrine phenotype’ in LNCaP prostate tumor cells. Oncogene 2008, 27, 506–518. [Google Scholar] [CrossRef] [Green Version]
  60. Maj, M.; Wagner, L.; Tretter, V. 20 Years of Secretagogin: Exocytosis and Beyond. Front. Mol. Neurosci. 2019, 12, 29. [Google Scholar] [CrossRef] [Green Version]
  61. Adolf, K.; Wagner, L.; Bergh, A.; Stattin, P.; Ottosen, P.; Borre, M.; Birkenkamp-Demtroder, K.; Orntoft, T.F.; Torring, N. Secretagogin is a new neuroendocrine marker in the human prostate. Prostate 2007, 67, 472–484. [Google Scholar] [CrossRef] [PubMed]
  62. Naafs, M. Parathyroid Hormone Related Peptide (PTHrP): A Mini-Review. Endocrinol. Metab. Int. J. 2017, 5, 1–9. [Google Scholar] [CrossRef] [Green Version]
  63. Asadi, F.; Farraj, M.; Sharifi, R.; Malakouti, S.; Antar, S.; Kukreja, S. Enhanced expression of parathyroid hormone-related protein in prostate cancer as compared with benign prostatic hyperplasia. Hum. Pathol. 1996, 27, 1319–1323. [Google Scholar] [CrossRef]
  64. DaSilva, J.; Gioeli, D.; Weber, M.J.; Parsons, S.J. The neuroendocrine-derived peptide parathyroid hormone-related protein promotes prostate cancer cell growth by stabilizing the androgen receptor. Cancer Res. 2009, 69, 7402–7411. [Google Scholar] [CrossRef] [Green Version]
  65. Cui, Y.; Sun, Y.; Hu, S.; Luo, J.; Li, L.; Li, X.; Yeh, S.; Jin, J.; Chang, C. Neuroendocrine prostate cancer (NEPCa) increased the neighboring PCa chemoresistance via altering the PTHrP/p38/Hsp27/androgen receptor (AR)/p21 signals. Oncogene 2016, 35, 6065–6076. [Google Scholar] [CrossRef] [Green Version]
  66. Park, S.I.; Lee, C.; Sadler, W.D.; Koh, A.J.; Jones, J.; Seo, J.W.; Soki, F.N.; Cho, S.W.; Daignault, S.D.; McCauley, L.K. Parathyroid hormone-related protein drives a CD11b+Gr1+ cell-mediated positive feedback loop to support prostate cancer growth. Cancer Res. 2013, 73, 6574–6583. [Google Scholar] [CrossRef] [Green Version]
  67. Ongkeko, W.M.; Burton, D.; Kiang, A.; Abhold, E.; Kuo, S.Z.; Rahimy, E.; Yang, M.; Hoffman, R.M.; Wang-Rodriguez, J.; Deftos, L.J. Parathyroid hormone related-protein promotes epithelial-to-mesenchymal transition in prostate cancer. PLoS ONE 2014, 9, e85803. [Google Scholar] [CrossRef] [Green Version]
  68. Tyler-McMahon, B.M.; Boules, M.; Richelson, E. Neurotensin: Peptide for the next millennium. Regul. Pept. 2000, 93, 125–136. [Google Scholar] [CrossRef]
  69. Sehgal, I.; Powers, S.; Huntley, B.; Powis, G.; Pittelkow, M.; Maihle, N.J. Neurotensin is an autocrine trophic factor stimulated by androgen withdrawal in human prostate cancer. Proc. Natl. Acad. Sci. USA 1994, 91, 4673–4677. [Google Scholar] [CrossRef] [Green Version]
  70. Vias, M.; Burtt, G.; Culig, Z.; Veerakumarasivam, A.; Neal, D.E.; Mills, I.G. A role for neurotensin in bicalutamide resistant prostate cancer cells. Prostate 2007, 67, 190–202. [Google Scholar] [CrossRef]
  71. Zhu, S.; Tian, H.; Niu, X.; Wang, J.; Li, X.; Jiang, N.; Wen, S.; Chen, X.; Ren, S.; Xu, C.; et al. Neurotensin and its receptors mediate neuroendocrine transdifferentiation in prostate cancer. Oncogene 2019, 38, 4875–4884. [Google Scholar] [CrossRef] [Green Version]
  72. He, T.; Wang, M.; Wang, H.; Tan, H.; Tang, Y.; Smith, E.; Wu, Z.; Liao, W.; Hu, S.; Li, Z. Evaluation of neurotensin receptor 1 as potential biomarker for prostate cancer theranostic use. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 2199–2207. [Google Scholar] [CrossRef]
  73. Morgat, C.; Chastel, A.; Molinie, V.; Schollhammer, R.; Macgrogan, G.; Velasco, V.; Malavaud, B.; Fernandez, P.; Hindie, E. Neurotensin Receptor-1 Expression in Human Prostate Cancer: A Pilot Study on Primary Tumors and Lymph Node Metastases. Int. J. Mol. Sci. 2019, 20, 1721. [Google Scholar] [CrossRef] [Green Version]
  74. Apte, R.S.; Chen, D.S.; Ferrara, N. VEGF in S.Signaling and Disease: Beyond Discovery and Development. Cell 2019, 176, 1248–1264. [Google Scholar] [CrossRef] [Green Version]
  75. Harper, M.E.; Glynne-Jones, E.; Goddard, L.; Thurston, V.J.; Griffiths, K. Vascular endothelial growth factor (VEGF) expression in prostatic tumours and its relationship to neuroendocrine cells. Br. J. Cancer 1996, 74, 910–916. [Google Scholar] [CrossRef] [Green Version]
  76. Grobholz, R.; Bohrer, M.H.; Siegsmund, M.; Jünemann, K.-P.; Bleyl, U.; Woenckhaus, M. Correlation between neovascularisation and neuroendocrine differentiation in prostatic carcinoma. Pathol.-Res. Pract. 2000, 196, 277–284. [Google Scholar] [CrossRef]
  77. Shariat, S.F.; Anwuri, V.A.; Lamb, D.J.; Shah, N.V.; Wheeler, T.M.; Slawin, K.M. Association of Preoperative Plasma Levels of Vascular Endothelial Growth Factor and Soluble Vascular Cell Adhesion Molecule-1 With Lymph Node Status and Biochemical Progression After Radical Prostatectomy. J. Clin. Oncol. 2004, 22, 1655–1663. [Google Scholar] [CrossRef]
  78. Polge, A.; Gaspard, C.; Mottet, N.; Guitton, C.; Boyer, J.C.; Choquet, A.; Combettes, S.; Bancel, E.; Costa, P.; Bali, J.P. Neurohormonal stimulation of histamine release from neuroendocrine cells of the human adenomatous prostate. Prostate 1998, 34, 1–9. [Google Scholar] [CrossRef]
  79. Chen, J.; Hu, X.Y. Inhibition of histamine receptor H3R suppresses prostate cancer growth, invasion and increases apoptosis via the AR pathway. Oncol. Lett. 2018, 16, 4921–4928. [Google Scholar] [CrossRef] [Green Version]
  80. Dizeyi, N.; Hedlund, P.; Bjartell, A.; Tinzl, M.; Austild-Tasken, K.; Abrahamsson, P.A. Serotonin activates MAP kinase and PI3K/Akt signaling pathways in prostate cancer cell lines. Urol. Oncol. 2011, 29, 436–445. [Google Scholar] [CrossRef]
  81. Shinka, T.; Onodera, D.; Tanaka, T.; Shoji, N.; Miyazaki, T.; Moriuchi, T.; Fukumoto, T. Serotonin synthesis and metabolism-related molecules in a human prostate cancer cell line. Oncol. Lett. 2011, 2, 211–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Oikawa, S.; Inuzuka, C.; Kuroki, M.; Matsuoka, Y.; Kosaki, G.; Nakazato, H. Cell adhesion activity of non-specific cross-reacting antigen (NCA) and carcinoembryonic antigen (CEA) expressed on CHO cell surface: Homophilic and heterophilic adhesion. Biochem. Biophys Res. Commun. 1989, 164, 39–45. [Google Scholar] [CrossRef]
  83. Taheri, M.; Saragovi, U.; Fuks, A.; Makkerh, J.; Mort, J.; Stanners, C.P. Self recognition in the Ig superfamily. Identification of precise subdomains in carcinoembryonic antigen required for intercellular adhesion. J. Biol. Chem. 2000, 275, 26935–26943. [Google Scholar] [CrossRef]
  84. Thompson, J.A.; Grunert, F.; Zimmermann, W. Carcinoembryonic antigen gene family: Molecular biology and clinical perspectives. J. Clin. Lab. Anal. 1991, 5, 344–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. DeLucia, D.C.; Cardillo, T.M.; Ang, L.; Labrecque, M.P.; Zhang, A.; Hopkins, J.E.; De Sarkar, N.; Coleman, I.; da Costa, R.M.G.; Corey, E.; et al. Regulation of CEACAM5 and Therapeutic Efficacy of an Anti-CEACAM5-SN38 Antibody-drug Conjugate in Neuroendocrine Prostate Cancer. Clin. Cancer Res. 2021, 27, 759–774. [Google Scholar] [CrossRef] [PubMed]
  86. Aloe, L.; Rocco, M.L.; Bianchi, P.; Manni, L. Nerve growth factor: From the early discoveries to the potential clinical use. J. Transl. Med. 2012, 10, 239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Di Donato, M.; Cernera, G.; Migliaccio, A.; Castoria, G. Nerve Growth Factor Induces Proliferation and Aggressiveness In Prostate Cancer Cells. Cancers 2019, 11, 784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Di Donato, M.; Cernera, G.; Auricchio, F.; Migliaccio, A.; Castoria, G. Cross-talk between androgen receptor and nerve growth factor receptor in prostate cancer cells: Implications for a new therapeutic approach. Cell Death Discov. 2018, 4, 5. [Google Scholar] [CrossRef]
  89. Chen, W.Y.; Wen, Y.C.; Lin, S.R.; Yeh, H.L.; Jiang, K.C.; Chen, W.H.; Lin, Y.S.; Zhang, Q.; Liew, P.L.; Hsiao, M.; et al. Nerve growth factor interacts with CHRM4 and promotes neuroendocrine differentiation of prostate cancer and castration resistance. Commun. Biol. 2021, 4, 22. [Google Scholar] [CrossRef]
  90. Tilan, J.; Kitlinska, J. Neuropeptide Y (NPY) in tumor growth and progression: Lessons learned from pediatric oncology. Neuropeptides 2016, 55, 55–66. [Google Scholar] [CrossRef] [Green Version]
  91. Alshalalfa, M.; Nguyen, P.L.; Beltran, H.; Chen, W.S.; Davicioni, E.; Zhao, S.G.; Rebbeck, T.R.; Schaeffer, E.M.; Lotan, T.L.; Feng, F.Y.; et al. Transcriptomic and Clinical Characterization of Neuropeptide Y Expression in Localized and Metastatic Prostate Cancer: Identification of Novel Prostate Cancer Subtype with Clinical Implications. Eur. Urol. Oncol. 2019, 2, 405–412. [Google Scholar] [CrossRef]
  92. Ding, Y.; Lee, M.; Gao, Y.; Bu, P.; Coarfa, C.; Miles, B.; Sreekumar, A.; Creighton, C.J.; Ayala, G. Neuropeptide Y nerve paracrine regulation of prostate cancer oncogenesis and therapy resistance. Prostate 2021, 81, 58–71. [Google Scholar] [CrossRef]
  93. Nadal, R.; Schweizer, M.; Kryvenko, O.N.; Epstein, J.I.; Eisenberger, M.A. Small cell carcinoma of the prostate. Nat. Rev. Urol. 2014, 11, 213–219. [Google Scholar] [CrossRef]
  94. Wang, W.; Epstein, J.I. Small cell carcinoma of the prostate. A morphologic and immunohistochemical study of 95 cases. Am. J. Surg. Pathol. 2008, 32, 65–71. [Google Scholar] [CrossRef]
  95. Beltran, H.; Tagawa, S.T.; Park, K.; MacDonald, T.; Milowsky, M.I.; Mosquera, J.M.; Rubin, M.A.; Nanus, D.M. Challenges in recognizing treatment-related neuroendocrine prostate cancer. J. Clin. Oncol. 2012, 30, e386–e389. [Google Scholar] [CrossRef]
  96. Beltran, H.; Hruszkewycz, A.; Scher, H.I.; Hildesheim, J.; Isaacs, J.; Yu, E.Y.; Kelly, K.; Lin, D.; Dicker, A.; Arnold, J.; et al. The Role of Lineage Plasticity in Prostate Cancer Therapy Resistance. Clin. Cancer Res. 2019, 25, 6916–6924. [Google Scholar] [CrossRef] [Green Version]
  97. Zou, M.; Toivanen, R.; Mitrofanova, A.; Floch, N.; Hayati, S.; Sun, Y.; Le Magnen, C.; Chester, D.; Mostaghel, E.A.; Califano, A.; et al. Transdifferentiation as a Mechanism of Treatment Resistance in a Mouse Model of Castration-Resistant Prostate Cancer. Cancer Discov. 2017, 7, 736–749. [Google Scholar] [CrossRef] [Green Version]
  98. Dong, B.; Miao, J.; Wang, Y.; Luo, W.; Ji, Z.; Lai, H.; Zhang, M.; Cheng, X.; Wang, J.; Fang, Y.; et al. Single-cell analysis supports a luminal-neuroendocrine transdifferentiation in human prostate cancer. Commun. Biol. 2020, 3, 778. [Google Scholar] [CrossRef]
  99. Nouri, M.; Caradec, J.; Lubik, A.A.; Li, N.; Hollier, B.G.; Takhar, M.; Altimirano-Dimas, M.; Chen, M.; Roshan-Moniri, M.; Butler, M.; et al. Therapy-induced developmental reprogramming of prostate cancer cells and acquired therapy resistance. Oncotarget 2017, 8, 18949–18967. [Google Scholar] [CrossRef] [Green Version]
  100. Gupta, K.; Gupta, S. Neuroendocrine differentiation in prostate cancer: Key epigenetic players. Transl. Cancer Res. 2017, 6, S104–S108. [Google Scholar] [CrossRef]
  101. Davies, A.; Zoubeidi, A.; Selth, L.A. The epigenetic and transcriptional landscape of neuroendocrine prostate cancer. Endocr.-Relat. Cancer 2020, 27, R35–R50. [Google Scholar] [CrossRef] [PubMed]
  102. Ge, R.; Wang, Z.; Montironi, R.; Jiang, Z.; Cheng, M.; Santoni, M.; Huang, K.; Massari, F.; Lu, X.; Cimadamore, A.; et al. Epigenetic modulations and lineage plasticity in advanced prostate cancer. Ann. Oncol. 2020, 31, 470–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Park, J.W.; Lee, J.K.; Sheu, K.M.; Wang, L.; Balanis, N.G.; Nguyen, K.; Smith, B.A.; Cheng, C.; Tsai, B.L.; Cheng, D.; et al. Reprogramming normal human epithelial tissues to a common, lethal neuroendocrine cancer lineage. Science 2018, 362, 91–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yamada, Y.; Beltran, H. Clinical and Biological Features of Neuroendocrine Prostate Cancer. Curr. Oncol. Rep. 2021, 23, 15. [Google Scholar] [CrossRef]
  105. Ayala, G.E.; Dai, H.; Powell, M.; Li, R.; Ding, Y.; Wheeler, T.M.; Shine, D.; Kadmon, D.; Thompson, T.; Miles, B.J.; et al. Cancer-related axonogenesis and neurogenesis in prostate cancer. Clin. Cancer Res. 2008, 14, 7593–7603. [Google Scholar] [CrossRef] [Green Version]
  106. Mauffrey, P.; Tchitchek, N.; Barroca, V.; Bemelmans, A.P.; Firlej, V.; Allory, Y.; Romeo, P.H.; Magnon, C. Progenitors from the central nervous system drive neurogenesis in cancer. Nature 2019, 569, 672–678. [Google Scholar] [CrossRef]
  107. Sejda, A.; Sigorski, D.; Gulczyński, J.; Wesołowski, W.; Kitlińska, J.; Iżycka-Świeszewska, E. Complexity of Neural Component of Tumor Microenvironment in Prostate Cancer. Pathobiology 2020, 87 (Suppl. S2), 87–99. [Google Scholar] [CrossRef]
  108. 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]
  109. Wu, C.T.; Altuwaijri, S.; Ricke, W.A.; Huang, S.P.; Yeh, S.; Zhang, C.; Niu, Y.; Tsai, M.Y.; Chang, C. Increased prostate cell proliferation and loss of cell differentiation in mice lacking prostate epithelial androgen receptor. Proc. Natl. Acad. Sci. USA 2007, 104, 12679–12684. [Google Scholar] [CrossRef] [Green Version]
  110. Schroeder, A.; Herrmann, A.; Cherryholmes, G.; Kowolik, C.; Buettner, R.; Pal, S.; Yu, H.; Muller-Newen, G.; Jove, R. Loss of androgen receptor expression promotes a stem-like cell phenotype in prostate cancer through STAT3 signaling. Cancer Res. 2014, 74, 1227–1237. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, X.; Coleman, I.M.; Brown, L.G.; True, L.D.; Kollath, L.; Lucas, J.M.; Lam, H.M.; Dumpit, R.; Corey, E.; Chery, L.; et al. SRRM4 Expression and the Loss of REST Activity May Promote the Emergence of the Neuroendocrine Phenotype in Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2015, 21, 4698–4708. [Google Scholar] [CrossRef] [Green Version]
  112. Jernberg, E.; Bergh, A.; Wikstrom, P. Clinical relevance of androgen receptor alterations in prostate cancer. Endocr. Connect. 2017, 6, R146–R161. [Google Scholar] [CrossRef] [Green Version]
  113. Fujita, K.; Nonomura, N. Role of Androgen Receptor in Prostate Cancer: A Review. World J. Mens. Health 2019, 37, 288–295. [Google Scholar] [CrossRef]
  114. Formaggio, N.; Rubin, M.A.; Theurillat, J.P. Loss and revival of androgen receptor signaling in advanced prostate cancer. Oncogene 2021, 40, 1205–1216. [Google Scholar] [CrossRef]
  115. Jamaspishvili, T.; Berman, D.M.; Ross, A.E.; Scher, H.I.; De Marzo, A.M.; Squire, J.A.; Lotan, T.L. Clinical implications of PTEN loss in prostate cancer. Nat. Rev. Urol. 2018, 15, 222–234. [Google Scholar] [CrossRef]
  116. Soundararajan, R.; Aparicio, A.M.; Logothetis, C.J.; Mani, S.A.; Maity, S.N. Function of Tumor Suppressors in Resistance to Antiandrogen Therapy and Luminal Epithelial Plasticity of Aggressive Variant Neuroendocrine Prostate Cancers. Front. Oncol. 2018, 8, 69. [Google Scholar] [CrossRef]
  117. Kallio, H.M.L.; Hieta, R.; Latonen, L.; Brofeldt, A.; Annala, M.; Kivinummi, K.; Tammela, T.L.; Nykter, M.; Isaacs, W.B.; Lilja, H.G.; et al. Constitutively active androgen receptor splice variants AR-V3, AR-V7 and AR-V9 are co-expressed in castration-resistant prostate cancer metastases. Br. J. Cancer 2018, 119, 347–356. [Google Scholar] [CrossRef] [Green Version]
  118. Dehm, S.M.; Tindall, D.J. Ligand-independent androgen receptor activity is activation function-2-independent and resistant to antiandrogens in androgen refractory prostate cancer cells. J. Biol. Chem. 2006, 281, 27882–27893. [Google Scholar] [CrossRef] [Green Version]
  119. Zhang, Y.; Coillie, S.V.; Fang, J.Y.; Xu, J. Gain of function of mutant p53: R282W on the peak? Oncogenesis 2016, 5, e196. [Google Scholar] [CrossRef] [Green Version]
  120. Ma, S.; McGuire, M.H.; Mangala, L.S.; Lee, S.; Stur, E.; Hu, W.; Bayraktar, E.; Villar-Prados, A.; Ivan, C.; Wu, S.Y.; et al. Gain-of-function p53 protein transferred via small extracellular vesicles promotes conversion of fibroblasts to a cancer-associated phenotype. Cell Rep. 2021, 34, 108726. [Google Scholar] [CrossRef]
  121. Beltran, H.; Rickman, D.S.; Park, K.; Chae, S.S.; Sboner, A.; MacDonald, T.Y.; Wang, Y.; Sheikh, K.L.; Terry, S.; Tagawa, S.T.; et al. Molecular characterization of neuroendocrine prostate cancer and identification of new drug targets. Cancer Discov. 2011, 1, 487–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Dardenne, E.; Beltran, H.; Benelli, M.; Gayvert, K.; Berger, A.; Puca, L.; Cyrta, J.; Sboner, A.; Noorzad, Z.; MacDonald, T.; et al. N-Myc Induces an EZH2-Mediated Transcriptional Program Driving Neuroendocrine Prostate Cancer. Cancer Cell 2016, 30, 563–577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wu, C.; Huang, J. Phosphatidylinositol 3-kinase-AKT-mammalian target of rapamycin pathway is essential for neuroendocrine differentiation of prostate cancer. J. Biol. Chem. 2007, 282, 3571–3583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Li, B.; Sun, A.; Youn, H.; Hong, Y.; Terranova, P.F.; Thrasher, J.B.; Xu, P.; Spencer, D. Conditional Akt activation promotes androgen-independent progression of prostate cancer. Carcinogenesis 2007, 28, 572–583. [Google Scholar] [CrossRef] [Green Version]
  125. Cangemi, R.; Mensah, A.; Albertini, V.; Jain, A.; Mello-Grand, M.; Chiorino, G.; Catapano, C.V.; Carbone, G.M. Reduced expression and tumor suppressor function of the ETS transcription factor ESE-3 in prostate cancer. Oncogene 2008, 27, 2877–2885. [Google Scholar] [CrossRef]
  126. Albino, D.; Longoni, N.; Curti, L.; Mello-Grand, M.; Pinton, S.; Civenni, G.; Thalmann, G.; D’Ambrosio, G.; Sarti, M.; Sessa, F.; et al. ESE3/EHF controls epithelial cell differentiation and its loss leads to prostate tumors with mesenchymal and stem-like features. Cancer Res. 2012, 72, 2889–2900. [Google Scholar] [CrossRef] [Green Version]
  127. Long, Z.; Deng, L.; Li, C.; He, Q.; He, Y.; Hu, X.; Cai, Y.; Gan, Y. Loss of EHF facilitates the development of treatment-induced neuroendocrine prostate cancer. Cell Death Dis. 2021, 12, 46. [Google Scholar] [CrossRef]
  128. Lovnicki, J.; Gan, Y.; Feng, T.; Li, Y.; Xie, N.; Ho, C.H.; Lee, A.R.; Chen, X.; Nappi, L.; Han, B.; et al. LIN28B promotes the development of neuroendocrine prostate cancer. J. Clin. Investig. 2020, 130, 5338–5348. [Google Scholar] [CrossRef]
  129. Albino, D.; Civenni, G.; Dallavalle, C.; Roos, M.; Jahns, H.; Curti, L.; Rossi, S.; Pinton, S.; D’Ambrosio, G.; Sessa, F.; et al. Activation of the Lin28/let-7 Axis by Loss of ESE3/EHF Promotes a Tumorigenic and Stem-like Phenotype in Prostate Cancer. Cancer Res. 2016, 76, 3629–3643. [Google Scholar] [CrossRef] [Green Version]
  130. Raj, B.; O’Hanlon, D.; Vessey, J.P.; Pan, Q.; Ray, D.; Buckley, N.J.; Miller, F.D.; Blencowe, B.J. Cross-regulation between an alternative splicing activator and a transcription repressor controls neurogenesis. Mol. Cell 2011, 43, 843–850. [Google Scholar] [CrossRef]
  131. Chang, Y.T.; Lin, T.P.; Campbell, M.; Pan, C.C.; Lee, S.H.; Lee, H.C.; Yang, M.H.; Kung, H.J.; Chang, P.C. REST is a crucial regulator for acquiring EMT-like and stemness phenotypes in hormone-refractory prostate cancer. Sci. Rep. 2017, 7, 42795. [Google Scholar] [CrossRef] [Green Version]
  132. Flores-Morales, A.; Bergmann, T.B.; Lavallee, C.; Batth, T.S.; Lin, D.; Lerdrup, M.; Friis, S.; Bartels, A.; Kristensen, G.; Krzyzanowska, A.; et al. Proteogenomic Characterization of Patient-Derived Xenografts Highlights the Role of REST in Neuroendocrine Differentiation of Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2019, 25, 595–608. [Google Scholar] [CrossRef] [Green Version]
  133. Yang, Y.A.; Yu, J. Current perspectives on FOXA1 regulation of androgen receptor signaling and prostate cancer. Genes Dis. 2015, 2, 144–151. [Google Scholar] [CrossRef] [Green Version]
  134. Yu, X.; Gupta, A.; Wang, Y.; Suzuki, K.; Mirosevich, J.; Orgebin-Crist, M.C.; Matusik, R.J. Foxa1 and Foxa2 interact with the androgen receptor to regulate prostate and epididymal genes differentially. Ann. N. Y. Acad. Sci. 2005, 1061, 77–93. [Google Scholar] [CrossRef]
  135. Pomerantz, M.M.; Li, F.; Takeda, D.Y.; Lenci, R.; Chonkar, A.; Chabot, M.; Cejas, P.; Vazquez, F.; Cook, J.; Shivdasani, R.A.; et al. The androgen receptor cistrome is extensively reprogrammed in human prostate tumorigenesis. Nat. Genet. 2015, 47, 1346–1351. [Google Scholar] [CrossRef]
  136. Jin, H.J.; Zhao, J.C.; Wu, L.; Kim, J.; Yu, J. Cooperativity and equilibrium with FOXA1 define the androgen receptor transcriptional program. Nat. Commun. 2014, 5, 3972. [Google Scholar] [CrossRef]
  137. Kim, J.; Jin, H.; Zhao, J.C.; Yang, Y.A.; Li, Y.; Yang, X.; Dong, X.; Yu, J. FOXA1 inhibits prostate cancer neuroendocrine differentiation. Oncogene 2017, 36, 4072–4080. [Google Scholar] [CrossRef] [Green Version]
  138. Amador-Arjona, A.; Cimadamore, F.; Huang, C.T.; Wright, R.; Lewis, S.; Gage, F.H.; Terskikh, A.V. SOX2 primes the epigenetic landscape in neural precursors enabling proper gene activation during hippocampal neurogenesis. Proc. Natl. Acad. Sci. USA 2015, 112, E1936–E1945. [Google Scholar] [CrossRef] [Green Version]
  139. Vasconcelos, F.F.; Castro, D.S. Transcriptional control of vertebrate neurogenesis by the proneural factor Ascl1. Front. Cell Neurosci. 2014, 8, 412. [Google Scholar] [CrossRef] [Green Version]
  140. Couillard-Despres, S.; Winner, B.; Schaubeck, S.; Aigner, R.; Vroemen, M.; Weidner, N.; Bogdahn, U.; Winkler, J.; Kuhn, H.G.; Aigner, L. Doublecortin expression levels in adult brain reflect neurogenesis. Eur. J. Neurosci. 2005, 21, 1–14. [Google Scholar] [CrossRef]
  141. Arenas, E. Foxa2: The rise and fall of dopamine neurons. Cell Stem. Cell 2008, 2, 110–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Hashizume, K.; Yamanaka, M.; Ueda, S. POU3F2 participates in cognitive function and adult hippocampal neurogenesis via mammalian-characteristic amino acid repeats. Genes Brain Behav. 2018, 17, 118–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Guo, J.; Cheng, X.; Zhang, L.; Wang, L.; Mao, Y.; Tian, G.; Xu, W.; Wu, Y.; Ma, Z.; Qin, J.; et al. Exploration of the Brn4-regulated genes enhancing adult hippocampal neurogenesis by RNA sequencing. J. Neurosci. Res. 2017, 95, 2071–2079. [Google Scholar] [CrossRef] [PubMed]
  144. Knoepfler, P.S.; Cheng, P.F.; Eisenman, R.N. N-myc is essential during neurogenesis for the rapid expansion of progenitor cell populations and the inhibition of neuronal differentiation. Genes Dev. 2002, 16, 2699–2712. [Google Scholar] [CrossRef] [Green Version]
  145. van der Raadt, J.; van Gestel, S.H.C.; Nadif Kasri, N.; Albers, C.A. ONECUT transcription factors induce neuronal characteristics and remodel chromatin accessibility. Nucleic Acids Res. 2019, 47, 5587–5602. [Google Scholar] [CrossRef]
  146. Mu, P.; Zhang, Z.; Benelli, M.; Karthaus, W.R.; Hoover, E.; Chen, C.C.; Wongvipat, J.; Ku, S.Y.; Gao, D.; Cao, Z.; et al. SOX2 promotes lineage plasticity and antiandrogen resistance in TP53- and RB1-deficient prostate cancer. Science 2017, 355, 84–88. [Google Scholar] [CrossRef] [Green Version]
  147. Rapa, I.; Ceppi, P.; Bollito, E.; Rosas, R.; Cappia, S.; Bacillo, E.; Porpiglia, F.; Berruti, A.; Papotti, M.; Volante, M. Human ASH1 expression in prostate cancer with neuroendocrine differentiation. Mod. Pathol. 2008, 21, 700–707. [Google Scholar] [CrossRef] [Green Version]
  148. Rapa, I.; Volante, M.; Migliore, C.; Farsetti, A.; Berruti, A.; Vittorio Scagliotti, G.; Giordano, S.; Papotti, M. Human ASH-1 promotes neuroendocrine differentiation in androgen deprivation conditions and interferes with androgen responsiveness in prostate cancer cells. Prostate 2013, 73, 1241–1249. [Google Scholar] [CrossRef]
  149. Fraser, J.A.; Sutton, J.E.; Tazayoni, S.; Bruce, I.; Poole, A.V. hASH1 nuclear localization persists in neuroendocrine transdifferentiated prostate cancer cells, even upon reintroduction of androgen. Sci. Rep. 2019, 9, 19076. [Google Scholar] [CrossRef] [Green Version]
  150. Tabrizi, S.; Alshalalfa, M.; Mahal, B.A.; Davicioni, E.; Liu, Y.; Mouw, K.W.; Feng, F.; Nguyen, P.L.; Muralidhar, V. Doublecortin Expression in Prostate Adenocarcinoma and Neuroendocrine Tumors. Int. J. Radiat. Oncol. Biol. Phys. 2020, 108, 936–940. [Google Scholar] [CrossRef]
  151. Park, J.W.; Lee, J.K.; Witte, O.N.; Huang, J. FOXA2 is a sensitive and specific marker for small cell neuroendocrine carcinoma of the prostate. Mod. Pathol. 2017, 30, 1262–1272. [Google Scholar] [CrossRef] [Green Version]
  152. Connelly, Z.M.; Yang, S.; Chen, F.; Yeh, Y.; Khater, N.; Jin, R.; Matusik, R.; Yu, X. Foxa2 activates the transcription of androgen receptor target genes in castrate resistant prostatic tumors. Am. J. Clin. Exp. Urol. 2018, 6, 172–181. [Google Scholar]
  153. Bishop, J.L.; Thaper, D.; Vahid, S.; Davies, A.; Ketola, K.; Kuruma, H.; Jama, R.; Nip, K.M.; Angeles, A.; Johnson, F.; et al. The Master Neural Transcription Factor BRN2 Is an Androgen Receptor-Suppressed Driver of Neuroendocrine Differentiation in Prostate Cancer. Cancer Discov. 2017, 7, 54–71. [Google Scholar] [CrossRef] [Green Version]
  154. Bhagirath, D.; Yang, T.L.; Tabatabai, Z.L.; Majid, S.; Dahiya, R.; Tanaka, Y.; Saini, S. BRN4 Is a Novel Driver of Neuroendocrine Differentiation in Castration-Resistant Prostate Cancer and Is Selectively Released in Extracellular Vesicles with BRN2. Clin. Cancer Res. 2019, 25, 6532–6545. [Google Scholar] [CrossRef] [Green Version]
  155. Lee, J.K.; Phillips, J.W.; Smith, B.A.; Park, J.W.; Stoyanova, T.; McCaffrey, E.F.; Baertsch, R.; Sokolov, A.; Meyerowitz, J.G.; Mathis, C.; et al. N-Myc Drives Neuroendocrine Prostate Cancer Initiated from Human Prostate Epithelial Cells. Cancer Cell 2016, 29, 536–547. [Google Scholar] [CrossRef] [Green Version]
  156. Zhang, W.; Liu, B.; Wu, W.; Li, L.; Broom, B.M.; Basourakos, S.P.; Korentzelos, D.; Luan, Y.; Wang, J.; Yang, G.; et al. Targeting the MYCN-PARP-DNA Damage Response Pathway in Neuroendocrine Prostate Cancer. Clin. Cancer Res. 2018, 24, 696–707. [Google Scholar] [CrossRef] [Green Version]
  157. Guo, H.; Ci, X.; Ahmed, M.; Hua, J.T.; Soares, F.; Lin, D.; Puca, L.; Vosoughi, A.; Xue, H.; Li, E.; et al. ONECUT2 is a driver of neuroendocrine prostate cancer. Nat. Commun. 2019, 10, 278. [Google Scholar] [CrossRef] [Green Version]
  158. Varambally, S.; Dhanasekaran, S.M.; Zhou, M.; Barrette, T.R.; Kumar-Sinha, C.; Sanda, M.G.; Ghosh, D.; Pienta, K.J.; Sewalt, R.G.; Otte, A.P.; et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature 2002, 419, 624–629. [Google Scholar] [CrossRef]
  159. Zhang, Y.; Zheng, D.; Zhou, T.; Song, H.; Hulsurkar, M.; Su, N.; Liu, Y.; Wang, Z.; Shao, L.; Ittmann, M.; et al. Androgen deprivation promotes neuroendocrine differentiation and angiogenesis through CREB-EZH2-TSP1 pathway in prostate cancers. Nat. Commun. 2018, 9, 4080. [Google Scholar] [CrossRef]
  160. Li, Y.; Donmez, N.; Sahinalp, C.; Xie, N.; Wang, Y.; Xue, H.; Mo, F.; Beltran, H.; Gleave, M.; Collins, C.; et al. SRRM4 Drives Neuroendocrine Transdifferentiation of Prostate Adenocarcinoma Under Androgen Receptor Pathway Inhibition. Eur. Urol. 2017, 71, 68–78. [Google Scholar] [CrossRef]
  161. Lee, A.R.; Gan, Y.; Tang, Y.; Dong, X. A novel mechanism of SRRM4 in promoting neuroendocrine prostate cancer development via a pluripotency gene network. EBioMedicine 2018, 35, 167–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Akamatsu, S.; Wyatt, A.W.; Lin, D.; Lysakowski, S.; Zhang, F.; Kim, S.; Tse, C.; Wang, K.; Mo, F.; Haegert, A.; et al. The Placental Gene PEG10 Promotes Progression of Neuroendocrine Prostate Cancer. Cell Rep. 2015, 12, 922–936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Kim, S.; Thaper, D.; Bidnur, S.; Toren, P.; Akamatsu, S.; Bishop, J.L.; Colins, C.; Vahid, S.; Zoubeidi, A. PEG10 is associated with treatment-induced neuroendocrine prostate cancer. J. Mol. Endocrinol. 2019, 63, 39–49. [Google Scholar] [CrossRef] [PubMed]
  164. Svensson, C.; Ceder, J.; Iglesias-Gato, D.; Chuan, Y.C.; Pang, S.T.; Bjartell, A.; Martinez, R.M.; Bott, L.; Helczynski, L.; Ulmert, D.; et al. REST mediates androgen receptor actions on gene repression and predicts early recurrence of prostate cancer. Nucleic Acids Res. 2014, 42, 999–1015. [Google Scholar] [CrossRef]
  165. Ballas, N.; Grunseich, C.; Lu, D.D.; Speh, J.C.; Mandel, G. REST and its corepressors mediate plasticity of neuronal gene chromatin throughout neurogenesis. Cell 2005, 121, 645–657. [Google Scholar] [CrossRef] [Green Version]
  166. Gao, Z.; Ure, K.; Ding, P.; Nashaat, M.; Yuan, L.; Ma, J.; Hammer, R.E.; Hsieh, J. The master negative regulator REST/NRSF controls adult neurogenesis by restraining the neurogenic program in quiescent stem cells. J. Neurosci. 2011, 31, 9772–9786. [Google Scholar] [CrossRef] [Green Version]
  167. Liu, K.; Wang, S.; Liu, Y.; Gu, J.; Gu, S.; Xu, Z.; Zhang, R.; Wang, Z.; Ma, H.; Chen, Y.; et al. Overexpression of MYCN promotes proliferation of non-small cell lung cancer. Tumour Biol. 2016, 37, 12855–12866. [Google Scholar] [CrossRef]
  168. Metz, E.P.; Wilder, P.J.; Dong, J.; Datta, K.; Rizzino, A. Elevating SOX2 in prostate tumor cells upregulates expression of neuroendocrine genes, but does not reduce the inhibitory effects of enzalutamide. J. Cell Physiol. 2020, 235, 3731–3740. [Google Scholar] [CrossRef]
  169. Niu, W.; Zang, T.; Smith, D.K.; Vue, T.Y.; Zou, Y.; Bachoo, R.; Johnson, J.E.; Zhang, C.L. SOX2 reprograms resident astrocytes into neural progenitors in the adult brain. Stem. Cell Rep. 2015, 4, 780–794. [Google Scholar] [CrossRef] [Green Version]
  170. Mirosevich, J.; Gao, N.; Gupta, A.; Shappell, S.B.; Jove, R.; Matusik, R.J. Expression and role of Foxa proteins in prostate cancer. Prostate 2006, 66, 1013–1028. [Google Scholar] [CrossRef]
  171. Baca, S.C.; Takeda, D.Y.; Seo, J.H.; Hwang, J.; Ku, S.Y.; Arafeh, R.; Arnoff, T.; Agarwal, S.; Bell, C.; O’Connor, E.; et al. Reprogramming of the FOXA1 cistrome in treatment-emergent neuroendocrine prostate cancer. Nat. Commun. 2021, 12, 1979. [Google Scholar] [CrossRef]
  172. Rotinen, M.; You, S.; Yang, J.; Coetzee, S.G.; Reis-Sobreiro, M.; Huang, W.C.; Huang, F.; Pan, X.; Yanez, A.; Hazelett, D.J.; et al. ONECUT2 is a targetable master regulator of lethal prostate cancer that suppresses the androgen axis. Nat. Med. 2018, 24, 1887–1898. [Google Scholar] [CrossRef]
  173. Burke, P.A.; Gregg, J.P.; Bakhtiar, B.; Beckett, L.A.; Denardo, G.L.; Albrecht, H.; De Vere White, R.W.; De Nardo, S.J. Characterization of MUC1 glycoprotein on prostate cancer for selection of targeting molecules. Int. J. Oncol. 2006, 29, 49–55. [Google Scholar] [CrossRef] [Green Version]
  174. Rajabi, H.; Kufe, D. MUC1-C Oncoprotein Integrates a Program of EMT, Epigenetic Reprogramming and Immune Evasion in Human Carcinomas. Biochim. Biophys Acta Rev. Cancer 2017, 1868, 117–122. [Google Scholar] [CrossRef]
  175. Yasumizu, Y.; Rajabi, H.; Jin, C.; Hata, T.; Pitroda, S.; Long, M.D.; Hagiwara, M.; Li, W.; Hu, Q.; Liu, S.; et al. MUC1-C regulates lineage plasticity driving progression to neuroendocrine prostate cancer. Nat. Commun. 2020, 11, 338. [Google Scholar] [CrossRef] [Green Version]
  176. Hagiwara, M.; Yasumizu, Y.; Yamashita, N.; Rajabi, H.; Fushimi, A.; Long, M.D.; Li, W.; Bhattacharya, A.; Ahmad, R.; Oya, M.; et al. MUC1-C Activates the BAF (mSWI/SNF) Complex in Prostate Cancer Stem Cells. Cancer Res. 2021, 81, 1111–1122. [Google Scholar] [CrossRef]
  177. Kwon, O.J.; Zhang, L.; Jia, D.; Zhou, Z.; Li, Z.; Haffner, M.; Lee, J.K.; True, L.; Morrissey, C.; Xin, L. De novo induction of lineage plasticity from human prostate luminal epithelial cells by activated AKT1 and c-Myc. Oncogene 2020, 39, 7142–7151. [Google Scholar] [CrossRef]
  178. Hsu, E.C.; Rice, M.A.; Bermudez, A.; Marques, F.J.G.; Aslan, M.; Liu, S.; Ghoochani, A.; Zhang, C.A.; Chen, Y.S.; Zlitni, A.; et al. Trop2 is a driver of metastatic prostate cancer with neuroendocrine phenotype via PARP1. Proc. Natl. Acad. Sci. USA 2020, 117, 2032–2042. [Google Scholar] [CrossRef] [Green Version]
  179. Abramovic, I.; Ulamec, M.; Katusic Bojanac, A.; Bulic-Jakus, F.; Jezek, D.; Sincic, N. miRNA in prostate cancer: Challenges toward translation. Epigenomics 2020, 12, 543–558. [Google Scholar] [CrossRef]
  180. Alles, J.; Fehlmann, T.; Fischer, U.; Backes, C.; Galata, V.; Minet, M.; Hart, M.; Abu-Halima, M.; Grasser, F.A.; Lenhof, H.P.; et al. An estimate of the total number of true human miRNAs. Nucleic Acids Res. 2019, 47, 3353–3364. [Google Scholar] [CrossRef] [Green Version]
  181. Kabekkodu, S.P.; Shukla, V.; Varghese, V.K.; Jeevitha, D.S.; Chakrabarty, S.; Satyamoorthy, K. Clustered miRNAs and their role in biological functions and diseases. Biol. Rev. Camb. Philos. Soc. 2018, 93, 1955–1986. [Google Scholar] [CrossRef] [PubMed]
  182. O’Brien, J.; Hayder, H.; Zayed, Y.; Peng, C. Overview of MicroRNA Biogenesis, Mechanisms of Actions, and Circulation. Front. Endocrinol. (Lausanne) 2018, 9, 402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Vasudevan, S.; Tong, Y.; Steitz, J.A. Switching from repression to activation: MicroRNAs can up-regulate translation. Science 2007, 318, 1931–1934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Bayraktar, R.; Bertilaccio, M.T.S.; Calin, G.A. The Interaction Between Two Worlds: MicroRNAs and Toll-Like Receptors. Front. Immunol. 2019, 10, 1053. [Google Scholar] [CrossRef] [Green Version]
  185. Quillet, A.; Saad, C.; Ferry, G.; Anouar, Y.; Vergne, N.; Lecroq, T.; Dubessy, C. Improving Bioinformatics Prediction of microRNA Targets by Ranks Aggregation. Front. Genet. 2019, 10, 1330. [Google Scholar] [CrossRef] [Green Version]
  186. Huang, H.Y.; Lin, Y.C.; Li, J.; Huang, K.Y.; Shrestha, S.; Hong, H.C.; Tang, Y.; Chen, Y.G.; Jin, C.N.; Yu, Y.; et al. miRTarBase 2020: Updates to the experimentally validated microRNA-target interaction database. Nucleic Acids Res. 2020, 48, D148–D154. [Google Scholar] [CrossRef] [Green Version]
  187. Jiang, X.; Guo, S.; Zhang, Y.; Zhao, Y.; Li, X.; Jia, Y.; Xu, Y.; Ma, B. LncRNA NEAT1 promotes docetaxel resistance in prostate cancer by regulating ACSL4 via sponging miR-34a-5p and miR-204-5p. Cell Signal. 2020, 65, 109422. [Google Scholar] [CrossRef]
  188. Thomson, D.W.; Dinger, M.E. Endogenous microRNA sponges: Evidence and controversy. Nat. Rev. Genet. 2016, 17, 272–283. [Google Scholar] [CrossRef]
  189. Lo, U.G.; Pong, R.C.; Yang, D.; Gandee, L.; Hernandez, E.; Dang, A.; Lin, C.J.; Santoyo, J.; Ma, S.; Sonavane, R.; et al. IFNgamma-Induced IFIT5 Promotes Epithelial-to-Mesenchymal Transition in Prostate Cancer via miRNA Processing. Cancer Res. 2019, 79, 1098–1112. [Google Scholar] [CrossRef]
  190. Fernandes, R.C.; Toubia, J.; Townley, S.; Hanson, A.R.; Dredge, B.K.; Pillman, K.A.; Bert, A.G.; Winter, J.M.; Iggo, R.; Das, R.; et al. Post-transcriptional Gene Regulation by MicroRNA-194 Promotes Neuroendocrine Transdifferentiation in Prostate Cancer. Cell Rep. 2021, 34, 108585. [Google Scholar] [CrossRef]
  191. Das, R.; Gregory, P.A.; Fernandes, R.C.; Denis, I.; Wang, Q.; Townley, S.L.; Zhao, S.G.; Hanson, A.R.; Pickering, M.A.; Armstrong, H.K.; et al. MicroRNA-194 Promotes Prostate Cancer Metastasis by Inhibiting SOCS2. Cancer Res. 2017, 77, 1021–1034. [Google Scholar] [CrossRef] [Green Version]
  192. Selth, L.A.; Townley, S.L.; Bert, A.G.; Stricker, P.D.; Sutherland, P.D.; Horvath, L.G.; Goodall, G.J.; Butler, L.M.; Tilley, W.D. Circulating microRNAs predict biochemical recurrence in prostate cancer patients. Br. J. Cancer 2013, 109, 641–650. [Google Scholar] [CrossRef] [Green Version]
  193. Tong, A.W.; Fulgham, P.; Jay, C.; Chen, P.; Khalil, I.; Liu, S.; Senzer, N.; Eklund, A.C.; Han, J.; Nemunaitis, J. MicroRNA profile analysis of human prostate cancers. Cancer Gene Ther. 2009, 16, 206–216. [Google Scholar] [CrossRef] [Green Version]
  194. Bhagirath, D.; Liston, M.; Patel, N.; Akoto, T.; Lui, B.; Yang, T.L.; To, D.M.; Majid, S.; Dahiya, R.; Tabatabai, Z.L.; et al. MicroRNA determinants of neuroendocrine differentiation in metastatic castration-resistant prostate cancer. Oncogene 2020, 39, 7209–7223. [Google Scholar] [CrossRef]
  195. Abdelmohsen, K.; Hutchison, E.R.; Lee, E.K.; Kuwano, Y.; Kim, M.M.; Masuda, K.; Srikantan, S.; Subaran, S.S.; Marasa, B.S.; Mattson, M.P.; et al. miR-375 inhibits differentiation of neurites by lowering HuD levels. Mol. Cell Biol. 2010, 30, 4197–4210. [Google Scholar] [CrossRef] [Green Version]
  196. Liu, Y.; Xing, R.; Zhang, X.; Dong, W.; Zhang, J.; Yan, Z.; Li, W.; Cui, J.; Lu, Y. miR-375 targets the p53 gene to regulate cellular response to ionizing radiation and etoposide in gastric cancer cells. DNA Repair (Amst) 2013, 12, 741–750. [Google Scholar] [CrossRef]
  197. Wang, Y.; Lieberman, R.; Pan, J.; Zhang, Q.; Du, M.; Zhang, P.; Nevalainen, M.; Kohli, M.; Shenoy, N.K.; Meng, H.; et al. miR-375 induces docetaxel resistance in prostate cancer by targeting SEC23A and YAP1. Mol. Cancer 2016, 15, 70. [Google Scholar] [CrossRef] [Green Version]
  198. He, S.; Shi, J.; Mao, J.; Luo, X.; Liu, W.; Liu, R.; Yang, F. The expression of miR-375 in prostate cancer: A study based on GEO, TCGA data and bioinformatics analysis. Pathol. Res. Pr. 2019, 215, 152375. [Google Scholar] [CrossRef]
  199. Valera, V.A.; Parra-Medina, R.; Walter, B.A.; Pinto, P.; Merino, M.J. microRNA Expression Profiling in Young Prostate Cancer Patients. J. Cancer 2020, 11, 4106–4114. [Google Scholar] [CrossRef] [Green Version]
  200. Costa-Pinheiro, P.; Ramalho-Carvalho, J.; Vieira, F.Q.; Torres-Ferreira, J.; Oliveira, J.; Goncalves, C.S.; Costa, B.M.; Henrique, R.; Jeronimo, C. MicroRNA-375 plays a dual role in prostate carcinogenesis. Clin. Epigenetics 2015, 7, 42. [Google Scholar] [CrossRef] [Green Version]
  201. Benoist, G.E.; van Oort, I.M.; Boerrigter, E.; Verhaegh, G.W.; van Hooij, O.; Groen, L.; Smit, F.; de Mol, P.; Hamberg, P.; Dezentje, V.O.; et al. Prognostic Value of Novel Liquid Biomarkers in Patients with Metastatic Castration-Resistant Prostate Cancer Treated with Enzalutamide: A Prospective Observational Study. Clin. Chem. 2020, 66, 842–851. [Google Scholar] [CrossRef] [PubMed]
  202. Selth, L.A.; Das, R.; Townley, S.L.; Coutinho, I.; Hanson, A.R.; Centenera, M.M.; Stylianou, N.; Sweeney, K.; Soekmadji, C.; Jovanovic, L.; et al. A ZEB1-miR-375-YAP1 pathway regulates epithelial plasticity in prostate cancer. Oncogene 2017, 36, 24–34. [Google Scholar] [CrossRef] [PubMed]
  203. Pickl, J.M.; Tichy, D.; Kuryshev, V.Y.; Tolstov, Y.; Falkenstein, M.; Schuler, J.; Reidenbach, D.; Hotz-Wagenblatt, A.; Kristiansen, G.; Roth, W.; et al. Ago-RIP-Seq identifies Polycomb repressive complex I member CBX7 as a major target of miR-375 in prostate cancer progression. Oncotarget 2016, 7, 59589–59603. [Google Scholar] [CrossRef] [PubMed]
  204. Huang, X.; Yuan, T.; Liang, M.; Du, M.; Xia, S.S.; Dittmar, R.; Wang, D.; See, W.; Costello, B.A.; Quevedo, F.; et al. Exosomal miR-1290 and miR-375 as prognostic markers in castration-resistant prostate cancer. Eur. Urol. 2015, 67, 33–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Zedan, A.H.; Osther, P.J.S.; Assenholt, J.; Madsen, J.S.; Hansen, T.F. Circulating miR-141 and miR-375 are associated with treatment outcome in metastatic castration resistant prostate cancer. Sci. Rep. 2020, 10, 227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Cheng, H.H.; Plets, M.; Li, H.; Higano, C.S.; Tangen, C.M.; Agarwal, N.; Vogelzang, N.J.; Hussain, M.; Thompson, I.M., Jr.; Tewari, M.; et al. Circulating microRNAs and treatment response in the Phase II SWOG S0925 study for patients with new metastatic hormone-sensitive prostate cancer. Prostate 2018, 78, 121–127. [Google Scholar] [CrossRef]
  207. Foj, L.; Ferrer, F.; Serra, M.; Arevalo, A.; Gavagnach, M.; Gimenez, N.; Filella, X. Exosomal and Non-Exosomal Urinary miRNAs in Prostate Cancer Detection and Prognosis. Prostate 2017, 77, 573–583. [Google Scholar] [CrossRef]
  208. Ciszkowicz, E.; Porzycki, P.; Semik, M.; Kaznowska, E.; Tyrka, M. MiR-93/miR-375: Diagnostic Potential, Aggressiveness Correlation and Common Target Genes in Prostate Cancer. Int. J. Mol. Sci. 2020, 21, 5667. [Google Scholar] [CrossRef]
  209. Jin, W.; Fei, X.; Wang, X.; Chen, F.; Song, Y. Circulating miRNAs as Biomarkers for Prostate Cancer Diagnosis in Subjects with Benign Prostatic Hyperplasia. J. Immunol. Res. 2020, 2020, 5873056. [Google Scholar] [CrossRef]
  210. Haldrup, C.; Kosaka, N.; Ochiya, T.; Borre, M.; Hoyer, S.; Orntoft, T.F.; Sorensen, K.D. Profiling of circulating microRNAs for prostate cancer biomarker discovery. Drug Deliv. Transl. Res. 2014, 4, 19–30. [Google Scholar] [CrossRef]
  211. Fredsoe, J.; Rasmussen, A.K.I.; Mouritzen, P.; Bjerre, M.T.; Ostergren, P.; Fode, M.; Borre, M.; Sorensen, K.D. Profiling of Circulating microRNAs in Prostate Cancer Reveals Diagnostic Biomarker Potential. Diagnostics 2020, 10, 188. [Google Scholar] [CrossRef] [Green Version]
  212. Bidarra, D.; Constancio, V.; Barros-Silva, D.; Ramalho-Carvalho, J.; Moreira-Barbosa, C.; Antunes, L.; Mauricio, J.; Oliveira, J.; Henrique, R.; Jeronimo, C. Circulating MicroRNAs as Biomarkers for Prostate Cancer Detection and Metastasis Development Prediction. Front. Oncol. 2019, 9, 900. [Google Scholar] [CrossRef] [Green Version]
  213. Paiva, R.M.; Zauli, D.A.G.; Neto, B.S.; Brum, I.S. Urinary microRNAs expression in prostate cancer diagnosis: A systematic review. Clin. Transl. Oncol. 2020, 22, 2061–2073. [Google Scholar] [CrossRef]
  214. Wach, S.; Al-Janabi, O.; Weigelt, K.; Fischer, K.; Greither, T.; Marcou, M.; Theil, G.; Nolte, E.; Holzhausen, H.J.; Stohr, R.; et al. The combined serum levels of miR-375 and urokinase plasminogen activator receptor are suggested as diagnostic and prognostic biomarkers in prostate cancer. Int. J. Cancer 2015, 137, 1406–1416. [Google Scholar] [CrossRef]
  215. Chu, M.; Chang, Y.; Li, P.; Guo, Y.; Zhang, K.; Gao, W. Androgen receptor is negatively correlated with the methylation-mediated transcriptional repression of miR-375 in human prostate cancer cells. Oncol. Rep. 2014, 31, 34–40. [Google Scholar] [CrossRef] [Green Version]
  216. Pillman, K.A.; Phillips, C.A.; Roslan, S.; Toubia, J.; Dredge, B.K.; Bert, A.G.; Lumb, R.; Neumann, D.P.; Li, X.; Conn, S.J.; et al. miR-200/375 control epithelial plasticity-associated alternative splicing by repressing the RNA-binding protein Quaking. EMBO J. 2018, 37, e99016. [Google Scholar] [CrossRef]
  217. Bhagirath, D.; Liston, M.; Akoto, T.; Lui, B.; Bensing, B.A.; Sharma, A.; Saini, S. Novel, non-invasive markers for detecting therapy induced neuroendocrine differentiation in castration-resistant prostate cancer patients. Sci. Rep. 2021, 11, 8279. [Google Scholar] [CrossRef]
  218. Ma, F.; Zhang, J.; Zhong, L.; Wang, L.; Liu, Y.; Wang, Y.; Peng, L.; Guo, B. Upregulated microRNA-301a in breast cancer promotes tumor metastasis by targeting PTEN and activating Wnt/beta-catenin signaling. Gene 2014, 535, 191–197. [Google Scholar] [CrossRef]
  219. Xie, H.; Li, L.; Zhu, G.; Dang, Q.; Ma, Z.; He, D.; Chang, L.; Song, W.; Chang, H.C.; Krolewski, J.J.; et al. Infiltrated pre-adipocytes increase prostate cancer metastasis via modulation of the miR-301a/androgen receptor (AR)/TGF-beta1/Smad/MMP9 signals. Oncotarget 2015, 6, 12326–12339. [Google Scholar] [CrossRef]
  220. Fan, L.; Wang, Y.; Huo, W.; Wang, W.H. MicroRNA301a3p overexpression promotes cell invasion and proliferation by targeting runtrelated transcription factor 3 in prostate cancer. Mol. Med. Rep. 2019, 20, 3755–3763. [Google Scholar] [CrossRef]
  221. Li, X.; Li, J.; Cai, Y.; Peng, S.; Wang, J.; Xiao, Z.; Wang, Y.; Tao, Y.; Leng, Q.; Wu, D.; et al. Hyperglycaemia-induced miR-301a promotes cell proliferation by repressing p21 and Smad4 in prostate cancer. Cancer Lett. 2018, 418, 211–220. [Google Scholar] [CrossRef] [PubMed]
  222. Wang, W.; Liu, M.; Guan, Y.; Wu, Q. Hypoxia-Responsive Mir-301a and Mir-301b Promote Radioresistance of Prostate Cancer Cells via Downregulating NDRG2. Med. Sci. Monit. 2016, 22, 2126–2132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Nam, R.K.; Benatar, T.; Wallis, C.J.; Amemiya, Y.; Yang, W.; Garbens, A.; Naeim, M.; Sherman, C.; Sugar, L.; Seth, A. MiR-301a regulates E-cadherin expression and is predictive of prostate cancer recurrence. Prostate 2016, 76, 869–884. [Google Scholar] [CrossRef] [PubMed]
  224. Damodaran, C.; Das, T.P.; Papu John, A.M.; Suman, S.; Kolluru, V.; Morris, T.J.; Faber, E.N.; Rai, S.N.; Messer, J.C.; Alatassi, H.; et al. miR-301a expression: A prognostic marker for prostate cancer. Urol. Oncol. 2016, 34, 336.e13–336.e20. [Google Scholar] [CrossRef] [Green Version]
  225. Nam, R.K.; Amemiya, Y.; Benatar, T.; Wallis, C.J.; Stojcic-Bendavid, J.; Bacopulos, S.; Sherman, C.; Sugar, L.; Naeim, M.; Yang, W.; et al. Identification and Validation of a Five MicroRNA Signature Predictive of Prostate Cancer Recurrence and Metastasis: A Cohort Study. J. Cancer 2015, 6, 1160–1171. [Google Scholar] [CrossRef]
  226. Kolluru, V.; Chandrasekaran, B.; Tyagi, A.; Dervishi, A.; Ankem, M.; Yan, X.; Maiying, K.; Alatassi, H.; Shaheen, S.P.; Messer, J.C.; et al. miR-301a expression: Diagnostic and prognostic marker for prostate cancer. Urol. Oncol. 2018, 36, 503.e509–503.e515. [Google Scholar] [CrossRef]
  227. Saran, U.; Chandrasekaran, B.; Kolluru, V.; Tyagi, A.; Nguyen, K.D.; Valadon, C.L.; Shaheen, S.P.; Kong, M.; Poddar, T.; Ankem, M.K.; et al. Diagnostic molecular markers predicting aggressive potential in low-grade prostate cancer. Transl. Res. 2021, 231, 92–101. [Google Scholar] [CrossRef]
  228. Dankert, J.T.; Wiesehofer, M.; Czyrnik, E.D.; Singer, B.B.; von Ostau, N.; Wennemuth, G. The deregulation of miR-17/CCND1 axis during neuroendocrine transdifferentiation of LNCaP prostate cancer cells. PLoS ONE 2018, 13, e0200472. [Google Scholar] [CrossRef]
  229. Arabi, L.; Gsponer, J.R.; Smida, J.; Nathrath, M.; Perrina, V.; Jundt, G.; Ruiz, C.; Quagliata, L.; Baumhoer, D. Upregulation of the miR-17-92 cluster and its two paraloga in osteosarcoma—Reasons and consequences. Genes Cancer 2014, 5, 56–63. [Google Scholar] [CrossRef] [Green Version]
  230. Fang, Y.; Shen, H.; Li, H.; Cao, Y.; Qin, R.; Long, L.; Zhu, X.; Xie, C.; Xu, W. miR-106a confers cisplatin resistance by regulating PTEN/Akt pathway in gastric cancer cells. Acta Biochim. Biophys Sin. (Shanghai) 2013, 45, 963–972. [Google Scholar] [CrossRef] [Green Version]
  231. Jiang, Y.; Wu, Y.; Greenlee, A.R.; Wu, J.; Han, Z.; Li, X.; Zhao, Y. miR-106a-mediated malignant transformation of cells induced by anti-benzo[a]pyrene-trans-7,8-diol-9,10-epoxide. Toxicol. Sci. 2011, 119, 50–60. [Google Scholar] [CrossRef] [Green Version]
  232. Hoey, C.; Ray, J.; Jeon, J.; Huang, X.; Taeb, S.; Ylanko, J.; Andrews, D.W.; Boutros, P.C.; Liu, S.K. miRNA-106a and prostate cancer radioresistance: A novel role for LITAF in ATM regulation. Mol. Oncol. 2018, 12, 1324–1341. [Google Scholar] [CrossRef] [Green Version]
  233. Cochetti, G.; Poli, G.; Guelfi, G.; Boni, A.; Egidi, M.G.; Mearini, E. Different levels of serum microRNAs in prostate cancer and benign prostatic hyperplasia: Evaluation of potential diagnostic and prognostic role. Onco Targets Ther. 2016, 9, 7545–7553. [Google Scholar] [CrossRef] [Green Version]
  234. Volinia, S.; Calin, G.A.; Liu, C.G.; Ambs, S.; Cimmino, A.; Petrocca, F.; Visone, R.; Iorio, M.; Roldo, C.; Ferracin, M.; et al. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc. Natl. Acad. Sci USA 2006, 103, 2257–2261. [Google Scholar] [CrossRef] [Green Version]
  235. Luo, B.; Kang, N.; Chen, Y.; Liu, L.; Zhang, Y. Oncogene miR-106a promotes proliferation and metastasis of prostate cancer cells by directly targeting PTEN in vivo and in vitro. Minerva Med. 2018, 109, 24–30. [Google Scholar] [CrossRef]
  236. 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]
  237. Sharova, E.; Grassi, A.; Marcer, A.; Ruggero, K.; Pinto, F.; Bassi, P.; Zanovello, P.; Zattoni, F.; D’Agostino, D.M.; Iafrate, M.; et al. A circulating miRNA assay as a first-line test for prostate cancer screening. Br. J. Cancer 2016, 114, 1362–1366. [Google Scholar] [CrossRef] [Green Version]
  238. Alhasan, A.H.; Scott, A.W.; Wu, J.J.; Feng, G.; Meeks, J.J.; Thaxton, C.S.; Mirkin, C.A. Circulating microRNA signature for the diagnosis of very high-risk prostate cancer. Proc. Natl. Acad. Sci. USA 2016, 113, 10655–10660. [Google Scholar] [CrossRef] [Green Version]
  239. Wei, P.; Yang, J.; Zhang, D.; Cui, M.; Li, L. lncRNA HAND2-AS1 Regulates Prostate Cancer Cell Growth Through Targeting the miR-106a-5p/RBM24 Axis. Onco Targets Ther. 2020, 13, 4523–4531. [Google Scholar] [CrossRef]
  240. Xia, T.; Liao, Q.; Jiang, X.; Shao, Y.; Xiao, B.; Xi, Y.; Guo, J. Long noncoding RNA associated-competing endogenous RNAs in gastric cancer. Sci. Rep. 2014, 4, 6088. [Google Scholar] [CrossRef] [Green Version]
  241. Zhang, G.; Zhou, H.; Xiao, H.; Liu, Z.; Tian, H.; Zhou, T. MicroRNA-92a functions as an oncogene in colorectal cancer by targeting PTEN. Dig. Dis. Sci. 2014, 59, 98–107. [Google Scholar] [CrossRef] [PubMed]
  242. Liao, G.; Xiong, H.; Tang, J.; Li, Y.; Liu, Y. MicroRNA-92a Inhibits the Cell Viability and Metastasis of Prostate Cancer by Targeting SOX4. Technol. Cancer Res. Treat 2020, 19, 1533033820959354. [Google Scholar] [CrossRef] [PubMed]
  243. Xiaoli, Z.; Yawei, W.; Lianna, L.; Haifeng, L.; Hui, Z. Screening of Target Genes and Regulatory Function of miRNAs as Prognostic Indicators for Prostate Cancer. Med. Sci. Monit. 2015, 21, 3748–3759. [Google Scholar] [CrossRef] [Green Version]
  244. Zhang, R.; Li, F.; Wang, Y.; Yao, M.; Chi, C. Prognostic value of microRNA-20b expression level in patients with prostate cancer. Histol. Histopathol. 2020, 35, 827–831. [Google Scholar] [CrossRef] [PubMed]
  245. Tian, L.; Fang, Y.X.; Xue, J.L.; Chen, J.Z. Four microRNAs promote prostate cell proliferation with regulation of PTEN and its downstream signals in vitro. PLoS ONE 2013, 8, e75885. [Google Scholar] [CrossRef]
  246. Konoshenko, M.Y.; Lekchnov, E.A.; Bryzgunova, O.E.; Zaporozhchenko, I.A.; Yarmoschuk, S.V.; Pashkovskaya, O.A.; Pak, S.V.; Laktionov, P.P. The Panel of 12 Cell-Free MicroRNAs as Potential Biomarkers in Prostate Neoplasms. Diagnostics 2020, 10, 38. [Google Scholar] [CrossRef] [Green Version]
  247. Huo, W.; Qi, F.; Wang, K. Long non-coding RNA FER1L4 inhibits prostate cancer progression via sponging miR-92a-3p and upregulation of FBXW7. Cancer Cell Int. 2020, 20, 64. [Google Scholar] [CrossRef]
  248. Buechner, J.; Tomte, E.; Haug, B.H.; Henriksen, J.R.; Lokke, C.; Flaegstad, T.; Einvik, C. Tumour-suppressor microRNAs let-7 and mir-101 target the proto-oncogene MYCN and inhibit cell proliferation in MYCN-amplified neuroblastoma. Br. J. Cancer 2011, 105, 296–303. [Google Scholar] [CrossRef] [Green Version]
  249. Olive, V.; Bennett, M.J.; Walker, J.C.; Ma, C.; Jiang, I.; Cordon-Cardo, C.; Li, Q.J.; Lowe, S.W.; Hannon, G.J.; He, L. miR-19 is a key oncogenic component of mir-17-92. Genes Dev. 2009, 23, 2839–2849. [Google Scholar] [CrossRef] [Green Version]
  250. Fan, Y.; Yin, S.; Hao, Y.; Yang, J.; Zhang, H.; Sun, C.; Ma, M.; Chang, Q.; Xi, J.J. miR-19b promotes tumor growth and metastasis via targeting TP53. Rna 2014, 20, 765–772. [Google Scholar] [CrossRef] [Green Version]
  251. Osip’yants, A.I.; Knyazev, E.N.; Galatenko, A.V.; Nyushko, K.M.; Galatenko, V.V.; Shkurnikov, M.Y.; Alekseev, B.Y. Changes in the Level of Circulating hsa-miR-297 and hsa-miR-19b-3p miRNA Are Associated with Generalization of Prostate Cancer. Bull. Exp. Biol. Med. 2017, 162, 379–382. [Google Scholar] [CrossRef]
  252. Stuopelyte, K.; Daniunaite, K.; Jankevicius, F.; Jarmalaite, S. Detection of miRNAs in urine of prostate cancer patients. Medicina 2016, 52, 116–124. [Google Scholar] [CrossRef]
  253. Duca, R.B.; Massillo, C.; Dalton, G.N.; Farré, P.L.; Graña, K.D.; Gardner, K.; De Siervi, A. MiR-19b-3p and miR-101-3p as potential biomarkers for prostate cancer diagnosis and prognosis. Am. J. Cancer Res. 2021, 11, 2802–2820. [Google Scholar]
  254. Li, D.; Ilnytskyy, Y.; Kovalchuk, A.; Khachigian, L.M.; Bronson, R.T.; Wang, B.; Kovalchuk, O. Crucial role for early growth response-1 in the transcriptional regulation of miR-20b in breast cancer. Oncotarget 2013, 4, 1373–1387. [Google Scholar] [CrossRef] [Green Version]
  255. Guo, J.; Xiao, Z.; Yu, X.; Cao, R. miR-20b promotes cellular proliferation and migration by directly regulating phosphatase and tensin homolog in prostate cancer. Oncol. Lett. 2017, 14, 6895–6900. [Google Scholar] [CrossRef]
  256. Qi, J.C.; Yang, Z.; Zhang, Y.P.; Lu, B.S.; Yin, Y.W.; Liu, K.L.; Xue, W.Y.; Qu, C.B.; Li, W. miR-20b-5p, TGFBR2, and E2F1 Form a Regulatory Loop to Participate in Epithelial to Mesenchymal Transition in Prostate Cancer. Front. Oncol. 2019, 9, 1535. [Google Scholar] [CrossRef] [Green Version]
  257. Pan, Z.; Mo, F.; Liu, H.; Zeng, J.; Huang, K.; Huang, S.; Cao, Z.; Xu, X.; Xu, J.; Liu, T.; et al. LncRNA prostate androgen-regulated transcript 1 (PART 1) functions as an oncogene in osteosarcoma via sponging miR-20b-5p to upregulate BAMBI. Ann. Transl. Med. 2021, 9, 488. [Google Scholar] [CrossRef]
  258. Pashaei, E.; Ahmady, M.; Ozen, M.; Aydin, N. Meta-analysis of miRNA expression profiles for prostate cancer recurrence following radical prostatectomy. PLoS ONE 2017, 12, e0179543. [Google Scholar] [CrossRef] [Green Version]
  259. Chen, Y.; Lu, X.; Wu, B.; Su, Y.; Li, J.; Wang, H. MicroRNA 363 mediated positive regulation of c-myc translation affect prostate cancer development and progress. Neoplasma 2015, 62, 191–198. [Google Scholar] [CrossRef] [Green Version]
  260. Liang, H.; Studach, L.; Hullinger, R.L.; Xie, J.; Andrisani, O.M. Down-regulation of RE-1 silencing transcription factor (REST) in advanced prostate cancer by hypoxia-induced miR-106b~25. Exp. Cell Res. 2014, 320, 188–199. [Google Scholar] [CrossRef] [Green Version]
  261. Cai, K.; Wang, Y.; Bao, X. MiR-106b promotes cell proliferation via targeting RB in laryngeal carcinoma. J. Exp. Clin. Cancer Res. 2011, 30, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Yang, T.S.; Yang, X.H.; Chen, X.; Wang, X.D.; Hua, J.; Zhou, D.L.; Zhou, B.; Song, Z.S. MicroRNA-106b in cancer-associated fibroblasts from gastric cancer promotes cell migration and invasion by targeting PTEN. FEBS Lett. 2014, 588, 2162–2169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Song, C.J.; Chen, H.; Chen, L.Z.; Ru, G.M.; Guo, J.J.; Ding, Q.N. The potential of microRNAs as human prostate cancer biomarkers: A meta-analysis of related studies. J. Cell Biochem. 2018, 119, 2763–2786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Yin, W.; Chen, J.; Wang, G.; Zhang, D. MicroRNA106b functions as an oncogene and regulates tumor viability and metastasis by targeting LARP4B in prostate cancer. Mol. Med. Rep. 2019, 20, 951–958. [Google Scholar] [CrossRef] [Green Version]
  265. Poliseno, L.; Salmena, L.; Riccardi, L.; Fornari, A.; Song, M.S.; Hobbs, R.M.; Sportoletti, P.; Varmeh, S.; Egia, A.; Fedele, G.; et al. Identification of the miR-106b~25 microRNA cluster as a proto-oncogenic PTEN-targeting intron that cooperates with its host gene MCM7 in transformation. Sci. Signal. 2010, 3, ra29. [Google Scholar] [CrossRef] [Green Version]
  266. Hudson, R.S.; Yi, M.; Esposito, D.; Glynn, S.A.; Starks, A.M.; Yang, Y.; Schetter, A.J.; Watkins, S.K.; Hurwitz, A.A.; Dorsey, T.H.; et al. MicroRNA-106b-25 cluster expression is associated with early disease recurrence and targets caspase-7 and focal adhesion in human prostate cancer. Oncogene 2013, 32, 4139–4147. [Google Scholar] [CrossRef] [Green Version]
  267. Li, B.; Shi, X.B.; Nori, D.; Chao, C.K.; Chen, A.M.; Valicenti, R.; White Rde, V. Down-regulation of microRNA 106b is involved in p21-mediated cell cycle arrest in response to radiation in prostate cancer cells. Prostate 2011, 71, 567–574. [Google Scholar] [CrossRef]
  268. Fu, X.; Tian, J.; Zhang, L.; Chen, Y.; Hao, Q. Involvement of microRNA-93, a new regulator of PTEN/Akt signaling pathway, in regulation of chemotherapeutic drug cisplatin chemosensitivity in ovarian cancer cells. FEBS Lett. 2012, 586, 1279–1286. [Google Scholar] [CrossRef] [Green Version]
  269. Pudova, E.A.; Krasnov, G.S.; Nyushko, K.M.; Kobelyatskaya, A.A.; Savvateeva, M.V.; Poloznikov, A.A.; Dolotkazin, D.R.; Klimina, K.M.; Guvatova, Z.G.; Simanovsky, S.A.; et al. miRNAs expression signature potentially associated with lymphatic dissemination in locally advanced prostate cancer. BMC Med. Genom. 2020, 13, 129. [Google Scholar] [CrossRef]
  270. Liu, J.-J.; Zhang, X.; Wu, X.-H. miR-93 Promotes the Growth and Invasion of Prostate Cancer by Upregulating Its Target Genes TGFBR2, ITGB8, and LATS2. Mol. Ther.-Oncolytics 2018, 11, 14–19. [Google Scholar] [CrossRef] [Green Version]
  271. Choi, N.; Park, J.; Lee, J.-S.; Yoe, J.; Park, G.Y.; Kim, E.; Jeon, H.; Cho, Y.M.; Roh, T.-Y.; Lee, Y. miR-93/miR-106b/miR-375-CIC-CRABP1: A novel regulatory axis in prostate cancer progression. Oncotarget 2015, 6, 23533. [Google Scholar] [CrossRef]
  272. Wang, C.; Tian, S.; Zhang, D.; Deng, J.; Cai, H.; Shi, C.; Yang, W. Increased expression of microRNA-93 correlates with progression and prognosis of prostate cancer. Medicine (Baltimore) 2020, 99, e18432. [Google Scholar] [CrossRef]
  273. Barceló, M.; Castells, M.; Pérez-Riba, M.; Bassas, L.; Vigués, F.; Larriba, S. Seminal plasma microRNAs improve diagnosis/prognosis of prostate cancer in men with moderately altered prostate-specific antigen. Am. J. Transl. Res. 2020, 12, 2041–2051. [Google Scholar]
  274. Martínez-González, L.J.; Sánchez-Conde, V.; González-Cabezuelo, J.M.; Antunez-Rodríguez, A.; Andrés-León, E.; Robles-Fernandez, I.; Lorente, J.A.; Vázquez-Alonso, F.; Alvarez-Cubero, M.J. Identification of MicroRNAs as Viable Aggressiveness Biomarkers for Prostate Cancer. Biomedicines 2021, 9, 646. [Google Scholar] [CrossRef]
  275. Zhang, S.; Liu, C.; Zou, X.; Geng, X.; Zhou, X.; Fan, X.; Zhu, D.; Zhang, H.; Zhu, W. MicroRNA panel in serum reveals novel diagnostic biomarkers for prostate cancer. PeerJ 2021, 9, e11441. [Google Scholar] [CrossRef]
  276. Li, Z.; Sun, Y.; Chen, X.; Squires, J.; Nowroozizadeh, B.; Liang, C.; Huang, J. p53 Mutation Directs AURKA Overexpression via miR-25 and FBXW7 in Prostatic Small Cell Neuroendocrine Carcinoma. Mol. Cancer Res. 2015, 13, 584–591. [Google Scholar] [CrossRef] [Green Version]
  277. Esposito, F.; Tornincasa, M.; Pallante, P.; Federico, A.; Borbone, E.; Pierantoni, G.M.; Fusco, A. Down-regulation of the miR-25 and miR-30d contributes to the development of anaplastic thyroid carcinoma targeting the polycomb protein EZH2. J. Clin. Endocrinol. Metab. 2012, 97, E710–E718. [Google Scholar] [CrossRef] [Green Version]
  278. Kumar, M.; Lu, Z.; Takwi, A.A.; Chen, W.; Callander, N.S.; Ramos, K.S.; Young, K.H.; Li, Y. Negative regulation of the tumor suppressor p53 gene by microRNAs. Oncogene 2011, 30, 843–853. [Google Scholar] [CrossRef] [Green Version]
  279. Zoni, E.; van der Horst, G.; van de Merbel, A.F.; Chen, L.; Rane, J.K.; Pelger, R.C.M.; Collins, A.T.; Visakorpi, T.; Snaar-Jagalska, B.E.; Maitland, N.J.; et al. miR-25 Modulates Invasiveness and Dissemination of Human Prostate Cancer Cells via Regulation of αv- and α6- Integrin Expression. Cancer Res. 2015, 75, 2326–2336. [Google Scholar] [CrossRef] [Green Version]
  280. Srivastava, A.; Goldberger, H.; Dimtchev, A.; Marian, C.; Soldin, O.; Li, X.; Collins, S.P.; Suy, S.; Kumar, D. Circulatory miR-628-5p is downregulated in prostate cancer patients. Tumour. Biol. 2014, 35, 4867–4873. [Google Scholar] [CrossRef] [Green Version]
  281. Cai, K.; Wan, Y.; Sun, G.; Shi, L.; Bao, X.; Wang, Z. Let-7a inhibits proliferation and induces apoptosis by targeting EZH2 in nasopharyngeal carcinoma cells. Oncol. Rep. 2012, 28, 2101–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Cai, J.; Yang, C.; Yang, Q.; Ding, H.; Jia, J.; Guo, J.; Wang, J.; Wang, Z. Deregulation of let-7e in epithelial ovarian cancer promotes the development of resistance to cisplatin. Oncogenesis 2013, 2, e75. [Google Scholar] [CrossRef] [PubMed]
  283. Murray, M.J.; Saini, H.K.; Siegler, C.A.; Hanning, J.E.; Barker, E.M.; van Dongen, S.; Ward, D.M.; Raby, K.L.; Groves, I.J.; Scarpini, C.G.; et al. LIN28 Expression in malignant germ cell tumors downregulates let-7 and increases oncogene levels. Cancer Res. 2013, 73, 4872–4884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Cimadamore, F.; Amador-Arjona, A.; Chen, C.; Huang, C.-T.; Terskikh, A.V. SOX2–LIN28/let-7 pathway regulates proliferation and neurogenesis in neural precursors. Proc. Natl. Acad. Sci. USA 2013, 110, E3017–E3026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Kong, D.; Heath, E.; Chen, W.; Cher, M.L.; Powell, I.; Heilbrun, L.; Li, Y.; Ali, S.; Sethi, S.; Hassan, O.; et al. Loss of Let-7 Up-Regulates EZH2 in Prostate Cancer Consistent with the Acquisition of Cancer Stem Cell Signatures That Are Attenuated by BR-DIM. PLoS ONE 2012, 7, e33729. [Google Scholar] [CrossRef] [Green Version]
  286. Schubert, M.; Spahn, M.; Kneitz, S.; Scholz, C.J.; Joniau, S.; Stroebel, P.; Riedmiller, H.; Kneitz, B. Distinct microRNA Expression Profile in Prostate Cancer Patients with Early Clinical Failure and the Impact of let-7 as Prognostic Marker in High-Risk Prostate Cancer. PLoS ONE 2013, 8, e65064. [Google Scholar] [CrossRef] [Green Version]
  287. Costanzi, E.; Romani, R.; Scarpelli, P.; Bellezza, I. Extracellular Vesicles-Mediated Transfer of miRNA Let-7b from PC3 Cells to Macrophages. Genes 2020, 11, 1495. [Google Scholar] [CrossRef]
  288. Guelfi, G.; Cochetti, G.; Stefanetti, V.; Zampini, D.; Diverio, S.; Boni, A.; Mearini, E. Next Generation Sequencing of urine exfoliated cells: An approach of prostate cancer microRNAs research. Sci. Rep. 2018, 8, 7111. [Google Scholar] [CrossRef]
  289. Xiao, G.a.; Yao, J.; Kong, D.; Ye, C.; Chen, R.; Li, L.; Zeng, T.; Wang, L.; Zhang, W.; Shi, X.; et al. The Long Noncoding RNA TTTY15, Which Is Located on the Y Chromosome, Promotes Prostate Cancer Progression by Sponging let-7. Eur. Urol. 2019, 76, 315–326. [Google Scholar] [CrossRef]
  290. Nadiminty, N.; Tummala, R.; Lou, W.; Zhu, Y.; Shi, X.-B.; Zou, J.X.; Chen, H.; Zhang, J.; Chen, X.; Luo, J.; et al. MicroRNA let-7c Is Downregulated in Prostate Cancer and Suppresses Prostate Cancer Growth. PLoS ONE 2012, 7, e32832. [Google Scholar] [CrossRef] [Green Version]
  291. Pernicová, Z.; Slabáková, E.; Fedr, R.; Šimečková, Š.; Jaroš, J.; Suchánková, T.; Bouchal, J.; Kharaishvili, G.; Král, M.; Kozubík, A.; et al. The role of high cell density in the promotion of neuroendocrine transdifferentiation of prostate cancer cells. Mol. Cancer 2014, 13, 113. [Google Scholar] [CrossRef] [Green Version]
  292. Akoto, T.; Bhagirath, D.; Saini, S. MicroRNAs in treatment-induced neuroendocrine differentiation in prostate cancer. Cancer Drug Resist. 2020, 3, 804–818. [Google Scholar] [CrossRef]
  293. Chen, J.; Bardes, E.E.; Aronow, B.J.; Jegga, A.G. ToppGene Suite for gene list enrichment analysis and candidate gene prioritization. Nucleic Acids Res. 2009, 37, W305–W311. [Google Scholar] [CrossRef]
  294. Kong, Q.; Chen, X.S.; Tian, T.; Xia, X.Y.; Xu, P. MicroRNA-194 suppresses prostate cancer migration and invasion by downregulating human nuclear distribution protein. Oncol. Rep. 2017, 37, 803–812. [Google Scholar] [CrossRef]
  295. Gao, S.; Zhao, Z.; Wu, R.; Wu, L.; Tian, X.; Zhang, Z. MicroRNA-194 regulates cell viability and apoptosis by targeting CDH2 in prostatic cancer. Onco Targets Ther. 2018, 11, 4837–4844. [Google Scholar] [CrossRef] [Green Version]
  296. Li, Z.; Li, L.X.; Diao, Y.J.; Wang, J.; Ye, Y.; Hao, X.K. Identification of Urinary Exosomal miRNAs for the Non-Invasive Diagnosis of Prostate Cancer. Cancer Manag. Res. 2021, 13, 25–35. [Google Scholar] [CrossRef]
  297. Konoshenko, M.Y.; Bryzgunova, O.E.; Lekchnov, E.A.; Amelina, E.V.; Yarmoschuk, S.V.; Pak, S.V.; Laktionov, P.P. The Influence of Radical Prostatectomy on the Expression of Cell-Free MiRNA. Diagnostics 2020, 10, 600. [Google Scholar] [CrossRef]
  298. Mosquera, J.M.; Beltran, H.; Park, K.; MacDonald, T.Y.; Robinson, B.D.; Tagawa, S.T.; Perner, S.; Bismar, T.A.; Erbersdobler, A.; Dhir, R.; et al. Concurrent AURKA and MYCN gene amplifications are harbingers of lethal treatment-related neuroendocrine prostate cancer. Neoplasia 2013, 15, 1–10. [Google Scholar] [CrossRef] [Green Version]
  299. Kishore, S.; Jaskiewicz, L.; Burger, L.; Hausser, J.; Khorshid, M.; Zavolan, M. A quantitative analysis of CLIP methods for identifying binding sites of RNA-binding proteins. Nat. Methods 2011, 8, 559–564. [Google Scholar] [CrossRef]
  300. Riley, K.J.; Rabinowitz, G.S.; Yario, T.A.; Luna, J.M.; Darnell, R.B.; Steitz, J.A. EBV and human microRNAs co-target oncogenic and apoptotic viral and human genes during latency. Embo J 2012, 31, 2207–2221. [Google Scholar] [CrossRef]
  301. Heo, I.; Joo, C.; Kim, Y.K.; Ha, M.; Yoon, M.J.; Cho, J.; Yeom, K.H.; Han, J.; Kim, V.N. TUT4 in concert with Lin28 suppresses microRNA biogenesis through pre-microRNA uridylation. Cell 2009, 138, 696–708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Mogilyansky, E.; Rigoutsos, I. The miR-17/92 cluster: A comprehensive update on its genomics, genetics, functions and increasingly important and numerous roles in health and disease. Cell Death Differ. 2013, 20, 1603–1614. [Google Scholar] [CrossRef] [PubMed]
  303. Liu, H.; Wu, Z.; Zhou, H.; Cai, W.; Li, X.; Hu, J.; Gao, L.; Feng, T.; Wang, L.; Peng, X.; et al. The SOX4/miR-17-92/RB1 Axis Promotes Prostate Cancer Progression. Neoplasia 2019, 21, 765–776. [Google Scholar] [CrossRef] [PubMed]
  304. Tan, H.L.; Sood, A.; Rahimi, H.A.; Wang, W.; Gupta, N.; Hicks, J.; Mosier, S.; Gocke, C.D.; Epstein, J.I.; Netto, G.J.; et al. Rb loss is characteristic of prostatic small cell neuroendocrine carcinoma. Clin. Cancer Res. 2014, 20, 890–903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Hu, H.; Li, H.; He, Y. MicroRNA-17 downregulates expression of the PTEN gene to promote the occurrence and development of adenomyosis. Exp. Ther. Med. 2017, 14, 3805–3811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Dong, P.; Xiong, Y.; Yu, J.; Chen, L.; Tao, T.; Yi, S.; Hanley, S.J.B.; Yue, J.; Watari, H.; Sakuragi, N. Control of PD-L1 expression by miR-140/142/340/383 and oncogenic activation of the OCT4-miR-18a pathway in cervical cancer. Oncogene 2018, 37, 5257–5268. [Google Scholar] [CrossRef] [Green Version]
  307. Liang, Z.; Li, Y.; Huang, K.; Wagar, N.; Shim, H. Regulation of miR-19 to breast cancer chemoresistance through targeting PTEN. Pharm Res. 2011, 28, 3091–3100. [Google Scholar] [CrossRef] [Green Version]
  308. Wang, F.; Li, T.; Zhang, B.; Li, H.; Wu, Q.; Yang, L.; Nie, Y.; Wu, K.; Shi, Y.; Fan, D. MicroRNA-19a/b regulates multidrug resistance in human gastric cancer cells by targeting PTEN. Biochem. Biophys Res. Commun. 2013, 434, 688–694. [Google Scholar] [CrossRef]
  309. Trompeter, H.I.; Abbad, H.; Iwaniuk, K.M.; Hafner, M.; Renwick, N.; Tuschl, T.; Schira, J.; Muller, H.W.; Wernet, P. MicroRNAs MiR-17, MiR-20a, and MiR-106b act in concert to modulate E2F activity on cell cycle arrest during neuronal lineage differentiation of USSC. PLoS ONE 2011, 6, e16138. [Google Scholar] [CrossRef]
  310. Jiang, Y.; Chang, H.; Chen, G. Effects of microRNA-20a on the proliferation, migration and apoptosis of multiple myeloma via the PTEN/PI3K/AKT signaling pathway. Oncol. Lett. 2018, 15, 10001–10007. [Google Scholar] [CrossRef] [Green Version]
  311. Lewis, B.P.; Shih, I.H.; Jones-Rhoades, M.W.; Bartel, D.P.; Burge, C.B. Prediction of mammalian microRNA targets. Cell 2003, 115, 787–798. [Google Scholar] [CrossRef] [Green Version]
  312. Lu, W.D.; Zuo, Y.; Xu, Z.; Zhang, M. MiR-19a promotes epithelial-mesenchymal transition through PI3K/AKT pathway in gastric cancer. World J. Gastroenterol. 2015, 21, 4564–4573. [Google Scholar] [CrossRef]
  313. Wiesehofer, M.; Czyrnik, E.D.; Spahn, M.; Ting, S.; Reis, H.; Dankert, J.T.; Wennemuth, G. Increased Expression of AKT3 in Neuroendocrine Differentiated Prostate Cancer Cells Alters the Response Towards Anti-Androgen Treatment. Cancers 2021, 13, 578. [Google Scholar] [CrossRef]
  314. Jiao, L.; Deng, Z.; Xu, C.; Yu, Y.; Li, Y.; Yang, C.; Chen, J.; Liu, Z.; Huang, G.; Li, L.C.; et al. miR-663 induces castration-resistant prostate cancer transformation and predicts clinical recurrence. J. Cell Physiol. 2014, 229, 834–844. [Google Scholar] [CrossRef]
  315. Cho, J.G.; Park, S.; Lim, C.H.; Kim, H.S.; Song, S.Y.; Roh, T.Y.; Sung, J.H.; Suh, W.; Ham, S.J.; Lim, K.H.; et al. ZNF224, Kruppel like zinc finger protein, induces cell growth and apoptosis-resistance by down-regulation of p21 and p53 via miR-663a. Oncotarget 2016, 7, 31177–31190. [Google Scholar] [CrossRef] [Green Version]
  316. Sadeghi, M.; Ranjbar, B.; Ganjalikhany, M.R.; Khan, F.M.; Schmitz, U.; Wolkenhauer, O.; Gupta, S.K. MicroRNA and Transcription Factor Gene Regulatory Network Analysis Reveals Key Regulatory Elements Associated with Prostate Cancer Progression. PLoS ONE 2016, 11, e0168760. [Google Scholar] [CrossRef]
  317. Dang, Q.; Li, L.; Xie, H.; He, D.; Chen, J.; Song, W.; Chang, L.S.; Chang, H.C.; Yeh, S.; Chang, C. Anti-androgen enzalutamide enhances prostate cancer neuroendocrine (NE) differentiation via altering the infiltrated mast cells → androgen receptor (AR) → miRNA32 signals. Mol. Oncol. 2015, 9, 1241–1251. [Google Scholar] [CrossRef] [Green Version]
  318. Jalava, S.E.; Urbanucci, A.; Latonen, L.; Waltering, K.K.; Sahu, B.; Janne, O.A.; Seppala, J.; Lahdesmaki, H.; Tammela, T.L.; Visakorpi, T. Androgen-regulated miR-32 targets BTG2 and is overexpressed in castration-resistant prostate cancer. Oncogene 2012, 31, 4460–4471. [Google Scholar] [CrossRef] [Green Version]
  319. Zhang, L.; Li, X.; Chao, Y.; He, R.; Liu, J.; Yuan, Y.; Zhao, W.; Han, C.; Song, X. KLF4, a miR-32-5p targeted gene, promotes cisplatin-induced apoptosis by upregulating BIK expression in prostate cancer. Cell Commun. Signal. 2018, 16, 53. [Google Scholar] [CrossRef] [Green Version]
  320. Latonen, L.; Scaravilli, M.; Gillen, A.; Hartikainen, S.; Zhang, F.P.; Ruusuvuori, P.; Kujala, P.; Poutanen, M.; Visakorpi, T. In Vivo Expression of miR-32 Induces Proliferation in Prostate Epithelium. Am. J. Pathol. 2017, 187, 2546–2557. [Google Scholar] [CrossRef] [Green Version]
  321. Liao, H.; Xiao, Y.; Hu, Y.; Yin, Z.; Liu, L. microRNA-32 induces radioresistance by targeting DAB2IP and regulating autophagy in prostate cancer cells. Oncol. Lett. 2015, 10, 2055–2062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Zhu, G.; Chai, J.; Ma, L.; Duan, H.; Zhang, H. Downregulated microRNA-32 expression induced by high glucose inhibits cell cycle progression via PTEN upregulation and Akt inactivation in bone marrow-derived mesenchymal stem cells. Biochem. Biophys Res. Commun. 2013, 433, 526–531. [Google Scholar] [CrossRef] [PubMed]
  323. Kong, D.; Heath, E.; Chen, W.; Cher, M.; Powell, I.; Heilbrun, L.; Li, Y.; Ali, S.; Sethi, S.; Hassan, O.; et al. Epigenetic silencing of miR-34a in human prostate cancer cells and tumor tissue specimens can be reversed by BR-DIM treatment. Am. J. Transl. Res. 2012, 4, 14–23. [Google Scholar] [PubMed]
  324. Chen, W.Y.; Liu, S.Y.; Chang, Y.S.; Yin, J.J.; Yeh, H.L.; Mouhieddine, T.H.; Hadadeh, O.; Abou-Kheir, W.; Liu, Y.N. MicroRNA-34a regulates WNT/TCF7 signaling and inhibits bone metastasis in Ras-activated prostate cancer. Oncotarget 2015, 6, 441–457. [Google Scholar] [CrossRef] [Green Version]
  325. Shu, Y.; Ren, L.; Xie, B.; Liang, Z.; Chen, J.M. MiR-204 enhances mitochondrial apoptosis in doxorubicin-treated prostate cancer cells by targeting SIRT1/p53 pathway. Oncotarget 2017, 8, 97313–97322. [Google Scholar] [CrossRef] [Green Version]
  326. Wa, Q.; Huang, S.; Pan, J.; Tang, Y.; He, S.; Fu, X.; Peng, X.; Chen, X.; Yang, C.; Ren, D.; et al. miR-204-5p Represses Bone Metastasis via Inactivating NF-kappaB Signaling in Prostate Cancer. Mol. Ther. Nucleic Acids 2019, 18, 567–579. [Google Scholar] [CrossRef] [Green Version]
  327. Kashat, M.; Azzouz, L.; Sarkar, S.H.; Kong, D.; Li, Y.; Sarkar, F.H. Inactivation of AR and Notch-1 signaling by miR-34a attenuates prostate cancer aggressiveness. Am. J. Transl. Res. 2012, 4, 432–442. [Google Scholar]
  328. Wei, J.S.; Song, Y.K.; Durinck, S.; Chen, Q.R.; Cheuk, A.T.C.; Tsang, P.; Zhang, Q.; Thiele, C.J.; Slack, A.; Shohet, J.; et al. The MYCN oncogene is a direct target of miR-34a. Oncogene 2008, 27, 5204–5213. [Google Scholar] [CrossRef] [Green Version]
  329. Choi, Y.J.; Lin, C.P.; Ho, J.J.; He, X.; Okada, N.; Bu, P.; Zhong, Y.; Kim, S.Y.; Bennett, M.J.; Chen, C.; et al. miR-34 miRNAs provide a barrier for somatic cell reprogramming. Nat. Cell Biol. 2011, 13, 1353–1360. [Google Scholar] [CrossRef]
  330. Javeri, A.; Ghaffarpour, M.; Taha, M.F.; Houshmand, M. Downregulation of miR-34a in breast tumors is not associated with either p53 mutations or promoter hypermethylation while it correlates with metastasis. Med. Oncol 2013, 30, 413. [Google Scholar] [CrossRef]
  331. Zheng, C.; Yinghao, S.; Li, J. MiR-221 expression affects invasion potential of human prostate carcinoma cell lines by targeting DVL2. Med. Oncol. 2012, 29, 815–822. [Google Scholar] [CrossRef]
  332. Kiener, M.; Chen, L.; Krebs, M.; Grosjean, J.; Klima, I.; Kalogirou, C.; Riedmiller, H.; Kneitz, B.; Thalmann, G.N.; Snaar-Jagalska, E.; et al. miR-221-5p regulates proliferation and migration in human prostate cancer cells and reduces tumor growth in vivo. BMC Cancer 2019, 19, 627. [Google Scholar] [CrossRef]
  333. Sun, T.; Du, S.-Y.; Armenia, J.; Qu, F.; Fan, J.; Wang, X.; Fei, T.; Komura, K.; Liu, S.X.; Lee, G.-S.M.; et al. Expression of lncRNA MIR222HG co-transcribed from the miR-221/222 gene promoter facilitates the development of castration-resistant prostate cancer. Oncogenesis 2018, 7, 30. [Google Scholar] [CrossRef] [Green Version]
  334. Xuan, H.; Xue, W.; Pan, J.; Sha, J.; Dong, B.; Huang, Y. Downregulation of miR-221, -30d, and -15a contributes to pathogenesis of prostate cancer by targeting Bmi-1. Biochemistry (Moscow) 2015, 80, 276–283. [Google Scholar] [CrossRef]
  335. Lupini, L.; Bassi, C.; Ferracin, M.; Bartonicek, N.; D’Abundo, L.; Zagatti, B.; Callegari, E.; Musa, G.; Moshiri, F.; Gramantieri, L.; et al. miR-221 affects multiple cancer pathways by modulating the level of hundreds messenger RNAs. Front. Genet. 2013, 4, 64. [Google Scholar] [CrossRef] [Green Version]
  336. Garofalo, M.; Di Leva, G.; Romano, G.; Nuovo, G.; Suh, S.-S.; Ngankeu, A.; Taccioli, C.; Pichiorri, F.; Alder, H.; Secchiero, P.; et al. miR-221&222 Regulate TRAIL Resistance and Enhance Tumorigenicity through PTEN and TIMP3 Downregulation. Cancer Cell 2009, 16, 498–509. [Google Scholar] [CrossRef] [Green Version]
  337. Boyle, G.M.; Woods, S.L.; Bonazzi, V.F.; Stark, M.S.; Hacker, E.; Aoude, L.G.; Dutton-Regester, K.; Cook, A.L.; Sturm, R.A.; Hayward, N.K. Melanoma cell invasiveness is regulated by miR-211 suppression of the BRN2 transcription factor. Pigment. Cell Melanoma Res. 2011, 24, 525–537. [Google Scholar] [CrossRef]
  338. Kumar, B.; Khaleghzadegan, S.; Mears, B.; Hatano, K.; Kudrolli, T.A.; Chowdhury, W.H.; Yeater, D.B.; Ewing, C.M.; Luo, J.; Isaacs, W.B.; et al. Identification of miR-30b-3p and miR-30d-5p as direct regulators of androgen receptor signaling in prostate cancer by complementary functional microRNA library screening. Oncotarget 2016, 7, 72593–72607. [Google Scholar] [CrossRef] [Green Version]
  339. Lin, P.C.; Chiu, Y.L.; Banerjee, S.; Park, K.; Mosquera, J.M.; Giannopoulou, E.; Alves, P.; Tewari, A.K.; Gerstein, M.B.; Beltran, H.; et al. Epigenetic repression of miR-31 disrupts androgen receptor homeostasis and contributes to prostate cancer progression. Cancer Res. 2013, 73, 1232–1244. [Google Scholar] [CrossRef] [Green Version]
  340. Ostano, P.; Mello-Grand, M.; Sesia, D.; Gregnanin, I.; Peraldo-Neia, C.; Guana, F.; Jachetti, E.; Farsetti, A.; Chiorino, G. Gene Expression Signature Predictive of Neuroendocrine Transformation in Prostate Adenocarcinoma. Int. J. Mol. Sci. 2020, 21, 1078. [Google Scholar] [CrossRef] [Green Version]
  341. Coarfa, C.; Fiskus, W.; Eedunuri, V.K.; Rajapakshe, K.; Foley, C.; Chew, S.A.; Shah, S.S.; Geng, C.; Shou, J.; Mohamed, J.S.; et al. Comprehensive proteomic profiling identifies the androgen receptor axis and other signaling pathways as targets of microRNAs suppressed in metastatic prostate cancer. Oncogene 2016, 35, 2345–2356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  342. Fletcher, C.E.; Sulpice, E.; Combe, S.; Shibakawa, A.; Leach, D.A.; Hamilton, M.P.; Chrysostomou, S.L.; Sharp, A.; Welti, J.; Yuan, W.; et al. Androgen receptor-modulatory microRNAs provide insight into therapy resistance and therapeutic targets in advanced prostate cancer. Oncogene 2019, 38, 5700–5724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Ebron, J.S.; Shankar, E.; Singh, J.; Sikand, K.; Weyman, C.M.; Gupta, S.; Lindner, D.J.; Liu, X.; Campbell, M.J.; Shukla, G.C. MiR-644a Disrupts Oncogenic Transformation and Warburg Effect by Direct Modulation of Multiple Genes of Tumor-Promoting Pathways. Cancer Res. 2019, 79, 1844–1856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Qu, H.W.; Jin, Y.; Cui, Z.L.; Jin, X.B. MicroRNA-373-3p inhibits prostate cancer progression by targeting AKT1. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 6252–6259. [Google Scholar] [CrossRef]
  345. Josson, S.; Gururajan, M.; Hu, P.; Shao, C.; Chu, G.Y.; Zhau, H.E.; Liu, C.; Lao, K.; Lu, C.L.; Lu, Y.T.; et al. miR-409-3p/-5p promotes tumorigenesis, epithelial-to-mesenchymal transition, and bone metastasis of human prostate cancer. Clin. Cancer Res. 2014, 20, 4636–4646. [Google Scholar] [CrossRef] [Green Version]
  346. 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]
  347. Yang, K.; Handorean, A.M.; Iczkowski, K.A. MicroRNAs 373 and 520c are downregulated in prostate cancer, suppress CD44 translation and enhance invasion of prostate cancer cells in vitro. Int. J. Clin. Exp. Pathol. 2009, 2, 361–369. [Google Scholar]
  348. Huang, Q.; Ma, B.; Su, Y.; Chan, K.; Qu, H.; Huang, J.; Wang, D.; Qiu, J.; Liu, H.; Yang, X.; et al. miR-197-3p Represses the Proliferation of Prostate Cancer by Regulating the VDAC1/AKT/beta-catenin Signaling Axis. Int. J. Biol. Sci. 2020, 16, 1417–1426. [Google Scholar] [CrossRef]
  349. Yang, J.; Song, Q.; Cai, Y.; Wang, P.; Wang, M.; Zhang, D. RLIP76-dependent suppression of PI3K/AKT/Bcl-2 pathway by miR-101 induces apoptosis in prostate cancer. Biochem. Biophys Res. Commun. 2015, 463, 900–906. [Google Scholar] [CrossRef]
  350. Wu, X.; Bhayani, M.K.; Dodge, C.T.; Nicoloso, M.S.; Chen, Y.; Yan, X.; Adachi, M.; Thomas, L.; Galer, C.E.; Jiffar, T.; et al. Coordinated targeting of the EGFR signaling axis by microRNA-27a *. Oncotarget 2013, 4, 1388–1398. [Google Scholar] [CrossRef] [Green Version]
  351. Fletcher, C.E.; Dart, D.A.; Sita-Lumsden, A.; Cheng, H.; Rennie, P.S.; Bevan, C.L. Androgen-regulated processing of the oncomir miR-27a, which targets Prohibitin in prostate cancer. Hum. Mol. Genet. 2012, 21, 3112–3127. [Google Scholar] [CrossRef] [Green Version]
  352. Wan, X.; Huang, W.; Yang, S.; Zhang, Y.; Zhang, P.; Kong, Z.; Li, T.; Wu, H.; Jing, F.; Li, Y. Androgen-induced miR-27A acted as a tumor suppressor by targeting MAP2K4 and mediated prostate cancer progression. Int. J. Biochem. Cell Biol. 2016, 79, 249–260. [Google Scholar] [CrossRef]
  353. Mo, W.; Zhang, J.; Li, X.; Meng, D.; Gao, Y.; Yang, S.; Wan, X.; Zhou, C.; Guo, F.; Huang, Y.; et al. Identification of Novel AR-Targeted MicroRNAs Mediating Androgen Signalling through Critical Pathways to Regulate Cell Viability in Prostate Cancer. PLoS ONE 2013, 8, e56592. [Google Scholar] [CrossRef] [Green Version]
  354. Barros-Silva, D.; Costa-Pinheiro, P.; Duarte, H.; Sousa, E.J.; Evangelista, A.F.; Graça, I.; Carneiro, I.; Martins, A.T.; Oliveira, J.; Carvalho, A.L.; et al. MicroRNA-27a-5p regulation by promoter methylation and MYC signaling in prostate carcinogenesis. Cell Death Dis. 2018, 9, 167. [Google Scholar] [CrossRef]
  355. Lin, S.-C.; Kao, C.-Y.; Lee, H.-J.; Creighton, C.J.; Ittmann, M.M.; Tsai, S.-J.; Tsai, S.Y.; Tsai, M.-J. Dysregulation of miRNAs-COUP-TFII-FOXM1-CENPF axis contributes to the metastasis of prostate cancer. Nat. Commun. 2016, 7, 11418. [Google Scholar] [CrossRef] [Green Version]
  356. Gao, W.; Hong, Z.; Huang, H.; Zhu, A.; Lin, S.; Cheng, C.; Zhang, X.; Zou, G.; Shi, Z. miR-27a in serum acts as biomarker for prostate cancer detection and promotes cell proliferation by targeting Sprouty2. Oncol. Lett. 2018, 16, 5291–5298. [Google Scholar] [CrossRef]
  357. Ku, A.; Fredsøe, J.; Sørensen, K.D.; Borre, M.; Evander, M.; Laurell, T.; Lilja, H.; Ceder, Y. High-Throughput and Automated Acoustic Trapping of Extracellular Vesicles to Identify microRNAs with Diagnostic Potential for Prostate Cancer. Front. Oncol. 2021, 11, 386. [Google Scholar] [CrossRef]
  358. Nam, R.K.; Wallis, C.J.D.; Amemiya, Y.; Benatar, T.; Seth, A. Identification of a Novel MicroRNA Panel Associated with Metastasis Following Radical Prostatectomy for Prostate Cancer. Anticancer Res. 2018, 38, 5027–5034. [Google Scholar] [CrossRef] [Green Version]
  359. Varambally, S.; Cao, Q.; Mani, R.S.; Shankar, S.; Wang, X.; Ateeq, B.; Laxman, B.; Cao, X.; Jing, X.; Ramnarayanan, K.; et al. Genomic loss of microRNA-101 leads to overexpression of histone methyltransferase EZH2 in cancer. Science 2008, 322, 1695–1699. [Google Scholar] [CrossRef] [Green Version]
  360. Li, K.; Liu, C.; Zhou, B.; Bi, L.; Huang, H.; Lin, T.; Xu, K. Role of EZH2 in the Growth of Prostate Cancer Stem Cells Isolated from LNCaP Cells. Int. J. Mol. Sci. 2013, 14, 11981–11993. [Google Scholar] [CrossRef] [Green Version]
  361. Gu, Z.; You, Z.; Yang, Y.; Ding, R.; Wang, M.; Pu, J.; Chen, J. Inhibition of MicroRNA miR-101-3p on prostate cancer progression by regulating Cullin 4B (CUL4B) and PI3K/AKT/mTOR signaling pathways. Bioengineered 2021, 12, 4719–4735. [Google Scholar] [CrossRef] [PubMed]
  362. Antognelli, C.; Cecchetti, R.; Riuzzi, F.; Peirce, M.J.; Talesa, V.N. Glyoxalase 1 sustains the metastatic phenotype of prostate cancer cells via EMT control. J. Cell. Mol. Med. 2018, 22, 2865–2883. [Google Scholar] [CrossRef] [PubMed]
  363. Hao, Y.; Gu, X.; Zhao, Y.; Greene, S.; Sha, W.; Smoot, D.T.; Califano, J.; Wu, T.C.; Pang, X. Enforced expression of miR-101 inhibits prostate cancer cell growth by modulating the COX-2 pathway in vivo. Cancer Prev. Res. 2011, 4, 1073–1083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Li, P.; You, S.; Nguyen, C.; Wang, Y.; Kim, J.; Sirohi, D.; Ziembiec, A.; Luthringer, D.; Lin, S.C.; Daskivich, T.; et al. Genes involved in prostate cancer progression determine MRI visibility. Theranostics 2018, 8, 1752–1765. [Google Scholar] [CrossRef]
  365. Lin, Y.; Chen, F.; Shen, L.; Tang, X.; Du, C.; Sun, Z.; Ding, H.; Chen, J.; Shen, B. Biomarker microRNAs for prostate cancer metastasis: Screened with a network vulnerability analysis model. J. Transl. Med. 2018, 16, 134. [Google Scholar] [CrossRef] [Green Version]
  366. Watahiki, A.; Wang, Y.; Morris, J.; Dennis, K.; O’Dwyer, H.M.; Gleave, M.; Gout, P.W.; Wang, Y. MicroRNAs Associated with Metastatic Prostate Cancer. PLoS ONE 2011, 6, e24950. [Google Scholar] [CrossRef]
  367. Chen, J.H.; Tong, W.; Pu, X.F.; Wang, J.Z. Long noncoding RNA CRNDE promotes proliferation, migration and invasion in prostate cancer through miR-101/Rap1A. Neoplasma 2020, 67, 584–594. [Google Scholar] [CrossRef]
  368. Cao, P.; Deng, Z.; Wan, M.; Huang, W.; Cramer, S.D.; Xu, J.; Lei, M.; Sui, G. MicroRNA-101 negatively regulates Ezh2 and its expression is modulated by androgen recept or and HIF-1α/HIF-1β. Mol. Cancer 2010, 9, 108. [Google Scholar] [CrossRef] [Green Version]
  369. Ramnarine, V.R.; Alshalalfa, M.; Mo, F.; Nabavi, N.; Erho, N.; Takhar, M.; Shukin, R.; Brahmbhatt, S.; Gawronski, A.; Kobelev, M.; et al. The long noncoding RNA landscape of neuroendocrine prostate cancer and its clinical implications. GigaScience 2018, 7, 1–23. [Google Scholar] [CrossRef] [Green Version]
  370. Luo, J.; Wang, K.; Yeh, S.; Sun, Y.; Liang, L.; Xiao, Y.; Xu, W.; Niu, Y.; Cheng, L.; Maity, S.N.; et al. LncRNA-p21 alters the antiandrogen enzalutamide-induced prostate cancer neuroendocrine differentiation via modulating the EZH2/STAT3 signaling. Nat. Commun. 2019, 10, 2571. [Google Scholar] [CrossRef] [Green Version]
  371. Işın, M.; Uysaler, E.; Özgür, E.; Köseoğlu, H.; Şanlı, Ö.; Yücel, Ö.B.; Gezer, U.; Dalay, N. Exosomal lncRNA-p21 levels may help to distinguish prostate cancer from benign disease. Front. Genet. 2015, 6, 168. [Google Scholar] [CrossRef] [Green Version]
  372. Liu, B.; Jiang, H.-Y.; Yuan, T.; Luo, J.; Zhou, W.-D.; Jiang, Q.-Q.; Wu, D. Enzalutamide-Induced Upregulation of PCAT6 Promotes Prostate Cancer Neuroendocrine Differentiation by Regulating miR-326/HNRNPA2B1 Axis. Front. Oncol. 2021, 11, 1803. [Google Scholar] [CrossRef]
  373. Lang, C.; Yin, C.; Lin, K.; Li, Y.; Yang, Q.; Wu, Z.; Du, H.; Ren, D.; Dai, Y.; Peng, X. m6A modification of lncRNA PCAT6 promotes bone metastasis in prostate cancer through IGF2BP2-mediated IGF1R mRNA stabilization. Clin. Transl. Med. 2021, 11, e426. [Google Scholar] [CrossRef]
  374. Li, Y.; Li, H.; Wei, X. Long noncoding RNA LINC00261 suppresses prostate cancer tumorigenesis through upregulation of GATA6-mediated DKK3. Cancer Cell Int. 2020, 20, 474. [Google Scholar] [CrossRef]
  375. Chang, Y.-T.; Lin, T.-P.; Tang, J.-T.; Campbell, M.; Luo, Y.-L.; Lu, S.-Y.; Yang, C.-P.; Cheng, T.-Y.; Chang, C.-H.; Liu, T.-T.; et al. HOTAIR is a REST-regulated lncRNA that promotes neuroendocrine differentiation in castration resistant prostate cancer. Cancer Lett. 2018, 433, 43–52. [Google Scholar] [CrossRef]
  376. Mather, R.L.; Wang, Y.; Crea, F. Is HOTAIR really involved in neuroendocrine prostate cancer differentiation? Epigenomics 2018, 10, 1259–1261. [Google Scholar] [CrossRef] [Green Version]
  377. Zhang, A.; Zhao, J.C.; Kim, J.; Fong, K.W.; Yang, Y.A.; Chakravarti, D.; Mo, Y.Y.; Yu, J. LncRNA HOTAIR Enhances the Androgen-Receptor-Mediated Transcriptional Program and Drives Castration-Resistant Prostate Cancer. Cell Rep. 2015, 13, 209–221. [Google Scholar] [CrossRef] [Green Version]
  378. Sowalsky, A.G.; Xia, Z.; Wang, L.; Zhao, H.; Chen, S.; Bubley, G.J.; Balk, S.P.; Li, W. Whole transcriptome sequencing reveals extensive unspliced mRNA in metastatic castration-resistant prostate cancer. Mol. Cancer Res. 2015, 13, 98–106. [Google Scholar] [CrossRef] [Green Version]
  379. Wang, R.; Sun, Y.; Li, L.; Niu, Y.; Lin, W.; Lin, C.; Antonarakis, E.S.; Luo, J.; Yeh, S.; Chang, C. Preclinical Study using Malat1 Small Interfering RNA or Androgen Receptor. Splicing Variant 7 Degradation Enhancer ASC-J9(®) to Suppress Enzalutamide-resistant Prostate Cancer Progression. Eur. Urol. 2017, 72, 835–844. [Google Scholar] [CrossRef] [Green Version]
  380. Cai, S.; Pataillot-Meakin, T.; Shibakawa, A.; Ren, R.; Bevan, C.L.; Ladame, S.; Ivanov, A.P.; Edel, J.B. Single-molecule amplification-free multiplexed detection of circulating microRNA cancer biomarkers from serum. Nat. Commun. 2021, 12, 3515. [Google Scholar] [CrossRef]
  381. Lakshmanan, V.K.; Ojha, S.; Jung, Y.D. A modern era of personalized medicine in the diagnosis, prognosis, and treatment of prostate cancer. Comput. Biol. Med. 2020, 126, 104020. [Google Scholar] [CrossRef] [PubMed]
  382. Ahadi, A.; Brennan, S.; Kennedy, P.J.; Hutvagner, G.; Tran, N. Long non-coding RNAs harboring miRNA seed regions are enriched in prostate cancer exosomes. Sci. Rep. 2016, 6, 24922. [Google Scholar] [CrossRef] [PubMed]
  383. Reda El Sayed, S.; Cristante, J.; Guyon, L.; Denis, J.; Chabre, O.; Cherradi, N. MicroRNA Therapeutics in Cancer: Current Advances and Challenges. Cancers 2021, 13, 2680. [Google Scholar] [CrossRef] [PubMed]
  384. Velagapudi, S.P.; Luo, Y.; Tran, T.; Haniff, H.S.; Nakai, Y.; Fallahi, M.; Martinez, G.J.; Childs-Disney, J.L.; Disney, M.D. Defining RNA-Small Molecule Affinity Landscapes Enables Design of a Small Molecule Inhibitor of an Oncogenic Noncoding RNA. ACS Cent. Sci. 2017, 3, 205–216. [Google Scholar] [CrossRef]
  385. Liu, X.; Haniff, H.S.; Childs-Disney, J.L.; Shuster, A.; Aikawa, H.; Adibekian, A.; Disney, M.D. Targeted Degradation of the Oncogenic MicroRNA 17-92 Cluster by Structure-Targeting Ligands. J. Am. Chem. Soc. 2020, 142, 6970–6982. [Google Scholar] [CrossRef]
  386. Ottman, R.; Levy, J.; Grizzle, W.E.; Chakrabarti, R. The other face of miR-17-92a cluster, exhibiting tumor suppressor effects in prostate cancer. Oncotarget 2016, 7, 73739–73753. [Google Scholar] [CrossRef] [Green Version]
  387. Zhou, P.; Ma, L.; Zhou, J.; Jiang, M.; Rao, E.; Zhao, Y.; Guo, F. miR-17-92 plays an oncogenic role and conveys chemo-resistance to cisplatin in human prostate cancer cells. Int. J. Oncol. 2016, 48, 1737–1748. [Google Scholar] [CrossRef] [Green Version]
  388. 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]
  389. Zhang, X.; Ladd, A.; Dragoescu, E.; Budd, W.T.; Ware, J.L.; Zehner, Z.E. MicroRNA-17-3p is a prostate tumor suppressor in vitro and in vivo, and is decreased in high grade prostate tumors analyzed by laser capture microdissection. Clin. Exp. Metastasis 2009, 26, 965–979. [Google Scholar] [CrossRef]
  390. Dai, H.; Wang, C.; Yu, Z.; He, D.; Yu, K.; Liu, Y.; Wang, S. MiR-17 Regulates Prostate Cancer Cell Proliferation and Apoptosis Through Inhibiting JAK-STAT3 Signaling Pathway. Cancer Biother. Radiopharm. 2018, 33, 103–109. [Google Scholar] [CrossRef]
  391. Xu, Z.; Zhang, Y.; Ding, J.; Hu, W.; Tan, C.; Wang, M.; Tang, J.; Xu, Y. miR-17-3p Downregulates Mitochondrial Antioxidant Enzymes and Enhances the Radiosensitivity of Prostate Cancer Cells. Mol. Ther. Nucleic Acids 2018, 13, 64–77. [Google Scholar] [CrossRef] [Green Version]
  392. Yang, X.; Du, W.W.; Li, H.; Liu, F.; Khorshidi, A.; Rutnam, Z.J.; Yang, B.B. Both mature miR-17-5p and passenger strand miR-17-3p target TIMP3 and induce prostate tumor growth and invasion. Nucleic Acids Res. 2013, 41, 9688–9704. [Google Scholar] [CrossRef] [Green Version]
  393. Gong, A.Y.; Eischeid, A.N.; Xiao, J.; Zhao, J.; Chen, D.; Wang, Z.Y.; Young, C.Y.; Chen, X.M. miR-17-5p targets the p300/CBP-associated factor and modulates androgen receptor transcriptional activity in cultured prostate cancer cells. BMC Cancer 2012, 12, 492. [Google Scholar] [CrossRef] [Green Version]
  394. Dyson, G.; Farran, B.; Bolton, S.; Craig, D.B.; Dombkowski, A.; Beebe-Dimmer, J.L.; Powell, I.J.; Podgorski, I.; Heilbrun, L.K.; Bock, C.H. The extrema of circulating miR-17 are identified as biomarkers for aggressive prostate cancer. Am. J. Cancer Res. 2018, 8, 2088–2095. [Google Scholar]
  395. Urabe, F.; Matsuzaki, J.; Yamamoto, Y.; Kimura, T.; Hara, T.; Ichikawa, M.; Takizawa, S.; Aoki, Y.; Niida, S.; Sakamoto, H.; et al. Large-scale Circulating microRNA Profiling for the Liquid Biopsy of Prostate Cancer. Clin. Cancer Res. 2019, 25, 3016–3025. [Google Scholar] [CrossRef] [Green Version]
  396. Wang, X.; Wang, R.; Wu, Z.; Bai, P. Circular RNA ITCH suppressed prostate cancer progression by increasing HOXB13 expression via spongy miR-17-5p. Cancer Cell Int. 2019, 19, 328. [Google Scholar] [CrossRef]
  397. Hu, Y.; Guo, B. Circ-MTO1 correlates with favorable prognosis and inhibits cell proliferation, invasion as well as miR-17-5p expression in prostate cancer. J. Clin. Lab. Anal. 2020, 34, e23086. [Google Scholar] [CrossRef] [Green Version]
  398. Xu, Y.; Fang, F.; Zhang, J.; Josson, S.; St Clair, W.H.; St Clair, D.K. miR-17* suppresses tumorigenicity of prostate cancer by inhibiting mitochondrial antioxidant enzymes. PLoS ONE 2010, 5, e14356. [Google Scholar] [CrossRef]
  399. Liang, B.; Zhou, C.; Cui, S.; Lu, H.; Xu, R.; Xue, D.; Zou, S.; He, X. Upregulation of miR-18a-5p promotes the proliferation of prostate cancer via inhibiting the expression of SLC40A1. Pathol.-Res. Pract. 2021, 224, 153448. [Google Scholar] [CrossRef]
  400. Hsu, T.I.; Hsu, C.H.; Lee, K.H.; Lin, J.T.; Chen, C.S.; Chang, K.C.; Su, C.Y.; Hsiao, M.; Lu, P.J. MicroRNA-18a is elevated in prostate cancer and promotes tumorigenesis through suppressing STK4 in vitro and in vivo. Oncogenesis 2014, 3, e99. [Google Scholar] [CrossRef] [Green Version]
  401. Ibrahim, N.H.; Abdellateif, M.S.; Kassem, S.H.; Abd El Salam, M.A.; El Gammal, M.M. Diagnostic significance of miR-21, miR-141, miR-18a and miR-221 as novel biomarkers in prostate cancer among Egyptian patients. Andrologia 2019, 51, e13384. [Google Scholar] [CrossRef] [PubMed]
  402. Al-Kafaji, G.; Al-Naieb, Z.T.; Bakhiet, M. Increased oncogenic microRNA-18a expression in the peripheral blood of patients with prostate cancer: A potential novel non-invasive biomarker. Oncol. Lett. 2016, 11, 1201–1206. [Google Scholar] [CrossRef] [PubMed]
  403. Yang, J.; Hao, T.; Sun, J.; Wei, P.; Zhang, H. Long noncoding RNA GAS5 modulates α-Solanine-induced radiosensitivity by negatively regulating miR-18a in human prostate cancer cells. Biomed. Pharm. 2019, 112, 108656. [Google Scholar] [CrossRef] [PubMed]
  404. Zhang, G.; Han, G.; Zhang, X.; Yu, Q.; Li, Z.; Li, Z.; Li, J. Long non-coding RNA FENDRR reduces prostate cancer malignancy by competitively binding miR-18a-5p with RUNX1. Biomarkers 2018, 23, 435–445. [Google Scholar] [CrossRef] [PubMed]
  405. Zhao, Y.; Zhang, Q.; Liu, H.; Wang, N.; Zhang, X.; Yang, S. lncRNA PART1, manipulated by transcriptional factor FOXP2, suppresses proliferation and invasion in ESCC by regulating the miR-18a-5p/SOX6 signaling axis. Oncol. Rep. 2021, 45, 1118–1132. [Google Scholar] [CrossRef] [PubMed]
  406. Feng, S.; Zhu, X.; Fan, B.; Xie, D.; Li, T.; Zhang, X. miR-19a-3p targets PMEPA1 and induces prostate cancer cell proliferation, migration and invasion. Mol. Med. Rep. 2016, 13, 4030–4038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  407. Wa, Q.; Li, L.; Lin, H.; Peng, X.; Ren, D.; Huang, Y.; He, P.; Huang, S. Downregulation of miR-19a-3p promotes invasion, migration and bone metastasis via activating TGF-β signaling in prostate cancer. Oncol. Rep. 2018, 39, 81–90. [Google Scholar] [CrossRef] [Green Version]
  408. Fu, F.; Wan, X.; Wang, D.; Kong, Z.; Zhang, Y.; Huang, W.; Wang, C.; Wu, H.; Li, Y. MicroRNA-19a acts as a prognostic marker and promotes prostate cancer progression via inhibiting VPS37A expression. Oncotarget 2017, 9, 1931. [Google Scholar] [CrossRef] [Green Version]
  409. Lu, K.; Liu, C.; Tao, T.; Zhang, X.; Zhang, L.; Sun, C.; Wang, Y.; Chen, S.; Xu, B.; Chen, M. MicroRNA-19a regulates proliferation and apoptosis of castration-resistant prostate cancer cells by targeting BTG1. FEBS Lett. 2015, 589, 1485–1490. [Google Scholar] [CrossRef] [Green Version]
  410. Wang, S.-Y.; Shiboski, S.; Belair, C.D.; Cooperberg, M.R.; Simko, J.P.; Stoppler, H.; Cowan, J.; Carroll, P.R.; Blelloch, R. miR-19, miR-345, miR-519c-5p Serum Levels Predict Adverse Pathology in Prostate Cancer Patients Eligible for Active Surveillance. PLoS ONE 2014, 9, e98597. [Google Scholar] [CrossRef]
  411. Qiang, X.F.; Zhang, Z.W.; Liu, Q.; Sun, N.; Pan, L.L.; Shen, J.; Li, T.; Yun, C.; Li, H.; Shi, L.H. miR-20a promotes prostate cancer invasion and migration through targeting ABL2. J. Cell Biochem. 2014, 115, 1269–1276. [Google Scholar] [CrossRef]
  412. Pesta, M.; Klecka, J.; Kulda, V.; Topolcan, O.; Hora, M.; Eret, V.; Ludvikova, M.; Babjuk, M.; Novak, K.; Stolz, J.; et al. Importance of miR-20a expression in prostate cancer tissue. Anticancer Res. 2010, 30, 3579–3583. [Google Scholar]
  413. 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]
  414. Hart, M.; Nolte, E.; Wach, S.; Szczyrba, J.; Taubert, H.; Rau, T.T.; Hartmann, A.; Grasser, F.A.; Wullich, B. Comparative microRNA profiling of prostate carcinomas with increasing tumor stage by deep sequencing. Mol. Cancer Res. 2014, 12, 250–263. [Google Scholar] [CrossRef] [Green Version]
  415. Li, X.; Pan, J.H.; Song, B.; Xiong, E.Q.; Chen, Z.W.; Zhou, Z.S.; Su, Y.P. Suppression of CX43 expression by miR-20a in the progression of human prostate cancer. Cancer Biol. Ther. 2012, 13, 890–898. [Google Scholar] [CrossRef] [Green Version]
  416. Sylvestre, Y.; De Guire, V.; Querido, E.; Mukhopadhyay, U.K.; Bourdeau, V.; Major, F.; Ferbeyre, G.; Chartrand, P. An E2F/miR-20a autoregulatory feedback loop. J. Biol. Chem. 2007, 282, 2135–2143. [Google Scholar] [CrossRef] [Green Version]
  417. Lin, H.M.; Castillo, L.; Mahon, K.L.; Chiam, K.; Lee, B.Y.; Nguyen, Q.; Boyer, M.J.; Stockler, M.R.; Pavlakis, N.; Marx, G.; et al. Circulating microRNAs are associated with docetaxel chemotherapy outcome in castration-resistant prostate cancer. Br. J. Cancer 2014, 110, 2462–2471. [Google Scholar] [CrossRef]
  418. Mohammadi Torbati, P.; Asadi, F.; Fard-Esfahani, P. Circulating miR-20a and miR-26a as Biomarkers in Prostate Cancer. Asian Pac. J. Cancer Prev. 2019, 20, 1453–1456. [Google Scholar] [CrossRef] [Green Version]
  419. Daniel, R.; Wu, Q.; Williams, V.; Clark, G.; Guruli, G.; Zehner, Z. A Panel of MicroRNAs as Diagnostic Biomarkers for the Identification of Prostate Cancer. Int. J. Mol. Sci. 2017, 18, 1281. [Google Scholar] [CrossRef] [Green Version]
  420. Li, J.; Yang, X.; Guan, H.; Mizokami, A.; Keller, E.T.; Xu, X.; Liu, X.; Tan, J.; Hu, L.; Lu, Y.; et al. Exosome-derived microRNAs contribute to prostate cancer chemoresistance. Int. J. Oncol. 2016, 49, 838–846. [Google Scholar] [CrossRef] [Green Version]
  421. Ambrozkiewicz, F.; Karczmarski, J.; Kulecka, M.; Paziewska, A.; Cybulska, M.; Szymanski, M.; Dobruch, J.; Antoniewicz, A.; Mikula, M.; Ostrowski, J. Challenges in Cancer Biomarker Discovery Exemplified by the Identification of Diagnostic MicroRNAs in Prostate Tissues. Biomed. Res. Int. 2020, 2020, 9086829. [Google Scholar] [CrossRef] [PubMed]
  422. Paziewska, A.; Mikula, M.; Dabrowska, M.; Kulecka, M.; Goryca, K.; Antoniewicz, A.; Dobruch, J.; Borowka, A.; Rutkowski, P.; Ostrowski, J. Candidate diagnostic miRNAs that can detect cancer in prostate biopsy. Prostate 2018, 78, 178–185. [Google Scholar] [CrossRef] [PubMed]
  423. Wu, G.; Wang, J.; Chen, G.; Zhao, X. microRNA-204 modulates chemosensitivity and apoptosis of prostate cancer cells by targeting zinc-finger E-box-binding homeobox 1 (ZEB1). Am. J. Transl. Res. 2017, 9, 3599–3610. [Google Scholar] [PubMed]
  424. Lin, Y.C.; Lin, J.F.; Tsai, T.F.; Chou, K.Y.; Chen, H.E.; Hwang, T.I. Tumor suppressor miRNA-204-5p promotes apoptosis by targeting BCL2 in prostate cancer cells. Asian J. Surg. 2017, 40, 396–406. [Google Scholar] [CrossRef] [Green Version]
  425. Fredsoe, J.; Rasmussen, A.K.I.; Mouritzen, P.; Borre, M.; Orntoft, T.; Sorensen, K.D. A five-microRNA model (pCaP) for predicting prostate cancer aggressiveness using cell-free urine. Int. J. Cancer 2019, 145, 2558–2567. [Google Scholar] [CrossRef]
  426. Koppers-Lalic, D.; Hackenberg, M.; de Menezes, R.; Misovic, B.; Wachalska, M.; Geldof, A.; Zini, N.; de Reijke, T.; Wurdinger, T.; Vis, A.; et al. Noninvasive prostate cancer detection by measuring miRNA variants (isomiRs) in urine extracellular vesicles. Oncotarget 2016, 7, 22566–22578. [Google Scholar] [CrossRef] [Green Version]
  427. He, C.; Lu, X.; Yang, F.; Qin, L.; Guo, Z.; Sun, Y.; Wu, J. LncRNA UCA1 acts as a sponge of miR-204 to up-regulate CXCR4 expression and promote prostate cancer progression. Biosci. Rep. 2019, 39, BSR20181465. [Google Scholar] [CrossRef] [Green Version]
  428. Zhang, S.; Dong, X.; Ji, T.; Chen, G.; Shan, L. Long non-coding RNA UCA1 promotes cell progression by acting as a competing endogenous RNA of ATF2 in prostate cancer. Am. J. Transl. Res. 2017, 9, 366–375. [Google Scholar]
  429. Todorova, K.; Metodiev, M.V.; Metodieva, G.; Mincheff, M.; Fernandez, N.; Hayrabedyan, S. Micro-RNA-204 Participates in TMPRSS2/ERG Regulation and Androgen Receptor Reprogramming in Prostate Cancer. Horm. Cancer 2017, 8, 28–48. [Google Scholar] [CrossRef]
  430. Panigrahi, G.K.; Ramteke, A.; Birks, D.; Abouzeid Ali, H.E.; Venkataraman, S.; Agarwal, C.; Vibhakar, R.; Miller, L.D.; Agarwal, R.; Abd Elmageed, Z.Y.; et al. Exosomal microRNA profiling to identify hypoxia-related biomarkers in prostate cancer. Oncotarget 2018, 9, 13894–13910. [Google Scholar] [CrossRef] [Green Version]
  431. Zhang, G.; Tian, X.; Li, Y.; Wang, Z.; Li, X.; Zhu, C. miR-27b and miR-34a enhance docetaxel sensitivity of prostate cancer cells through inhibiting epithelial-to-mesenchymal transition by targeting ZEB1. Biomed. Pharmacother. 2018, 97, 736–744. [Google Scholar] [CrossRef]
  432. Rokavec, M.; Oner, M.G.; Li, H.; Jackstadt, R.; Jiang, L.; Lodygin, D.; Kaller, M.; Horst, D.; Ziegler, P.K.; Schwitalla, S.; et al. IL-6R/STAT3/miR-34a feedback loop promotes EMT-mediated colorectal cancer invasion and metastasis. J. Clin. Investig. 2014, 124, 1853–1867. [Google Scholar] [CrossRef] [Green Version]
  433. Duan, K.; Ge, Y.C.; Zhang, X.P.; Wu, S.Y.; Feng, J.S.; Chen, S.L.; Zhang, L.I.; Yuan, Z.H.; Fu, C.H. miR-34a inhibits cell proliferation in prostate cancer by downregulation of SIRT1 expression. Oncol. Lett. 2015, 10, 3223–3227. [Google Scholar] [CrossRef] [Green Version]
  434. Liang, J.; Li, Y.; Daniels, G.; Sfanos, K.; De Marzo, A.; Wei, J.; Li, X.; Chen, W.; Wang, J.; Zhong, X.; et al. LEF1 Targeting EMT in Prostate Cancer Invasion Is Regulated by miR-34a. Mol. Cancer Res. 2015, 13, 681–688. [Google Scholar] [CrossRef] [Green Version]
  435. Liao, H.; Xiao, Y.; Hu, Y.; Yin, Z.; Liu, L.; Kang, X.; Chen, Y. Methylation-induced silencing of miR-34a enhances chemoresistance by directly upregulating ATG4B-induced autophagy through AMPK/mTOR pathway in prostate cancer. Oncol. Rep. 2016, 35, 64–72. [Google Scholar] [CrossRef] [Green Version]
  436. Liu, X.; Luo, X.; Wu, Y.; Xia, D.; Chen, W.; Fang, Z.; Deng, J.; Hao, Y.; Yang, X.; Zhang, T.; et al. MicroRNA-34a Attenuates Paclitaxel Resistance in Prostate Cancer Cells via Direct Suppression of JAG1/Notch1 Axis. Cell Physiol. Biochem. 2018, 50, 261–276. [Google Scholar] [CrossRef]
  437. Corcoran, C.; Rani, S.; O’Driscoll, L. miR-34a is an intracellular and exosomal predictive biomarker for response to docetaxel with clinical relevance to prostate cancer progression. Prostate 2014, 74, 1320–1334. [Google Scholar] [CrossRef] [Green Version]
  438. Ma, Y.; Fan, B.; Ren, Z.; Liu, B.; Wang, Y. Long noncoding RNA DANCR contributes to docetaxel resistance in prostate cancer through targeting the miR-34a-5p/JAG1 pathway. Onco Targets Ther. 2019, 12, 5485–5497. [Google Scholar] [CrossRef] [Green Version]
  439. Ma, E.; Wang, Q.; Li, J.; Zhang, X.; Guo, Z.; Yang, X. LINC01006 facilitates cell proliferation, migration and invasion in prostate cancer through targeting miR-34a-5p to up-regulate DAAM1. Cancer Cell. Int. 2020, 20, 515. [Google Scholar] [CrossRef]
  440. Li, N.; Zhang, L.Y.; Qiao, Y.H.; Song, R.J. Long noncoding RNA LINC00662 functions as miRNA sponge to promote the prostate cancer tumorigenesis through targeting miR-34a. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 3688–3698. [Google Scholar] [CrossRef]
  441. Lodygin, D.; Tarasov, V.; Epanchintsev, A.; Berking, C.; Knyazeva, T.; Körner, H.; Knyazev, P.; Diebold, J.; Hermeking, H. Inactivation of miR-34a by aberrant CpG methylation in multiple types of cancer. Cell Cycle 2008, 7, 2591–2600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  442. Krebs, M.; Behrmann, C.; Kalogirou, C.; Sokolakis, I.; Kneitz, S.; Kruithof-de Julio, M.; Zoni, E.; Rech, A.; Schilling, B.; Kubler, H.; et al. miR-221 Augments TRAIL-Mediated Apoptosis in Prostate Cancer Cells by Inducing Endogenous TRAIL Expression and Targeting the Functional Repressors SOCS3 and PIK3R1. Biomed. Res. Int. 2019, 2019, 6392748. [Google Scholar] [CrossRef] [PubMed]
  443. Kurul, N.O.; Ates, F.; Yilmaz, I.; Narli, G.; Yesildal, C.; Senkul, T. The association of let-7c, miR-21, miR-145, miR-182, and miR-221 with clinicopathologic parameters of prostate cancer in patients diagnosed with low-risk disease. Prostate 2019, 79, 1125–1132. [Google Scholar] [CrossRef] [PubMed]
  444. Maqbool, R.; Lone, S.N.; Ul Hussain, M. Post-transcriptional regulation of the tumor suppressor p53 by a novel miR-27a, with implications during hypoxia and tumorigenesis. Biochem. J. 2016, 473, 3597–3610. [Google Scholar] [CrossRef]
  445. Cao, Z.; Xu, L.; Zhao, S. Exosome-derived miR-27a produced by PSC-27 cells contributes to prostate cancer chemoresistance through p53. Biochem. Biophys Res. Commun. 2019, 515, 345–351. [Google Scholar] [CrossRef]
  446. Lyu, J.; Zhao, L.; Wang, F.; Ji, J.; Cao, Z.; Xu, H.; Shi, X.; Zhu, Y.; Zhang, C.; Guo, F.; et al. Discovery and Validation of Serum MicroRNAs as Early Diagnostic Biomarkers for Prostate Cancer in Chinese Population. Biomed. Res. Int. 2019, 2019, 9306803. [Google Scholar] [CrossRef] [Green Version]
  447. Sun, F.; Wu, K.; Yao, Z.; Mu, X.; Zheng, Z.; Sun, M.; Wang, Y.; Liu, Z.; Zhu, Y. Long Noncoding RNA PVT1 Promotes Prostate Cancer Metastasis by Increasing NOP2 Expression via Targeting Tumor Suppressor MicroRNAs. Onco Targets Ther. 2020, 13, 6755–6765. [Google Scholar] [CrossRef]
  448. Cui, X.; Piao, C.; Lv, C.; Lin, X.; Zhang, Z.; Liu, X. ZNFX1 anti-sense RNA 1 promotes the tumorigenesis of prostate cancer by regulating c-Myc expression via a regulatory network of competing endogenous RNAs. Cell Mol. Life Sci. 2020, 77, 1135–1152. [Google Scholar] [CrossRef]
  449. Wu, J.; Zheng, C.; Fan, Y.; Zeng, C.; Chen, Z.; Qin, W.; Zhang, C.; Zhang, W.; Wang, X.; Zhu, X.; et al. Downregulation of microRNA-30 facilitates podocyte injury and is prevented by glucocorticoids. J. Am. Soc. Nephrol. 2014, 25, 92–104. [Google Scholar] [CrossRef] [Green Version]
  450. Ren, Q.; Liang, J.; Wei, J.; Basturk, O.; Wang, J.; Daniels, G.; Gellert, L.L.; Li, Y.; Shen, Y.; Osman, I.; et al. Epithelial and stromal expression of miRNAs during prostate cancer progression. Am. J. Transl. Res. 2014, 6, 329–339. [Google Scholar]
  451. 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] [Green Version]
  452. Zhao, Z.; Weickmann, S.; Jung, M.; Lein, M.; Kilic, E.; Stephan, C.; Erbersdobler, A.; Fendler, A.; Jung, K. A Novel Predictor Tool of Biochemical Recurrence after Radical Prostatectomy Based on a Five-MicroRNA Tissue Signature. Cancers 2019, 11, 1603. [Google Scholar] [CrossRef] [Green Version]
  453. Chen, W.; Yao, G.; Zhou, K. miR-103a-2-5p/miR-30c-1-3p inhibits the progression of prostate cancer resistance to androgen ablation therapy via targeting androgen receptor variant 7. J. Cell Biochem. 2019, 120, 14055–14064. [Google Scholar] [CrossRef]
  454. Song, Y.; Song, C.; Yang, S. Tumor-Suppressive Function of miR-30d-5p in Prostate Cancer Cell Proliferation and Migration by Targeting NT5E. Cancer Biother. Radiopharm. 2018, 33, 203–211. [Google Scholar] [CrossRef]
  455. Fuse, M.; Kojima, S.; Enokida, H.; Chiyomaru, T.; Yoshino, H.; Nohata, N.; Kinoshita, T.; Sakamoto, S.; Naya, Y.; Nakagawa, M.; et al. Tumor suppressive microRNAs (miR-222 and miR-31) regulate molecular pathways based on microRNA expression signature in prostate cancer. J. Hum. Genet. 2012, 57, 691–699. [Google Scholar] [CrossRef] [Green Version]
  456. Tsuchiyama, K.; Ito, H.; Taga, M.; Naganuma, S.; Oshinoya, Y.; Nagano, K.; Yokoyama, O.; Itoh, H. Expression of microRNAs associated with Gleason grading system in prostate cancer: miR-182-5p is a useful marker for high grade prostate cancer. Prostate 2013, 73, 827–834. [Google Scholar] [CrossRef]
  457. Bian, X.; Shen, Y.; Zhang, G.; Gu, C.; Cai, Y.; Wang, C.; Zhu, Y.; Zhu, Y.; Zhang, H.; Dai, B.; et al. Expression of Dicer and Its Related MiRNAs in the Progression of Prostate Cancer. PLoS ONE 2015, 10, e0120159. [Google Scholar] [CrossRef] [Green Version]
  458. Zhao, J.; Xu, H.; Duan, Z.; Chen, X.; Ao, Z.; Chen, Y.; Ruan, Y.; Ni, M. miR-31-5p Regulates 14-3-3 ɛ to Inhibit Prostate Cancer 22RV1 Cell Survival and Proliferation via PI3K/AKT/Bcl-2 Signaling Pathway. Cancer Manag. Res. 2020, 12, 6679–6694. [Google Scholar] [CrossRef]
  459. Bhatnagar, N.; Li, X.; Padi, S.K.R.; Zhang, Q.; Tang, M.s.; Guo, B. Downregulation of miR-205 and miR-31 confers resistance to chemotherapy-induced apoptosis in prostate cancer cells. Cell Death. Dis. 2010, 1, e105. [Google Scholar] [CrossRef] [Green Version]
  460. Daniunaite, K.; Dubikaityte, M.; Gibas, P.; Bakavicius, A.; Rimantas Lazutka, J.; Ulys, A.; Jankevicius, F.; Jarmalaite, S. Clinical significance of miRNA host gene promoter methylation in prostate cancer. Hum. Mol. Genet. 2017, 26, 2451–2461. [Google Scholar] [CrossRef]
  461. Lekchnov, E.A.; Amelina, E.V.; Bryzgunova, O.E.; Zaporozhchenko, I.A.; Konoshenko, M.Y.; Yarmoschuk, S.V.; Murashov, I.S.; Pashkovskaya, O.A.; Gorizkii, A.M.; Zheravin, A.A.; et al. Searching for the Novel Specific Predictors of Prostate Cancer in Urine: The Analysis of 84 miRNA Expression. Int. J. Mol. Sci. 2018, 19, 4088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  462. Liu, B.; Sun, Y.; Tang, M.; Liang, C.; Huang, C.P.; Niu, Y.; Wang, Z.; Chang, C. The miR-361-3p increases enzalutamide (Enz) sensitivity via targeting the ARv7 and MKNK2 to better suppress the Enz-resistant prostate cancer. Cell Death Dis. 2020, 11, 807. [Google Scholar] [CrossRef] [PubMed]
  463. Liu, D.; Tao, T.; Xu, B.; Chen, S.; Liu, C.; Zhang, L.; Lu, K.; Huang, Y.; Jiang, L.; Zhang, X.; et al. MiR-361-5p acts as a tumor suppressor in prostate cancer by targeting signal transducer and activator of transcription-6(STAT6). Biochem. Biophys Res. Commun. 2014, 445, 151–156. [Google Scholar] [CrossRef] [PubMed]
  464. Zhu, J.; Wang, S.; Zhang, W.; Qiu, J.; Shan, Y.; Yang, D.; Shen, B. Screening key microRNAs for castration-resistant prostate cancer based on miRNA/mRNA functional synergistic network. Oncotarget 2015, 6, 43819–43830. [Google Scholar] [CrossRef] [PubMed]
  465. Ju, G.; Zhu, Y.; Du, T.; Cao, W.; Lin, J.; Li, C.; Xu, D.; Wang, Z. MiR-197 Inhibitor Loaded AbCD133@MSNs@GNR Affects the Development of Prostate Cancer Through Targeting ITGAV. Front. Cell Dev. Biol. 2021, 9, 1289. [Google Scholar] [CrossRef]
  466. Walter, B.A.; Valera, V.A.; Pinto, P.A.; Merino, M.J. Comprehensive microRNA Profiling of Prostate Cancer. J. Cancer 2013, 4, 350–357. [Google Scholar] [CrossRef] [Green Version]
  467. Qiu, X.; Zhu, J.; Sun, Y.; Fan, K.; Yang, D.R.; Li, G.; Yang, G.; Chang, C. TR4 nuclear receptor increases prostate cancer invasion via decreasing the miR-373-3p expression to alter TGFβR2/p-Smad3 signals. Oncotarget 2015, 6, 15397–15409. [Google Scholar] [CrossRef] [Green Version]
  468. Zhang, G.; Liu, Z.; Xu, H.; Yang, Q. miR-409-3p suppresses breast cancer cell growth and invasion by targeting Akt1. Biochem. Biophys. Res. Commun. 2016, 469, 189–195. [Google Scholar] [CrossRef]
  469. Nguyen, H.C.; Xie, W.; Yang, M.; Hsieh, C.L.; Drouin, S.; Lee, G.S.; Kantoff, P.W. Expression differences of circulating microRNAs in metastatic castration resistant prostate cancer and low-risk, localized prostate cancer. Prostate 2013, 73, 346–354. [Google Scholar] [CrossRef] [Green Version]
  470. Yu, Q.; Li, P.; Weng, M.; Wu, S.; Zhang, Y.; Chen, X.; Zhang, Q.; Shen, G.; Ding, X.; Fu, S. Nano-Vesicles are a Potential Tool to Monitor Therapeutic Efficacy of Carbon Ion Radiotherapy in Prostate Cancer. J. Biomed. Nanotechnol. 2018, 14, 168–178. [Google Scholar] [CrossRef]
  471. Xiang, S.; Zou, P.; Tang, Q.; Zheng, F.; Wu, J.; Chen, Z.; Hann, S.S. HOTAIR-mediated reciprocal regulation of EZH2 and DNMT1 contribute to polyphyllin I-inhibited growth of castration-resistant prostate cancer cells in vitro and in vivo. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2018, 1862, 589–599. [Google Scholar] [CrossRef]
  472. Li, T.; Liu, N.; Gao, Y.; Quan, Z.; Hao, Y.; Yu, C.; Li, L.; Yuan, M.; Niu, L.; Luo, C.; et al. Long noncoding RNA HOTAIR regulates the invasion and metastasis of prostate cancer by targeting hepaCAM. Br. J. Cancer 2021, 124, 247–258. [Google Scholar] [CrossRef]
  473. Wang, N.; Jiang, Y.; Lv, S.; Wen, H.; Wu, D.; Wei, Q.; Dang, Q. HOTAIR expands the population of prostatic cancer stem-like cells and causes Docetaxel resistance via activating STAT3 signaling. Aging (Albany NY) 2020, 12, 12771–12782. [Google Scholar] [CrossRef]
  474. Ling, Z.; Wang, X.; Tao, T.; Zhang, L.; Guan, H.; You, Z.; Lu, K.; Zhang, G.; Chen, S.; Wu, J.; et al. Involvement of aberrantly activated HOTAIR/EZH2/miR-193a feedback loop in progression of prostate cancer. J. Exp. Clin. Cancer Res. 2017, 36, 159. [Google Scholar] [CrossRef] [Green Version]
  475. Wang, D.; Ding, L.; Wang, L.; Zhao, Y.; Sun, Z.; Karnes, R.J.; Zhang, J.; Huang, H. LncRNA MALAT1 enhances oncogenic activities of EZH2 in castration-resistant prostate cancer. Oncotarget 2015, 6, 41045. [Google Scholar] [CrossRef]
  476. Hao, T.; Wang, Z.; Yang, J.; Zhang, Y.; Shang, Y.; Sun, J. MALAT1 knockdown inhibits prostate cancer progression by regulating miR-140/BIRC6 axis. Biomed. Pharmacother. 2020, 123, 109666. [Google Scholar] [CrossRef]
  477. Chang, J.; Xu, W.; Du, X.; Hou, J. MALAT1 silencing suppresses prostate cancer progression by upregulating miR-1 and downregulating KRAS. Onco Targets Ther. 2018, 11, 3461–3473. [Google Scholar] [CrossRef] [Green Version]
  478. Xue, D.; Lu, H.; Xu, H.-Y.; Zhou, C.-X.; He, X.-Z. Long noncoding RNA MALAT1 enhances the docetaxel resistance of prostate cancer cells via miR-145-5p-mediated regulation of AKAP12. J. Cell. Mol. Med. 2018, 22, 3223–3237. [Google Scholar] [CrossRef]
  479. Li, Y.; Ji, J.; Lyu, J.; Jin, X.; He, X.; Mo, S.; Xu, H.; He, J.; Cao, Z.; Chen, X.; et al. A Novel Urine Exosomal lncRNA Assay to Improve the Detection of Prostate Cancer at Initial Biopsy: A Retrospective Multicenter Diagnostic Feasibility Study. Cancers 2021, 13, 4075. [Google Scholar] [CrossRef]
  480. Wang, F.; Ren, S.; Chen, R.; Lu, J.; Shi, X.; Zhu, Y.; Zhang, W.; Jing, T.; Zhang, C.; Shen, J.; et al. Development and prospective multicenter evaluation of the long noncoding RNA MALAT-1 as a diagnostic urinary biomarker for prostate cancer. Oncotarget 2014, 5, 11091. [Google Scholar] [CrossRef] [Green Version]
  481. Ren, S.; Wang, F.; Shen, J.; Sun, Y.; Xu, W.; Lu, J.; Wei, M.; Xu, C.; Wu, C.; Zhang, Z.; et al. Long non-coding RNA metastasis associated in lung adenocarcinoma transcript 1 derived miniRNA as a novel plasma-based biomarker for diagnosing prostate cancer. Eur. J. Cancer 2013, 49, 2949–2959. [Google Scholar] [CrossRef] [PubMed]
  482. Dai, X.; Liang, Z.; Liu, L.; Guo, K.; Xu, S.; Wang, H. Silencing of MALAT1 inhibits migration and invasion by sponging miR-1-3p in prostate cancer cells. Mol. Med. Rep. 2019, 20, 3499–3508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  483. Eke, I.; Bylicky, M.A.; Sandfort, V.; Chopra, S.; Martello, S.; Graves, E.E.; Coleman, C.N.; Aryankalayil, M.J. The lncRNAs LINC00261 and LINC00665 are upregulated in long-term prostate cancer adaptation after radiotherapy. Mol. Ther. Nucleic Acids 2021, 24, 175–187. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Components of regulatory circuits and driving events involved in NEPC development.
Figure 1. Components of regulatory circuits and driving events involved in NEPC development.
Ncrna 07 00075 g001
Figure 2. Non-coding RNAs in the regulation of NEPC. microRNAs and lncRNAs associated with the NE phenotype in clinical PCa samples (left) and NE-like changes in cellular models (right) were experimentally identified to target NED markers and regulators (bottom; dark pink, light blue, and green colours visualise driving events of NEPC based on their mode of action, i.e., upregulation, downregulation, or altered signalling status, respectively. Orange depicts NED markers. The font size reflects the frequency of a particular NED regulator/marker as a target of tested group of miRNA molecules.
Figure 2. Non-coding RNAs in the regulation of NEPC. microRNAs and lncRNAs associated with the NE phenotype in clinical PCa samples (left) and NE-like changes in cellular models (right) were experimentally identified to target NED markers and regulators (bottom; dark pink, light blue, and green colours visualise driving events of NEPC based on their mode of action, i.e., upregulation, downregulation, or altered signalling status, respectively. Orange depicts NED markers. The font size reflects the frequency of a particular NED regulator/marker as a target of tested group of miRNA molecules.
Ncrna 07 00075 g002
Table 1. Markers associated with NED of PCa cells, their biological function, and expression in clinical PCa and exper-imental models.
Table 1. Markers associated with NED of PCa cells, their biological function, and expression in clinical PCa and exper-imental models.
NameBiological FunctionRefRole in PCa/NEPCRef
Chromogranin A (CgA)
  • member of granin family
  • biogenesis of secretory granules
  • glucose and calcium homeostasis
[20,21]
  • NED marker
  • elevated plasma levels associate with poor prognosis in hormone-refractory PCa
  • IHC staining correlates with both grade and stage
  • independent predictor of overall survival and progression-free survival in CRPC
[22,23,24,25]
Chromogranin B (CgB)
  • member of granin family
  • secretory protein
[21]
  • marker of NED in prostate adenocarcinoma
  • increased level in transdifferentiated LNCaP subclones in vitro
[26,27]
Neuron-specific enolase (γ-enolase, NSE)
  • isoenzyme of glycolytic enzyme enolase
  • catalyzes conversion of 2-phosphoglycerate to phosphoenolpyruvate and its reverse reaction during gluconeogenesis
  • expressed in cytoplasm of neurons and NE cells, erythrocytes, and platelets
[28]
  • increased serum level correlates with prognosis in advanced PCa, mainly mCRPC
  • increased pretreatment NSE serum level in metastatic PCa patients correlates with poor survival
  • elevated in mCRPC compared to clinically localized and hormone-naïve PCa
[29,30,31]
Synaptophysin (Syn)
  • membrane protein of small synaptic vesicles
  • found also in dense-core chromaffin and neurosecretory granules
  • incorporated in lipid bilayer forms a cation channel essential for neurosecretion
[32]
  • detected in metastasis of CRPC patients
  • detected on circulating tumour cells in CRPC patients; expression correlated with resistance to enzalutamide and abiraterone acetate
[33,34]
CD56 (N-CAM, neural cell adhesion molecule-1)
  • member of immunoglobulin superfamily
  • involved in homophilic and heterophilic interaction
  • expression on surface of neural cells and some cells of immune system
  • aberrant expression in haematological malignancies and solid tumours
[35]
  • specific NED marker in endocrine lung cancer
  • specific surface marker of NEPC
[36,37]
L-dopa decarboxylase (DDC)
  • decarboxylation of L-Dopa to dopamine, 5-hydroxytryptophan (5-HTP) to serotonin and also other aromatic acids to corresponding amines
  • supply organism with essential neurotransmitters
  • implication in Parkinson’s disease
[38]
  • AR coactivator
  • NED marker
  • modulator of AR-regulated genes
[39,40,41]
Class III β-tubulin (TUBBIII)
  • tubulin formation (heterodimers with
  • α-tubulin)
  • constitutive expression in central and peripheral nervous system and in testes
  • important for neural development
  • expression induced by hypoxia and poor nutrient supply
[42]
  • increased after ADT in vitro
  • expressed in CRPC patients
  • taxane-based chemotherapy resistance
[43,44]
Gastrin-releasing peptide (GRP)
  • neuropeptide analogous to amphibian bombesin
  • stimulation of all gastrointestinal hormones’ secretion, intestinal and pancreatic secretion, and motility
  • exocrine and endocrine secretion, smooth muscle contraction, pain transmission
  • mitogen, morphogen, pro-angiogenic factor in cancers
[45,46]
  • increased expression of GRP and receptor GRPR in response to androgen ablation in vitro
  • GRP/GRPR signalling supports AI growth of LNCaP by increasing AR-V7 expression
  • GRPR amplification/overexpression in CRPC
  • GRP secretion from NE-like cells induced by GABA through GABBR1 receptor
  • GRPR overexpression in primary PCa compared to non-neoplastic tissue (attractive target for PCa treatment)
[47,48,49]
Calcitonin gene-related peptide (CGRP)
  • result of an alternative RNA processing of the calcitonin gene
  • vasodilator
  • involved in cardiovascular regulation, pathophysiology of migraine, arthritis, wound healing
[50]
  • expressed in prostate gland in NE cells and autonomic and sensory nerves
  • serum levels correlated with clinical stage in patients receiving hormonal therapy
  • CGRP increases invasion of PC-3 cell line in vitro
[51,52,53]
Proadrenomedullin N-terminal 20-peptide (PAMP)
  • member of calcitonin family of peptides
  • potent angiogenic factor
[54]
  • detected in CgA-positive NE cells in both normal and neoplastic prostate
[55,56]
Adrenomedullin (AM)
  • member of CGRP family
  • produced and secreted by adrenal medulla cells, tumour cells
  • vasodilation, cell growth, regulation of hormone secretion, apoptosis modulation, inflammatory regulation
[57]
  • expressed by basal cells
  • secreted by AI cell lines in vitro
  • production of AM by LNCaP in response to androgen withdrawal
  • AM mediates NED in vitro and in xenografts in vivo
[55,56,58,59]
Secretagogin
  • calcium-binding protein
  • expressed in brain, GI tract, pancreas, thyroid, adrenal medulla
  • exocytosis, insulin synthesis and function, stress-hormone release
[60]
  • colocalization with CgA and NSE in both benign and cancer NE cells
  • not stored in secretory vesicles
[61]
Parathyroid hormone-related peptide (PTHrP)
  • produced in low concentration in virtually all tissues
  • function in transepithelial calcium transport in kidney and mammary gland, smooth muscle relaxation in uterus, bladder, GI tract, arterial wall
  • cellular differentiation and apoptosis
[62]
  • increased expression in NE-transdifferentiated subclones of LNCaP in vitro
  • protection of neighbouring PCa cells from dox-induced apoptosis
  • stimulation of MDSC in bone marrow, which recruited to tumour tissue, stimulated PCa growth, and angiogenesis
  • promotion of aggressive and metastatic progression of PCa through EMT induction
[27,63,64,65,66,67]
Neurotensin (NTS)
  • neurotransmitter found in CNS and GI tract
  • paracrine or endocrine peptide in digestive and cardiovascular system
  • growth stimulatory effect on cancer cells
[68]
  • induction of NTS expression in response to androgen withdrawal in LNCaP
  • NE-transdifferentiated subclones express NTS, while parental not
  • induction by castration in vivo
  • NED induction in LNCaP through receptors NTSR1 and NTSR3
  • NTSR1 expressed in 91.8% of PCa compared to 8% of BPH
  • NTSR1 expressed also in lymph node metastasis
[27,69,70,71,72,73]
Vascular endothelial growth factor (VEGF)
  • important factor in vasculogenesis and angiogenesis
  • upregulation in cancers, affects tumour angiogenesis
  • secretion by cancer cells and stroma supports endothelial cells and leads to formation of new vessels
[74]
  • detected in CgA-positive NE cells in PCa
  • NEPC phenotype and angiogenesis correlation
  • higher plasma levels in clinically localized PCa compared to healthy, and in metastatic patients compared to clinically localized
  • preoperative plasma levels associated with biochemical progression after radical prostatectomy and LN metastasis
[75,76,77]
Histamine
  • neurotransmitter
  • 4 types of receptors H1R/H4R
[78]
  • H3R overexpression in PCa vs. normal tissue, correlation with Gleason score
  • H3R stimulates growth of LNCaP
  • H3R expression associated with AR expression present in mast cells and in NE cells in adenomatous prostate
[78,79]
Serotonin (5-hydroxy-tryptamine, 5-HT)
  • neurotransmitter
  • treatment of LNCaP with 5-HT induced NED
  • growth factor in PCa cell lines
[80,81]
Carcinoembryonic antigen-related cell adhesion molecule 5 (CEACAM5)
  • glycoprotein belonging to the family of carcinoembryonic antigen
  • involved in adhesion and migration
  • overexpressed in 90% of gastrointestinal, colorectal and pancreatic cancer
[82,83,84]
  • potential specific surface antigen of NEPC
  • expression detected in over 60% of NEPC including patients with end-stage disease
  • CEACAM5 antibody-drug conjugate labetuzuman govitecan showed therapeutic potential in PCa and particularly NEPC
[37,85]
Nerve growth factor (NGF)
  • member of neurotrophins
  • regulation of growth, maintenance, and survival of certain types of neurons, control of synthesis of neuropeptides and neurotransmitters
[86]
  • stimulation of EMT through TrkA receptor in CRPC cell lines
  • crosstalk between AR and NGF receptor TrkA in LNCaP
  • increase of NGF in response to androgen deprivation promotes NED
[87,88,89]
Neuropeptide Y (NPY)
  • member of NPY family of biologically active peptides
  • one of the most abundant neuropeptides in brain
  • growth promoting factor in various malignancies
[90]
  • high expression in PCa vs. other cancers
  • bimodal distribution in CRPC with lower levels associated with NED mCRPC
  • regulator of nerve-PCa cells interaction, NPY-neural axis regulates apoptosis, metabolism, therapy resistance
[91,92]
Table 2. Cancer-related effects of miRNAs associated with NEPC.
Table 2. Cancer-related effects of miRNAs associated with NEPC.
miRNAAssociationwith NEDValidated TargetExpression in PCa Clinical SamplesCancer-Related EffectPrognosis Correlation with Clinical DataBiomarkerOther Findings
NED MarkerPositive NED RegulatorNegative NED RegulatorExperimental FindingsSource: Indication
hsa-miR-194
  • ⇗ in clinical NEPC [190]
---
  • ⇗ in advanced disease [191]
  • ⇗ expression in primary tumour: poor prognosis [192]
  • serum: BCR prediction [192]
  • prostate biopsy: relapse prediction [193]
  • induces NED through FOXA1 [190]
hsa-miR-375
  • ⇗ in NEPC tissues [194]
  • ⇗ NE in cells [194]
-
  • ⇗ in PCa [197,198,199]
  • ⇗ in advanced PCa [200]
  • ⇗ in metastatic CRPC [201]
  • ⇗ docetaxel resistance [197]
  • associated with epithelial phenotype [202]
  • ⇗ proliferation, migration, tumour growth [203]
  • dual effect on malignant phenotype [200]
  • poor overall survival [204]
  • relapse after radiotherapy, shorter overall survival [205]
  • early progression [201]
  • association with baseline CTC count and PSA response [206]
  • serum: PCa vs healthy [122,207]
  • serum: BPH vs PCa [208]
  • urine: BPH vs PCa [209]
  • urine: disseminated vs localized [210]
  • plasma: treatment outcome prediction [204,205]
  • plasma: disease staging [211]
  • plasma: metastasis prediction [212,213]
  • serum: advanced disease [214]
  • correlates with CTCs in metastatic patients [202]
  • positive correlation with AR expression [215]
  • enriched in epithelial cells [216]
  • ⇗ in NEPC patient datasets [121,194,217]
hsa-miR-301a
  • enriched in NEPC tissues [194]
  • induces NE in cells [194]
--
  • ⇗ in prostate tumour relative to adjacent tissue [220]
  • ⇗ proliferation [221]
  • ⇗ radioresistance [222]
  • ⇗ EMT [223]
  • ⇗ migration, invasion [224]
  • increased risk of BCR [223]
  • high predicts metastasis [225]
  • serum, tumour: BPH vs PCa [226]
  • serum, needle biopsy: low grade tumours [227]
  • ⇗ by hypoxia [222]
  • ⇗ by hyperglycemia [221]
hsa-miR-106a
  • ⇘ in NEPC tissues [194]
  • ⇘ in experimental NED [228]
--
  • ⇗ in high grade tumours [232]
  • ⇘ expression with ⇗ malignity [233]
  • ⇗ in solid tumours [234]
  • ⇗ proliferation and metastasis [235]
  • confers radioresistance [232]
  • ⇗ expression - BCR [236]
  • blood: ⇗ predicts BCR [236]
  • serum: localized PCa vs BPH [237]
  • serum: low risk vs aggressive PCa [238]
  • regulated by lncRNAs HAND2-AS1 [239]
  • FER1LR
  • [240]
hsa-miR-92a---
  • ⇘ [242,243]
  • ⇗ in solid tumours of different origin [234]
  • ⇘ viability, migration, invasion [242]
  • ⇗ viability, migration, invasion [244]
  • ⇗ proliferation [245]
-
  • urine: PCa vs BPH vs healthy [246]
  • regulated by lncRNA FER1LR [247]
hsa-miR-19b
  • ⇘ in NEPC tissues [194]
--
  • ⇗ proliferation [245]
-
  • plasma: localized vs metastatic PCa [251]
  • urine: BCR [252]
  • urine, urine EVs, plasma: PCa vs BPH vs healthy [246]
  • biopsy: tumour vs adjacent tissue [253]
-
hsa-miR-20b
  • ⇘ in NED in vitro [228]
--
  • ⇗ in PCa vs adjacent tissue [244,255]
  • ⇘ migration, invasion, EMT [256]
  • ⇗ proliferation, migration [255]
  • poor survival [244]
  • tissue: ⇗ predicts BCR [236]
  • regulated by lncRNA PART1 [257]
hsa-miR-363
  • ⇘ in NEPC tissues [194]
- --
  • ⇗ in recurrent PCa [258]
  • ⇘ in young PCa patients [199]
  • ⇘ in expression concomitantly with an ⇗ in malignancy [233]
  • ⇗ proliferation and EMT [259]
--
  • miR-363 biogenesis regulated by IFIT5, downstream of IFNgamma - antiviral response) [189]
hsa-miR-106b
  • ⇗ in NED from hypoxia [260]
  • ⇘ in experimental NED [228]
--
  • ⇗ in PCa vs BPH [263]
  • ⇗ in PCa [264] [265]
  • ⇗ in PCa and metastases [266]
  • ⇗ viability, migration, invasion [264]
  • overrides radiation-induced cell cycle arrest [267]
  • associated with disease recurrence [266]
--
hsa-miR-93
  • ⇘ in experimental NED [228]
--
  • ⇗ in PCa [199]
  • ⇗ in patients with LN metastases [269]
  • promotes PCa progression [270,271]
  • ⇗ expression predicts poor survival [272]
  • blood: BPH vs PCa [208]
  • seminal plasma: disease aggressiveness [273]
  • plasma: disease prediction [274]
  • serum: PCa diagnosis [275]
-
hsa-miR-25
  • ⇗ in NED from hypoxia [260]
  • ⇗ in small cell neuroendocrine carcinoma [276]
-
  • ⇗ in PCa [199]
  • ⇗ in advanced PCa [260]
  • ⇗ in patients with LN metastases [269]
  • ⇘ invasiveness [279]
-
  • serum: disease stage and risk [233]
  • serum: decreased in cancer [280]
-
hsa-let-7
  • ⇘ in NEPC [128]
--
  • ⇘ in advanced PCa [285]
  • ⇘ favors progression and self-renewal [129]
  • ⇘ correlates with early clinical failure [286]
  • urine: cancer cell - macrophage signalling [287]
  • urine: PCa vs healthy [288]
  • negative regulation by lncRNA TTTY15 [289]
  • suppresses AR via Myc [290]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Slabáková, E.; Kahounová, Z.; Procházková, J.; Souček, K. Regulation of Neuroendocrine-like Differentiation in Prostate Cancer by Non-Coding RNAs. Non-Coding RNA 2021, 7, 75. https://doi.org/10.3390/ncrna7040075

AMA Style

Slabáková E, Kahounová Z, Procházková J, Souček K. Regulation of Neuroendocrine-like Differentiation in Prostate Cancer by Non-Coding RNAs. Non-Coding RNA. 2021; 7(4):75. https://doi.org/10.3390/ncrna7040075

Chicago/Turabian Style

Slabáková, Eva, Zuzana Kahounová, Jiřina Procházková, and Karel Souček. 2021. "Regulation of Neuroendocrine-like Differentiation in Prostate Cancer by Non-Coding RNAs" Non-Coding RNA 7, no. 4: 75. https://doi.org/10.3390/ncrna7040075

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