Next Article in Journal
HER2 Expression in Bladder Cancer: A Focused View on Its Diagnostic, Prognostic, and Predictive Role
Next Article in Special Issue
Cancer Stem Cells and Prostate Cancer: A Narrative Review
Previous Article in Journal
Impact of the Voltage-Gated Calcium Channel Antagonist Nimodipine on the Development of Oligodendrocyte Precursor Cells
Previous Article in Special Issue
Transcripts of the Prostate Cancer-Associated Gene ANO7 Are Retained in the Nuclei of Prostatic Epithelial Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dependence on MUC1-C in Progression of Neuroendocrine Prostate Cancer

Dana-Farber Cancer Institute, Harvard Medical School, Boston, MA 02215, USA
Int. J. Mol. Sci. 2023, 24(4), 3719; https://doi.org/10.3390/ijms24043719
Submission received: 26 January 2023 / Accepted: 11 February 2023 / Published: 13 February 2023
(This article belongs to the Special Issue Molecular Biology of Prostate Cancer)

Abstract

:
Castration resistant prostate cancer (CRPC) is responsive to androgen receptor (AR) axis targeted agents; however, patients invariably relapse with resistant disease that often progresses to neuroendocrine prostate cancer (NEPC). Treatment-related NEPC (t-NEPC) is highly aggressive with limited therapeutic options and poor survival outcomes. The molecular basis for NEPC progression remains incompletely understood. The MUC1 gene evolved in mammals to protect barrier tissues from loss of homeostasis. MUC1 encodes the transmembrane MUC1-C subunit, which is activated by inflammation and contributes to wound repair. However, chronic activation of MUC1-C contributes to lineage plasticity and carcinogenesis. Studies in human NEPC cell models have demonstrated that MUC1-C suppresses the AR axis and induces the Yamanaka OSKM pluripotency factors. MUC1-C interacts directly with MYC and activates the expression of the BRN2 neural transcription factor (TF) and other effectors, such as ASCL1, of the NE phenotype. MUC1-C also induces the NOTCH1 stemness TF in promoting the NEPC cancer stem cell (CSC) state. These MUC1-C-driven pathways are coupled with activation of the SWI/SNF embryonic stem BAF (esBAF) and polybromo-BAF (PBAF) chromatin remodeling complexes and global changes in chromatin architecture. The effects of MUC1-C on chromatin accessibility integrate the CSC state with the control of redox balance and induction of self-renewal capacity. Importantly, targeting MUC1-C inhibits NEPC self-renewal, tumorigenicity and therapeutic resistance. This dependence on MUC1-C extends to other NE carcinomas, such as SCLC and MCC, and identify MUC1-C as a target for the treatment of these aggressive malignancies with the anti-MUC1 agents now under clinical and preclinical development.

1. Background

Castration resistant prostate cancer (CRPC) is effectively treated with agents, such as enzalutamide and abiraterone, that target the androgen receptor (AR) axis [1]. However, patients with CRPC invariably develop resistance to AR pathway-targeted therapy and often progress to a more aggressive form with neuroendocrine (NE) features [2,3,4,5]. As a result, the incidence of treatment associated neuroendocrine prostate cancer (t-NEPC) has been increasing with the now widespread use of AR-targeted agents [4,5]. Poorly differentiated NE carcinoma of the prostate can also develop de novo with pathologies similar to that of small-cell lung cancer (SCLC) and other types of small cell carcinomas [6,7]. Despite treatment with docetaxel, cabazitaxel and platinum-based chemotherapy, the median overall survival (OS) of patients with t-NEPC and de novo NEPC has remained poor. Immune checkpoint inhibitors have also had limited success in the treatment of CRPC and NEPC [8,9,10]. In this regard, patients who progress to t-NEPC have a median OS of <1 year, whereas those with de novo NEPC have a median OS of 16.8 months from diagnosis [4,5,11]. Notably, there are presently few effective targeted agents for NEPC treatment.
Resistance to AR pathway-targeted agents is associated with significant increases in truncated AR splice variant-7 (AR-V7) expression [12,13,14]. Anti-AR-V7 agents have been under development, although they have not been approved as yet for advanced CRPC or NEPC therapy [15]. DNA damage repair pathways are dysregulated in metastatic PC and therefore represent potential vulnerabilities for treatment [16,17]. Along these lines, PARP inhibitors, such as olaparib, have extended relapse-free survival of patients with CRPCs that harbor mutations in homologous recombination repair (HRR) genes [18]. Based on these findings, olaparib was approved for the treatment of enzalutamide or abiraterone unresponsive mCRPCs with certain HRR gene mutations [19]. Trials targeting PI3K/AKT [20], Aurora kinase A [21] and MEK [22] are underway based on their dysregulation in advanced CRPC and in progression to NEPC. In addition, RET has been identified as a target based on the overexpression of this tyrosine kinase in NEPC [23].
As another potential target, the MUC1 gene is aberrantly expressed in advanced CRPC and NEPC [24]. MUC1 is amplified in 30% of a CRPC cohort with NEPC enrichment compared to 6% in the SU2C CRPC cohort and 2% in the TCGA primary PC cohort [24]. In addition, MUC1 expression is significantly increased in CRPCs compared to localized, hormone-naïve PCs [24]. Upregulation of MUC1 in advanced CRPC is associated with (i) Gleason grades ≥ 7, aggressive disease and increased risk of recurrence [25,26,27], (ii) early biochemical failure and PC-related death [28,29], and (iii) bone metastases [30]. These findings supported the potential of MUC1 as a target for advanced CRPC and NEPC treatment.

1.1. Evolution of the MUC1 Gene for the Adaptation of Barrier Tissues

MUC1 was discovered based on the overexpression of this gene in human breast cancers [31,32]. The cloning of the MUC1 gene identified a unique structure with variable numbers of conserved 60 base-pair tandem repeats (TRs) [33,34]. Further characterization demonstrated frequent alterations of the MUC1 gene in breast cancers [35]. Of interest were the subsequent findings that MUC1 first appeared in mammals, which supported a fundamental role in placentation and lactation [36,37].
MUC1 encodes a polypeptide that undergoes autocleavage into N-terminal (MUC1-N) and C-terminal (MUC1-C) subunits [38]. In turn, MUC1-N and MUC1-C form a noncovalent complex which localizes to the cell membrane [38]. MUC1-N contains 20 aa TRs that are abundantly modified by O-glycosylation and extends from the cell surface into a protective mucous barrier [38]. MUC1-C is the transmembrane component of the heterodimer, containing a 58 aa extracellular domain, 28 aa transmembrane region and a 72 aa cytoplasmic tail [38]. The MUC1-N/MUC1-C complex plays a role in physically protecting barrier tissues from infections [39]. This barrier function also plays a role in embryo implantation and placentation [40,41]. Importantly, further evidence has indicated that MUC1-C evolved in mammals to confer the adaptation of barrier tissues, including resident stem cells and immune cells, to loss of homeostasis (Figure 1) [42].
The activation of MUC1-C in the response of barrier tissues to stress induces inflammatory, proliferative and remodeling pathways that are associated with wound healing and repair [42,44]. MUC1-C represses CRB3 and E-cadherin expression in promoting loss of polarity [44]. MUC1-C also induces (i) the epithelial-mesenchymal transition (EMT) by activating the expression of the ZEB1, TWIST1 and SNAIL EMT TFs, and (ii) epigenetic reprogramming by the activation of Polycomb Repressive Complex 1 (PRC1) and PRC2 (Figure 1) [44,45]. These responses to loss of homeostasis are conceptually reversible with wound repair; however, prolonged MUC1-C activation in response to chronic inflammation imprints these changes with progression to cancer (Figure 1) [42,43,44]. As a consequence, MUC1-C has the capacity to promote carcinogenesis in barrier tissues, such as the prostatic epithelium, that are subject to chronic inflammatory responses [42]. Moreover, chronic inflammation associated with diet, injury and an altered microbiome has been implicated in driving prostate carcinogenesis [46]. Prostatitis is associated with EMT [47,48] and prostate cancer [49], indicating that the prolonged activation of MUC1-C in settings of chronic prostatitis could also contribute to PC progression.

1.2. Importance of MUC1-C in Suppressing the AR Axis and Driving CRPC→NEPC Progression

MUC1-C drives lineage plasticity and progression to the CSC state in human cancers [42,44]. Studies in AR-dependent LNCaP cells selected for proliferation in androgen-depleted medium demonstrated upregulation of MUC1-C expression [24]. The androgen-independent LNCaP-AI cells also exhibited downregulation of AR axis signaling [24], in concert with the finding that MUC1-C suppresses AR expression in PC cells [50]. In further support for the repression of the AR axis, silencing MUC1-C in this model activated PSA/KLK3, NKX3.1 and TMPRSS2 expression [24]. This inverse relationship between MUC1 and AR signaling was extended by the observation that MUC1-high CRPC tumors associate with decreased AR, KLK3, TMPRSS2, HERC3 and NKX3-1 expression levels.
BRN2 is a neural transcription factor that drives SOX2, induces NE markers and enriches for an NEPC gene signature [51]. DU-145 PC cells, derived from a brain metastasis [52], express BRN2 and other genes associated with the NE phenotype [51,53]. Studies in LNCaP-AI, DU-145 and H660 NEPC [21] cells uncovered a common role for MUC1-C in activating BRN2 expression in association with upregulation of MYC, SOX2 and the NE phenotype (Figure 2) [24]. The finding that MUC1 expression in CRPC tumors associates with upregulation of BRN2, SOX2 and the NEPC score provided additional evidence that, in parallel with the suppression of the AR axis, MUC1-C promotes CRPC→NEPC progression [24]. In further support for the notion that MUC1-C drives NEPC dedifferentiation, MUC1-C was necessary for the induction of MYCN, EZH2 and the ASCL1, AURKA and SYP NE markers, which have been associated with progression to NEPC (Figure 2) [21,54]. NE dedifferentiation vs. transdifferentiation is used here in that MUC1-C induces the NE phenotype in association with stemness, as evidenced by the induction of self-renewal capacity and tumorigenicity [24].
t-NEPC is associated with the activation of gene programs that drive lineage plasticity, the CSC state and NE dedifferentiation [2,3,4,55,56]. The observation that MUC1-C induces MYC and SOX2 was extended to include KLF4 and OCT4, which collectively represent the Yamanaka OSKM pluripotency factors that are sufficient to confer lineage plasticity and dedifferentiation of fibroblasts [57]. Pluripotency factors are repressed in somatic cells to maintain lineage specification, whereas they are induced transiently in wound healing and persistently in cancer [58]. Lineage plasticity and stemness in cancer contributes to progression and treatment resistance [3,59,60]. In line with promoting lineage infidelity, the upregulation of MUC1-C in NEPC progression supported the notion that MUC1-C is also necessary for the NEPC CSC state. Indeed, targeting MUC1-C genetically and pharmacologically suppressed NEPC cell self-renewal and tumorigenicity [24].
Figure 2. MUC1-C drives NEPC dedifferentiation. MUC1-C binds directly to the MYC HLH/LZ domain and contributes to the induction of MYC target genes. MUC1-C/MYC complexes occupy the BRN2 promoter and induce BRN2 expression. BRN2 induces SOX2 expression [51]. In addition, MUC1-C drives KLF4 and OCT4, which are collectively referred to as OSKM factors, and are sufficient for inducing pluripotency and dedifferentiation of somatic cells [57]. In addition, MUC1-C activates the inflammatory NF-κB p65 pathway and, by binding directly to NF-κB p65, promotes the activation of NF-κB p65 target genes [61], including (i) ZEB1 and thereby EMT and stemness, and (ii) EZH2 and epigenetic reprogramming [45,62]. In this way, MUC1-C integrates activation of the MYC and NF-κB p65 pathways to drive NEPC dedifferentiation and self-renewal. Figure modified from [24].
Figure 2. MUC1-C drives NEPC dedifferentiation. MUC1-C binds directly to the MYC HLH/LZ domain and contributes to the induction of MYC target genes. MUC1-C/MYC complexes occupy the BRN2 promoter and induce BRN2 expression. BRN2 induces SOX2 expression [51]. In addition, MUC1-C drives KLF4 and OCT4, which are collectively referred to as OSKM factors, and are sufficient for inducing pluripotency and dedifferentiation of somatic cells [57]. In addition, MUC1-C activates the inflammatory NF-κB p65 pathway and, by binding directly to NF-κB p65, promotes the activation of NF-κB p65 target genes [61], including (i) ZEB1 and thereby EMT and stemness, and (ii) EZH2 and epigenetic reprogramming [45,62]. In this way, MUC1-C integrates activation of the MYC and NF-κB p65 pathways to drive NEPC dedifferentiation and self-renewal. Figure modified from [24].
Ijms 24 03719 g002

1.3. MUC1-C Drives NEPC Cell Stemness by Activating the SWI/SNF BAF Chromatin Remodeling Complex

The seminal findings that NEPC CSCs are dependent on MUC1-C for self-renewal capacity uncovered a new line of investigation; that is, more precisely determining how MUC1-C drives the NEPC CSC state. Certain insights in this regard emerged from work on the involvement of MUC1-C in activating the PRC1/2 complexes in cancer cells [43,62,63]. The mammalian SWI/SNF BRG/BRAHMA factor (BAF) chromatin remodeling complex intersects with modifications of nucleosomes by PRC1/2 in regulating gene expression and cell fate [64,65,66,67]. Canonical core BAF subunits include (i) a BRG1/SMARCA4 or BRM/SMARCA2 ATPase, (ii) SMARCB1, which is essential for targeting enhancers, (iii) ARID1A/B, which maintain BAF on enhancers, and (iv) SMARCC1/2 leucine zipper proteins [68,69,70,71]. The ETS family transcription factor ERG requires BAF for mediating prostate oncogenesis [72], whereas little had been known about the regulation of BAF in PC cells.
The finding that MUC1-C activates BAF in NEPC cells thus opened a new line of investigation for the involvement of this SWI/SNF complex in NEPC progression [73]. The binding of MUC1-C to the E2F1 TF in NEPC cells was found to induce the expression of BRG1 and ARID1A, as well as SMARCD1/BAF60a, SMARCC1/BAF155 and SMARCC2/BAF170 [73], which are components of the embryonic stem cell (ESC)-specific BAF (esBAF) complex [74,75]. esBAF is required for regulating ESC gene expression and thereby ESC self-renewal and differentiation [76]. Notably in this regard, MUC1-C associates with ESC and CSC gene signatures in NEPC cells [73]. From these results, NOTCH1, which has been linked to stemness and driving CRPC [77], was identified as a MUC1-C-activated gene that is also dependent on BRG1 and ARID1A for expression (Figure 3) [73]. In addition, the MUC1-C→esBAF pathway was found to induce the NANOG pluripotency factor which promotes stemness (Figure 3) [78]. Consistent with these findings, MUC1-C→esBAF signaling was necessary for NEPC cell self-renewal capacity [73]. These findings further highlighted the involvement of MUC1-C in activating (i) MYC with the induction of pluripotency and the NE phenotype [24], and (ii) E2F1→esBAF in inducing NOTCH1, NANOG and the CSC state (Figure 3) [73].

1.4. MUC1-C Activates the SWI/SNF PBAF Chromatin Remodeling Complex in Integrating Redox Balance with Stemness of NEPC Cells

The polybromo-associated BAF (PBAF) chromatin remodeling complex includes the BRG1 ATPase, PBRM1/BAF180, ARID2/BAF200 and BRD7 [64]. BRG1 is shared with BAF and PBAF [64], whereas PBRM1 has a distinguishing capacity in PBAF with functions in (i) DNA damage-associated transcriptional repression and DNA repair [79,80], and (ii) regulation of genes involved in the oxidative stress response and apoptosis [81]. In addition, ARID2 regulates IFN-induced genes [82] and BRD7 interacts with BRCA1-mediated transcription [83]. Interestingly, as found for esBAF [73], the MUC1-C→E2F1 pathway was essential for the induction of PBRM1, ARID2 and BRD7 expression in NEPC cells (Figure 4) [84]. As also reported for esBAF (BRG1, ARID1A), MUC1-C/E2F1 complexes were detectable on the PBRM1, ARID2 and BRD7 promoter regions containing E2F binding motifs [84]. These findings, and those demonstrating that MUC1-C→E2F1 signaling activates genes encoding PRC2 (EZH2, SUZ12 and EED) subunits, indicated that MUC1-C plays a role in integrating functions of the PRC2 and SWI/SNF BAF and PBAF complexes in NEPC cells [45,62,73,84].
The MUC1-C→esBAF pathway contributes to the NEPC CSC state [73]. CSCs are effective in repairing DNA damage and controlling ROS levels [85,86]. In support of MUC1-C in maintaining redox balance in CRPC CSCs, MUC1-C-induced PBRM1 expression associated with the activation of NRF2 target genes (Figure 4) [73]. The NRF2 TF is a master regulator of the anti-oxidant response and driver of cancer progression and resistance to therapy [87,88]. MUC1-C formed a complex with NRF2 and PBRM1 on the SLC7A11 gene [84], which encodes the xCT cystine-glutamate antiporter, a subunit of the Xc- system that confers cysteine uptake for GSH synthesis [89]. In this way, MUC1-C, NRF2 and PBRM1 were necessary for increasing the chromatin accessibility of the SLC7A11 gene and for xCT expression [84]. MUC1-C/NRF2/PBRM1 complexes also contributed to the chromatin accessibility of the G6PD gene and expression of glucose-6-phosphate dehydrogenase (G6PD), which converts NADP+ to NADPH [84]. In accordance with this MUC1-C-driven activation of SLC7A11, G6PD and other antioxidant genes, silencing MUC1-C and PBRM1 in NEPC cells decreased GSH, GSH/GSSG and NADP/NADPH levels in association with increases in sensitivity to oxidative stress [84]. Of further interest, MUC1-C functioned in cross-talk between the esBAF and PBAF complexes in integrating the expression of pluripotency and stemness factors with effectors of redox balance (Figure 4) [84]. These findings for MUC1-C-induced regulation of esBAF and PBAF formed the basis for studies of MUC1-C involvement in the remodeling of chromatin in NEPC CSCs.

1.5. MUC1-C Regulates Chromatin Accessibility across the Genome of NEPC Cells in Promoting the CSC State

Chromatin remodeling is critical for lineage plasticity, EMT and the CSC state [90]. MUC1-C drives the esBAF and PBAF chromatin remodeling complexes in NEPC CSCs [73,84]. By extension, ATAC-seq studies showed that MUC1-C is associated with global changes in chromatin accessibility across the genome of DU-145 cells [90]. Silencing MUC1-C identified Differentially Accessible Regions (DARs) with increases and decreases in chromatin accessibility [90]. DARs were located at proximal regions within 1–3 kb and distal intergenic regions within 50–500 kb from annotated TSSs [90]. Associations of MUC1-C-induced DARs with Differentially Expressed Genes (DEGs) identified genes enriched for FOS, JUN and NEF2 binding motifs that are recognized by members of the AP-1 family of TFs. NOTCH1 was identified among these genes in concert with the demonstration that MUC1-C induces NOTCH1 in driving the self-renewal of NEPC CSCs [73,90]. In further support of esBAF involvement in activating NOTCH1 [73], MUC1-C occupied a NOTCH1 proximal enhancer-like sequence (pELS) with JUN and ARID1A in association with increases in chromatin accessibility, H3K4 trimethylation and NOTCH1 expression [90]. Similar results were obtained for the EGR1 stemness gene, indicating that MUC1-C activates a JUN- and esBAF-dependent pathway of importance for increasing the chromatin accessibility of genes that promote the NEPC state [90]. These findings indicated that MUC1-C drives NEPC progression by inducing JUN-mediated chromatin remodeling, which parenthetically is essential for the wound healing response and maintaining tissue homeostasis [42,91].
The distinction between MUC1-C-induced activation of the esBAF and PBAF complexes in driving NEPC progression was further investigated with studies of their potential involvement in inflammatory signaling pathways (Figure 5). Along this line of investigation, MUC1-C contributes to the inflammatory wound healing response and prolonged activation of MUC1-C by chronic inflammation promotes oncogenesis [42,44]. An analysis of the TCGA-PRAD and SU2C-CRPC datasets demonstrated that MUC1-high PCs significantly associate with the activation of the type II IFN response gene signature [92]. Moreover, MUC1-high PCs had significantly increased levels of IFNGR1, STAT1 and IRF1 [92], which drive IFN response genes (ISGs) and chronic inflammation in cancer cells [93]. MUC1-C was found to be necessary for IFNGR1, STAT1 and IRF1 expression in NEPC cells (Figure 5) [92]. In further support for the interactions among MUC1-C, JUN and esBAF [90], MUC1-C/JUN/ARID1A complexes were detectable on the IFNGR1 gene at a distal enhancer-like signature (dELS) in association with increases in chromatin accessibility, H3K4me3 levels and IFNGR1 expression [92]. The stimulation of the IFNGR1 complex by IFN-gamma (IFNG) activates the STAT1→IRF1 pathway [93]. Silencing MUC1-C decreased the chromatin accessibility of the STAT1 and IRF1 genes and their expression (Figure 5) [92]. In addition, MUC1-C and PBRM1 were necessary for the expression of the downstream immunosuppressive effectors (i) indoleamine-2,3-dioxygenase (IDO1) [94], (ii) tryptophanyl-tRNA synthetase (WARS) [95], and (iii) prostaglandin E synthase (PTGES) [96] (Figure 5). In support of these results in NEPC cells, MUC1-high PC tumors significantly associate with upregulation of IDO1, WARS and PTGES expression [92]. Of additional importance, MUC1-high CRPC tumors associate with the suppression of the TME, as evidenced by (i) negative regulation T cell and NK cell mediated immunity, and (ii) decreases in CD4+ memory T cells, Th2 cells, M2 macrophages and the ImmuneScore [92]. An analysis of a scRNA-seq dataset further demonstrated that MUC1 associates with CSC and IFN signatures across individual CRPC cells [92].

1.6. Involvement of MUC1-C in Lineage Plasticity of NE Cancers

The lineage plasticity of the NE phenotype is of importance to the progression and response of prostate and other types of cancers [97]. An analysis of CRPC, NEPC and small cell lung cancer (SCLC) tumors identified pan-cancer convergence to a small cell NE phenotype, characterized by TP53 and RB1 loss, common epigenetic alterations and the expression of NE markers [98]. The findings that MUC1-C drives lineage plasticity in the progression to NEPC suggested that MUC1-C may play a role in other cancers with NE dedifferentiation or transdifferentiation. In support of that notion, MUC1-C dictates NE lineage specification in pancreatic ductal carcinomas (PDAC-NE) (Table 1) [99]. As found in NEPC [24,92], MUC1-C induces the Yamanaka OSKM pluripotency factors, inflammatory IFN signaling and the NOTCH1/2 TFs in PDAC-NE cells [99]. In addition, MUC1-C interacted with MYC in activating the BRN2 and ASCL1 neural TFs and thereby the expression of NE markers [99]. In SCLC, MUC1-C activates the MYC pathway in classic NE SCLC-ASCL1 (SCLC-A), variant NE SCLC-NEUROD1 (SCLC-N) and non-NE SCLC-POU2F3 (SCLC-P) subtypes (Table 1) [100]. MUC1-C→MYC signaling was necessary for the induction of (i) NOTCH2, which is a marker of pulmonary NE stem cells, and (ii) the ASCL1 and NEUROD1 neural TFs [100]. Merkel cell carcinoma (MCC) is another aggressive NE malignancy [101]. MUC1 is dysregulated in Merkel cell polyomavirus (MCPyV)-positive and -negative MCCs [101]. In both MCC subtypes, MUC1-C interacts with MYCL in inducing the expression of the OSKM + NANOG pluripotency factors and the neural BRN2, NEUROD1 and ATOH1 TFs (Table 1) [101].
These studies have identified common MUC1-C-driven pathways in NE cancers that include the activation of (i) MYC family members, (ii) Yamanaka pluripotency factors, (iii) NE lineage dictating TFs and (iv) NOTCH stemness TFs. Moreover, targeting MUC1-C genetically and pharmacologically in NEPC, PDAC-NE, SCLC and MCC cells inhibits tumorsphere formation and tumorigenicity (Table 1) [24,92,99,100,101], in support of their dependency on MUC1-C for self-renewal. These findings have identified MUC1-C as a potential target for the treatment of NEPC and other recalcitrant NE cancers that have limited therapeutic options. To that end, the generation of MAb 3D1 against the MUC1-C extracellular domain provided an opportunity for the development of agents that target MUC1-C on the cancer cell surface [102]. As one example, an allogeneic anti-MUC1-C CAR T cell using MAb 3D1 sequences is undergoing Phase I evaluation for the treatment of MUC1-C-expressing cancers (NCT05239143: P-MUC1C-ALLO1 Allogeneic CAR-T Cells in the Treatment of Subjects with Advanced or Metastatic Solid Tumors). In addition, anti-MUC1-C huMAb3D1-MMAE ADCs are under development by the NCI NExT Program for IND-enabling studies and performing early phase clinical trials in recalcitrant cancers.

2. Conclusions

In summary, the discoveries described here indicate that MUC1-C contributes to NEPC progression by activating pathways that drive lineage infidelity, epigenetic reprogramming and chromatin remodeling [24,73,84,91,92]. In this way, MUC1-C promotes the NEPC CSC state, DNA damage resistance and immune evasion. These findings highlight MUC1-C as a target for the treatment of NEPC and certain other aggressive NE cancers with the anti-MUC1-C agents that are now under development.

Author Contributions

Wrote and revised the review: D.K. The author has read and agreed to the published version of the manuscript.

Funding

The research reported in this publication was supported by the National Cancer Institute of the National Institutes of Health under grant numbers CA097098, CA233084, CA267138 and CA262991, awarded to D.K.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author acknowledges the seminal contributions of Yota Yasumizu, Masayuki Hagiwara, Atrayee Bhattacharya, Zhou Luan, Atsushi Fushimi, Yoshihiro Morimoto and Nami Yamashita to the findings described in this review.

Conflicts of Interest

D.K. has equity interests in Reata Pharmaceuticals and Hillstream Biopharma and is a paid consultant to Reata and CanBas.

References

  1. Sartor, O.; de Bono, J.S. Metastatic prostate cancer. N. Engl. J. Med. 2018, 378, 645–657. [Google Scholar] [CrossRef] [PubMed]
  2. Puca, L.; Vlachostergios, P.J.; Beltran, H. Neuroendocrine Differentiation in Prostate Cancer: Emerging Biology, Models, and Therapies. Cold Spring Harb. Perspect. Med. 2019, 9, a030593. [Google Scholar] [CrossRef]
  3. 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]
  4. Aggarwal, R.; Huang, J.; Alumkal, J.J.; Zhang, L.; Feng, F.Y.; Thomas, G.V.; Weinstein, A.S.; Friedl, V.; Zhang, C.; Witte, O.N.; et al. Clinical and Genomic Characterization of Treatment-Emergent Small-Cell Neuroendocrine Prostate Cancer: A Multi-institutional Prospective Study. J. Clin. Oncol. 2018, 36, 2492–2503. [Google Scholar] [CrossRef]
  5. Abida, W.; Cyrta, J.; Heller, G.; Prandi, D.; Armenia, J.; Coleman, I.; Cieslik, M.; Benelli, M.; Robinson, D.; Van Allen, E.M.; et al. Genomic correlates of clinical outcome in advanced prostate cancer. Proc. Natl. Acad. Sci. USA 2019, 116, 11428–11436. [Google Scholar] [CrossRef]
  6. 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] [PubMed]
  7. Epstein, J.I.; Amin, M.B.; Beltran, H.; Lotan, T.L.; Mosquera, J.-M.; Reuter, V.E.; Robinson, B.D.; Troncoso, P.; Rubin, M.A. Proposed Morphologic Classification of Prostate Cancer With Neuroendocrine Differentiation. Am. J. Surg. Pathol. 2014, 38, 756–767. [Google Scholar] [CrossRef]
  8. Cha, H.-R.; Lee, J.H.; Ponnazhagan, S. Revisiting Immunotherapy: A Focus on Prostate Cancer. Cancer Res. 2020, 80, 1615–1623. [Google Scholar] [CrossRef]
  9. Rizzo, A.; Mollica, V.; Cimadamore, A.; Santoni, M.; Scarpelli, M.; Giunchi, F.; Cheng, L.; Lopez-Beltran, A.; Fiorentino, M.; Montironi, R.; et al. Is there a role for immunotherapy in prostate cancer? Cells 2020, 9, 2051. [Google Scholar] [CrossRef]
  10. Kim, T.J.; Koo, K.C. Current Status and Future Perspectives of Checkpoint Inhibitor Immunotherapy for Prostate Cancer: A Comprehensive Review. Int. J. Mol. Sci. 2020, 21, 5484. [Google Scholar] [CrossRef] [PubMed]
  11. Conteduca, V.; Oromendia, C.; Eng, K.W.; Bareja, R.; Sigouros, M.; Molina, A.; Faltas, B.M.; Sboner, A.; Mosquera, J.M.; Elemento, O.; et al. Clinical features of neuroendocrine prostate cancer. Eur. J. Cancer 2019, 121, 7–18. [Google Scholar] [CrossRef]
  12. Antonarakis, E.S.; Lu, C.; Wang, H.; Luber, B.; Nakazawa, M.; Roeser, J.C.; Chen, Y.; Mohammad, T.A.; Chen, Y.; Fedor, H.L.; et al. AR-V7 and Resistance to Enzalutamide and Abiraterone in Prostate Cancer. N. Engl. J. Med. 2014, 371, 1028–1038. [Google Scholar] [CrossRef]
  13. Sharp, A.; Coleman, I.; Yuan, W.; Sprenger, C.; Dolling, D.; Rodrigues, D.N.; Russo, J.W.; Figueiredo, I.; Bertan, C.; Seed, G.; et al. Androgen receptor splice variant-7 expression emerges with castration resistance in prostate cancer. J. Clin. Investig. 2019, 129, 192–208. [Google Scholar] [CrossRef]
  14. Del Re, M.; Conteduca, V.; Crucitta, S.; Gurioli, G.; Casadei, C.; Restante, G.; Schepisi, G.; Lolli, C.; Cucchiara, F.; Danesi, R.; et al. Androgen receptor gain in circulating free DNA and splicing variant 7 in exosomes predict clinical outcome in CRPC patients treated with abiraterone and enzalutamide. Prostate Cancer Prostatic Dis. 2021, 24, 524–531. [Google Scholar] [CrossRef]
  15. Liu, C.; Lou, W.; Zhu, Y.; Nadiminty, N.; Schwartz, C.T.; Evans, C.P.; Gao, A.C. Niclosamide Inhibits Androgen Receptor Variants Expression and Overcomes Enzalutamide Resistance in Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2014, 20, 3198–3210. [Google Scholar] [CrossRef]
  16. Abida, W.; Armenia, J.; Gopalan, A.; Brennan, R.; Walsh, M.; Barron, D.; Danila, D.; Rathkopf, D.; Morris, M.; Slovin, S.; et al. Prospective genomic profiling of prostate cancer across disease states reveals germline and somatic alterations that may affect clinical decision making. JCO Precis Oncol. 2017, 1, PO.17.00029. [Google Scholar] [CrossRef] [PubMed]
  17. Pritchard, C.C.; Mateo, J.; Walsh, M.F.; De Sarkar, N.; Abida, W.; Beltran, H.; Garofalo, A.; Gulati, R.; Carreira, S.; Eeles, R.; et al. Inherited DNA-Repair Gene Mutations in Men with Metastatic Prostate Cancer. N. Engl. J. Med. 2016, 375, 443–453. [Google Scholar] [CrossRef]
  18. De Bono, J.; Mateo, J.; Fizazi, K.; Saad, F.; Shore, N.; Sandhu, S.; Chi, K.N.; Sartor, O.; Agarwal, N.; Olmos, D.; et al. Olaparib for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2020, 382, 2091–2102. [Google Scholar] [CrossRef] [PubMed]
  19. Schaeffer, E.; Srinivas, S.; Antonarakis, E.S.; Armstrong, A.J.; Bekelman, J.E.; Cheng, H.; D’Amico, A.V.; Davis, B.J.; Desai, N.; Dorff, T.; et al. NCCN Guidelines Insights: Prostate Cancer, Version 1.2021. J. Natl. Compr. Cancer Netw. 2021, 19, 134–143. [Google Scholar] [CrossRef]
  20. Sweeney, C.; Bracarda, S.; Sternberg, C.N.; Chi, K.N.; Olmos, D.; Sandhu, S.; Massard, C.; Matsubara, N.; Alekseev, B.; Parnis, F.; et al. Ipatasertib plus abiraterone and prednisolone in metastatic castration-resistant prostate cancer (IPATential150): A multicentre, randomised, double-blind, phase 3 trial. Lancet 2021, 398, 131–142. [Google Scholar] [CrossRef] [PubMed]
  21. 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]
  22. Nickols, N.G.; Nazarian, R.; Zhao, S.G.; Tan, V.; Uzunangelov, V.; Xia, Z.; Baertsch, R.; Neeman, E.; Gao, A.C.; Thomas, G.; et al. MEK-ERK signaling is a therapeutic target in metastatic castration resistant prostate cancer. Prostate Cancer Prostatic Dis. 2019, 22, 531–538. [Google Scholar] [CrossRef] [PubMed]
  23. VanDeusen, H.R.; Ramroop, J.R.; Morel, K.L.; Bae, S.Y.; Sheahan, A.V.; Sychev, Z.; Lau, N.A.; Cheng, L.C.; Tan, V.M.; Li, Z.; et al. Targeting RET Kinase in Neuroendocrine Prostate Cancer. Mol. Cancer Res. 2020, 18, 1176–1188. [Google Scholar] [CrossRef]
  24. 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]
  25. Lapointe, J.; Li, C.; Higgins, J.P.; van de Rijn, M.; Bair, E.; Montgomery, K.; Ferrari, M.; Egevad, L.; Rayford, W.; Bergerheim, U.; et al. Gene expression profiling identifies clinically relevant subtypes of prostate cancer. Proc. Natl. Acad. Sci. USA 2004, 101, 811–816. [Google Scholar] [CrossRef]
  26. Andrén, O.; Fall, K.; Andersson, S.-O.; Rubin, M.A.; Bismar, T.; Karlsson, M.G.; Johansson, J.-E.; Mucci, L.A. MUC-1 gene is associated with prostate cancer death: A 20-year follow-up of a population-based study in Sweden. Br. J. Cancer 2007, 97, 730–734. [Google Scholar] [CrossRef]
  27. Eminaga, O.; Wei, W.; Hawley, S.J.; Auman, H.; Newcomb, L.F.; Simko, J.; Hurtado-Coll, A.; Troyer, D.A.; Carroll, P.R.; Gleave, M.E.; et al. MUC1 expression by immunohistochemistry is associated with adverse pathologic features in prostate cancer: A mul-ti-institutional study. PLoS ONE 2016, 11, e0165236. [Google Scholar] [CrossRef] [PubMed]
  28. Genitsch, V.; Zlobec, I.; Thalmann, G.N.; Fleischmann, A. MUC1 is upregulated in advanced prostate cancer and is an inde-pendent prognostic factor. Prostate Cancer Prostatic Dis. 2016, 19, 242–247. [Google Scholar] [CrossRef]
  29. Jiang, P.; Gu, S.; Pan, D.; Fu, J.; Sahu, A.; Hu, X.; Li, Z.; Traugh, N.; Bu, X.; Li, B.; et al. Signatures of T cell dysfunction and exclusion predict cancer immunotherapy response. Nat. Med. 2018, 24, 1550–1558. [Google Scholar] [CrossRef] [PubMed]
  30. Lin, X.; Gu, Y.; Kapoor, A.; Wei, F.; Aziz, T.; Ojo, D.; Jiang, Y.; Bonert, M.; Shayegan, B.; Yang, H.; et al. Overexpression of MUC1 and Genomic Alterations in Its Network Associate with Prostate Cancer Progression. Neoplasia 2017, 19, 857–867. [Google Scholar] [CrossRef]
  31. Kufe, D.; Inghirami, G.; Abe, M.; Hayes, D.; Justi-Wheeler, H.; Schlom, J. Differential Reactivity of a Novel Monoclonal Antibody (DF3) with Human Malignant versus Benign Breast Tumors. Hybridoma 1984, 3, 223–232. [Google Scholar] [CrossRef]
  32. Ormerod, M.G.; Steele, K.A.; Edwards, P.; Taylor-Papadimitriou, J. Monoclonal antibodies that react with epithelial membrane antigen. J. Exp. Pathol. 1984, 1, 263–271. [Google Scholar]
  33. Siddiqui, J.; Abe, M.; Hayes, D.; Shani, E.; Yunis, E.; Kufe, D. Isolation and sequencing of a cDNA coding for the human DF3 breast carcinoma-associated antigen. Proc. Natl. Acad. Sci. USA 1988, 85, 2320–2323. [Google Scholar] [CrossRef] [PubMed]
  34. Gendler, S.; Taylor-Papadimitriou, J.; Duhig, T.; Rothbard, J.; Burchell, J.A. A highly immunogenic region of a human poly-morphic epithelial mucin expressed by carcinomas is made up of tandem repeats. J. Biol. Chem. 1988, 263, 12820–12823. [Google Scholar] [CrossRef] [PubMed]
  35. Merlo, G.R.; Siddiqui, J.; Cropp, C.S.; Liscia, D.S.; Lidereau, R.; Callahan, R.; Kufe, D.W. Frequent alteration of the DF3 tumor-associated antigen gene in primary human breast carcinomas. Cancer Res. 1989, 49, 6966–6971. [Google Scholar] [PubMed]
  36. Duraisamy, S.; Ramasamy, S.; Kharbanda, S.; Kufe, D. Distinct evolution of the human carcinoma-associated transmembrane mucins, MUC1, MUC4 AND MUC16. Gene 2006, 373, 28–34. [Google Scholar] [CrossRef]
  37. Duraisamy, S.; Kufe, T.; Ramasamy, S.; Kufe, N. Evolution of the human MUC1 oncoprotein. Int. J. Oncol. 2007, 31, 671–677. [Google Scholar] [CrossRef] [PubMed]
  38. Kufe, D. Mucins in cancer: Function, prognosis and therapy. Nat. Rev. Cancer 2009, 9, 874–885. [Google Scholar] [CrossRef]
  39. McAuley, J.; Linden, S.K.; Png, C.W.; King, R.M.; Pennington, H.L.; Gendler, S.J.; Florin, T.H.; Hill, G.; Korolik, V.; McGuckin, M.A. MUC1 cell surface mucin is a critical element of the mucosal barrier to infection. J. Clin. Investig. 2007, 117, 2313–2324. [Google Scholar] [CrossRef]
  40. Constantinou, P.E.; Morgado, M.; Carson, D.D. Transmembrane mucin expression and function in mmbryo implantation and placentation. Adv. Anat. Embryol. Cell Biol. 2015, 216, 51–68. [Google Scholar]
  41. Griffith, O.W.; Chavan, A.R.; Protopapas, S.; Maziarz, J.; Romero, R.; Wagner, G.P. Embryo implantation evolved from an an-cestral inflammatory attachment reaction. Proc. Natl. Acad. Sci. USA 2017, 114, E6566–E6575. [Google Scholar] [CrossRef]
  42. Kufe, D. Emergence of MUC1 in mammals for adaptation of barrier epithelia. Cancers 2022, 14, 4805. [Google Scholar] [CrossRef] [PubMed]
  43. Kufe, D.W. Chronic activation of MUC1-C in wound repair promotes progression to cancer stem cells. J. Cancer Metastasis Treat. 2022, 8, 12. [Google Scholar] [CrossRef]
  44. Kufe, D.W. MUC1-C in chronic inflammation and carcinogenesis; emergence as a target for cancer treatment. Carcinogenesis 2020, 41, 1173–1183. [Google Scholar] [CrossRef]
  45. Rajabi, H.; Hiraki, M.; Kufe, D. MUC1-C activates polycomb repressive complexes and downregulates tumor suppressor genes in human cancer cells. Oncogene 2018, 37, 2079–2088. [Google Scholar] [CrossRef]
  46. Sfanos, K.S.; Yegnasubramanian, S.; Nelson, W.G.; De Marzo, A.M. The inflammatory microenvironment and microbiome in prostate cancer development. Nat. Rev. Urol. 2018, 15, 11–24. [Google Scholar] [CrossRef]
  47. Shao, R.; Shi, J.; Liu, H.; Shi, X.; Du, X.; Klocker, H.; Lee, C.; Zhu, Y.; Zhang, J. Epithelial-to-mesenchymal transition and estrogen receptor α mediated epithelial dedifferentiation mark the development of benign prostatic hyperplasia. Prostate 2014, 74, 970–982. [Google Scholar] [CrossRef] [PubMed]
  48. Alonso-Magdalena, P.; Brossner, C.; Reiner, A.; Cheng, G.; Sugiyama, N.; Warner, M.; Gustafsson, J.A. A role for epitheli-al-mesenchymal transition in the etiology of benign prostatic hyperplasia. Proc. Natl. Acad. Sci. USA 2009, 106, 2859–2863. [Google Scholar] [CrossRef] [PubMed]
  49. Perletti, G.; Monti, E.; Magri, V.; Cai, T.; Cleves, A.; Trinchieri, A.; Montanari, E. The association between prostatitis and prostate cancer. Systematic review and meta-analysis. Arch. Ital. Urol. Androl. 2017, 89, 259–265. [Google Scholar] [CrossRef]
  50. Rajabi, H.; Ahmad, R.; Jin, C.; Joshi, M.; Guha, M.; Alam, M.; Kharbanda, S.; Kufe, D. MUC1-C oncoprotein confers andro-gen-independent growth of human prostate cancer cells. Prostate 2012, 72, 1659–1668. [Google Scholar] [CrossRef]
  51. 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] [PubMed]
  52. Stone, K.R.; Mickey, D.D.; Wunderli, H.; Mickey, G.H.; Paulson, D.F. Isolation of a human prostate carcinoma cell line (DU 145). Int. J. Cancer 1978, 21, 274–281. [Google Scholar] [CrossRef]
  53. 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, 650054. [Google Scholar] [CrossRef]
  54. Berger, A.; Brady, N.J.; Bareja, R.; Robinson, B.D.; Conteduca, V.; Augello, M.A.; Puca, L.; Ahmed, A.; Dardenne, E.; Lu, X.; et al. N-Myc-mediated epigenetic reprogramming drives lineage plasticity in advanced prostate cancer. J. Clin. Investig. 2019, 130, 3924–3940. [Google Scholar] [CrossRef]
  55. 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] [PubMed]
  56. Akamatsu, S.; Inoue, T.; Ogawa, O.E.; Gleave, M. Clinical and molecular features of treatment-related neuroendocrine prostate cancer. Int. J. Urol. 2018, 25, 345–351. [Google Scholar] [CrossRef]
  57. Takahashi, K.; Yamanaka, S. A decade of transcription factor-mediated reprogramming to pluripotency. Nat. Rev. Mol. Cell Biol. 2016, 17, 183–193. [Google Scholar] [CrossRef]
  58. Wollenzien, H.; Voigt, E.; Kareta, M.S. Somatic Pluripotent Genes in Tissue Repair, Developmental Disease, and Cancer. SPG BioMed 2018, 1, 10.32392/biomed.18. [Google Scholar] [CrossRef]
  59. Le Magnen, C.; Shen, M.M.; Abate-Shen, C. Lineage Plasticity in Cancer Progression and Treatment. Annu. Rev. Cancer Biol. 2018, 2, 271–289. [Google Scholar] [CrossRef]
  60. Gupta, P.B.; Pastushenko, I.; Skibinski, A.; Blanpain, C.; Kuperwasser, C. Phenotypic Plasticity: Driver of Cancer Initiation, Progression, and Therapy Resistance. Cell Stem Cell 2019, 24, 65–78. [Google Scholar] [CrossRef] [PubMed]
  61. Takahashi, H.; Jin, C.; Rajabi, H.; Pitroda, S.P.; Alam, M.; Ahmad, R.; Raina, D.; Hasegawa, M.; Suzuki, Y.; Tagde, A.; et al. MUC1-C activates the TAK1 inflammatory pathway in colon cancer. Oncogene 2015, 34, 5187–5197. [Google Scholar] [CrossRef]
  62. Rajabi, H.; Hiraki, M.; Tagde, A.; Alam, M.; Bouillez, A.; Christensen, C.L.; Samur, M.; Wong, K.-K.; Kufe, D. MUC1-C activates EZH2 expression and function in human cancer cells. Sci. Rep. 2017, 7, 7481. [Google Scholar] [CrossRef] [Green Version]
  63. Hiraki, M.; Maeda, T.; Bouillez, A.; Alam, M.; Tagde, A.; Hinohara, K.; Suzuki, Y.; Markert, T.; Miyo, M.; Komura, K.; et al. MUC1-C activates BMI1 in human cancer cells. Oncogene 2016, 36, 2791–2801. [Google Scholar] [CrossRef] [PubMed]
  64. Hodges, H.C.; Kirkland, J.G.; Crabtree, G.R. The Many Roles of BAF (mSWI/SNF) and PBAF Complexes in Cancer. Cold Spring Harb. Perspect. Med. 2016, 6, a026930. [Google Scholar] [CrossRef]
  65. Schuettengruber, B.; Bourbon, H.-M.; Di Croce, L.; Cavalli, G. Genome Regulation by Polycomb and Trithorax: 70 Years and Counting. Cell 2017, 171, 34–57. [Google Scholar] [CrossRef]
  66. Bracken, A.P.; Brien, G.L.; Verrijzer, C.P. Dangerous liaisons: Interplay between SWI/SNF, NuRD, and Polycomb in chromatin regulation and cancer. Genes Dev. 2019, 33, 936–959. [Google Scholar] [CrossRef]
  67. Clapier, C. Sophisticated Conversations between Chromatin and Chromatin Remodelers, and Dissonances in Cancer. Int. J. Mol. Sci. 2021, 22, 5578. [Google Scholar] [CrossRef] [PubMed]
  68. Pierre, R.S.; Kadoch, C. Mammalian SWI/SNF complexes in cancer: Emerging therapeutic opportunities. Curr. Opin. Genet. Dev. 2017, 42, 56–67. [Google Scholar] [CrossRef]
  69. Wang, X.; Lee, R.S.; Alver, B.; Haswell, J.R.; Wang, S.; Mieczkowski, J.; Drier, Y.; Gillespie, S.; Archer, T.C.; Wu, J.N.; et al. SMARCB1-mediated SWI/SNF complex function is essential for enhancer regulation. Nat. Genet. 2017, 49, 289–295. [Google Scholar] [CrossRef]
  70. Mathur, R.; Alver, B.H.; Roman, A.K.S.; Wilson, B.G.; Wang, X.; Agoston, A.T.; Park, P.J.; Shivdasani, R.A.; Roberts, C.W.M. ARID1A loss impairs enhancer-mediated gene regulation and drives colon cancer in mice. Nat. Genet. 2017, 49, 296–302. [Google Scholar] [CrossRef] [PubMed]
  71. Trizzino, M.; Barbieri, E.; Petracovici, A.; Wu, S.; Welsh, S.A.; Owens, T.A.; Licciulli, S.; Zhang, R.; Gardini, A. The Tumor Suppressor ARID1A Controls Global Transcription via Pausing of RNA Polymerase II. Cell Rep. 2018, 23, 3933–3945. [Google Scholar] [CrossRef]
  72. Sandoval, G.J.; Pulice, J.; Pakula, H.; Schenone, M.; Takeda, D.Y.; Pop, M.; Boulay, G.; Williamson, K.E.; McBride, M.; Pan, J.; et al. Binding of TMPRSS2-ERG to BAF Chromatin Remodeling Complexes Mediates Prostate Oncogenesis. Mol. Cell 2018, 71, 554–566.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. 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]
  74. Lessard, J.A.; Crabtree, G.R. Chromatin regulatory mechanisms in pluripotency. Annu. Rev. Cell Dev. Biol. 2010, 26, 503–532. [Google Scholar] [CrossRef]
  75. Alfert, A.; Moreno, N.; Kerl, K. The BAF complex in development and disease. Epigenet. Chromatin 2019, 12, 19. [Google Scholar] [CrossRef]
  76. Ho, L.; Ronan, J.L.; Wu, J.; Staahl, B.T.; Chen, L.; Kuo, A.; Lessard, J.; Nesvizhskii, A.I.; Ranish, J.; Crabtree, G.R. An embryonic stem cell chromatin remodeling complex, esBAF, is essential for embryonic stem cell self-renewal and pluripotency. Proc. Natl. Acad. Sci. USA 2009, 106, 5181–5186. [Google Scholar] [CrossRef] [PubMed]
  77. Stoyanova, T.; Riedinger, M.; Lin, S.; Faltermeier, C.M.; Smith, B.A.; Zhang, K.X.; Going, C.C.; Goldstein, A.S.; Lee, J.K.; Drake, J.M.; et al. Activation of Notch1 synergizes with multiple pathways in promoting castration-resistant prostate cancer. Proc. Natl. Acad. Sci. USA 2016, 113, E6457–E6466. [Google Scholar] [CrossRef]
  78. Zhang, W.; Sui, Y.; Ni, J.; Yang, T. Insights into the Nanog gene: A propeller for stemness in primitive stem cells. Int. J. Biol. Sci. 2016, 12, 1372–1381. [Google Scholar] [CrossRef]
  79. Kakarougkas, A.; Ismail, A.; Chambers, A.L.; Riballo, E.; Herbert, A.D.; Kunzel, J.; Lobrich, M.; Jeggo, P.A.; Downs, J.A. Re-quirement for PBAF in transcriptional repression and repair at DNA breaks in actively transcribed regions of chromatin. Mol. Cell 2014, 55, 723–732. [Google Scholar] [CrossRef]
  80. Hopson, S.; Thompson, M.J. BAF180: Its Roles in DNA Repair and Consequences in Cancer. ACS Chem. Biol. 2017, 12, 2482–2490. [Google Scholar] [CrossRef] [PubMed]
  81. Porter, E.G.; Dhiman, A.; Chowdhury, B.; Carter, B.C.; Lin, H.; Stewart, J.C.; Kazemian, M.; Wendt, M.K.; Dykhuizen, E.C. PBRM1 Regulates Stress Response in Epithelial Cells. Iscience 2019, 15, 196–210. [Google Scholar] [CrossRef] [PubMed]
  82. Yan, Z.; Cui, K.; Murray, D.M.; Ling, C.; Xue, Y.; Gerstein, A.; Parsons, R.; Zhao, K.; Wang, W. PBAF chromatin-remodeling complex requires a novel specificity subunit, BAF200, to regulate expression of selective interferon-responsive genes. Genes Dev. 2005, 19, 1662–1667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Harte, M.T.; O’Brien, G.J.; Ryan, N.M.; Gorski, J.J.; Savage, K.I.; Crawford, N.T.; Mullan, P.B.; Harkin, D.P. BRD7, a Subunit of SWI/SNF Complexes, Binds Directly to BRCA1 and Regulates BRCA1-Dependent Transcription. Cancer Res 2010, 70, 2538–2547. [Google Scholar] [CrossRef]
  84. Hagiwara, M.; Fushimi, A.; Yamashita, N.; Bhattacharya, A.; Rajabi, H.; Long, M.D.; Yasumizu, Y.; Oya, M.; Liu, S.; Kufe, D. MUC1-C activates the PBAF chromatin remodeling complex in integrating redox balance with progression of human prostate cancer stem cells. Oncogene 2021, 40, 4930–4940. [Google Scholar] [CrossRef]
  85. Vitale, I.; Manic, G.; De Maria, R.; Kroemer, G.; Galluzzi, L. DNA Damage in Stem Cells. Mol. Cell 2017, 66, 306–319. [Google Scholar] [CrossRef]
  86. Tanabe, A.; Sahara, H. The Metabolic Heterogeneity and Flexibility of Cancer Stem Cells. Cancers 2020, 12, 2780. [Google Scholar] [CrossRef]
  87. De La Vega, M.R.; Chapman, E.; Zhang, D.D. NRF2 and the Hallmarks of Cancer. Cancer Cell 2018, 34, 21–43. [Google Scholar] [CrossRef] [PubMed]
  88. He, F.; Antonucci, L.; Karin, M. NRF2 as a regulator of cell metabolism and inflammation in cancer. Carcinogenesis 2020, 41, 405–416. [Google Scholar] [CrossRef]
  89. Lim, J.K.M.; Delaidelli, A.; Minaker, S.W.; Zhang, H.F.; Colovic, M.; Yang, H.; Negri, G.L.; von Karstedt, S.; Lockwood, W.W.; Schaffer, P.; et al. Cystine/glutamate antiporter xCT (SLC7A11) facilitates oncogenic RAS transformation by preserving intra-cellular redox balance. Proc. Natl. Acad. Sci. USA 2019, 116, 9433–9442. [Google Scholar] [CrossRef]
  90. Bhattacharya, A.; Fushimi, A.; Yamashita, N.; Hagiwara, M.; Morimoto, Y.; Rajabi, H.; Long, M.D.; Abdulla, M.; Ahmad, R.; Street, K.; et al. MUC1-C Dictates JUN and BAF-Mediated Chromatin Remodeling at Enhancer Signatures in Cancer Stem Cells. Mol. Cancer Res. 2022, 20, 556–567. [Google Scholar] [CrossRef]
  91. Larsen, S.B.; Cowley, C.J.; Sajjath, S.M.; Barrows, D.; Yang, Y.; Carroll, T.S.; Fuchs, E. Establishment, maintenance, and recall of inflammatory memory. Cell Stem. Cell 2021, 28, 1758–1774.e8. [Google Scholar] [CrossRef]
  92. Hagiwara, M.; Fushimi, A.; Bhattacharya, A.; Yamashita, N.; Morimoto, Y.; Oya, M.; Withers, H.G.; Hu, Q.; Liu, T.; Liu, S.; et al. MUC1-C integrates type II interferon and chromatin remodeling pathways in immunosuppression of prostate cancer. Oncoimmunology 2022, 11, e2029298. [Google Scholar] [CrossRef] [PubMed]
  93. Stark, G.R.; Cheon, H.; Wang, Y. Responses to Cytokines and Interferons that Depend upon JAKs and STATs. Cold Spring Harb. Perspect. Biol. 2018, 10, a028555. [Google Scholar] [CrossRef] [PubMed]
  94. Prendergast, G.C.; Malachowski, W.P.; DuHadaway, J.B.; Muller, A.J. Discovery of IDO1 Inhibitors: From Bench to Bedside. Cancer Res. 2017, 77, 6795–6811. [Google Scholar] [CrossRef] [PubMed]
  95. Adam, I.; Dewi, D.L.; Mooiweer, J.; Sadik, A.; Mohapatra, S.R.; Berdel, B.; Keil, M.; Sonner, J.K.; Thedieck, K.; Rose, A.J.; et al. Up-regulation of tryptophanyl-tRNA synthethase adapts human cancer cells to nutritional stress caused by tryptophan degra-dation. Oncoimmunology 2018, 7, e1486353. [Google Scholar] [CrossRef] [PubMed]
  96. Wu, A.A.; Drake, V.; Huang, H.S.; Chiu, S.; Zheng, L. Reprogramming the tumor microenvironment: Tumor-induced immu-nosuppressive factors paralyze T cells. Oncoimmunology 2015, 4, e1016700. [Google Scholar] [CrossRef]
  97. Rubin, M.A.; Bristow, R.G.; Thienger, P.D.; Dive, C.; Imielinski, M. Impact of Lineage Plasticity to and from a Neuroendocrine Phenotype on Progression and Response in Prostate and Lung Cancers. Mol. Cell 2020, 80, 562–577. [Google Scholar] [CrossRef]
  98. Balanis, N.G.; Sheu, K.M.; Esedebe, F.N.; Patel, S.J.; Smith, B.A.; Park, J.W.; Alhani, S.; Gomperts, B.N.; Huang, J.; Witte, O.N.; et al. Pan-cancer convergence to a small-cell neuroendocrine phenotype that shares susceptibilities with hematological malig-nancies. Cancer Cell 2019, 36, 17–34.e7. [Google Scholar] [CrossRef]
  99. Luan, Z.; Morimoto, Y.; Fushimi, A.; Yamashita, N.; Suo, W.; Bhattacharya, A.; Hagiwara, M.; Jin, C.; Kufe, D. MUC1-C dictates neuroendocrine lineage specification in pancreatic ductal adenocarcinomas. Carcinogenesis 2021, 43, 67–76. [Google Scholar] [CrossRef]
  100. Fushimi, A.; Morimoto, Y.; Ishikawa, S.; Yamashita, N.; Bhattacharya, A.; Daimon, T.; Rajabi, H.; Jin, C.; Hagiwara, M.; Yasumizu, Y.; et al. Dependence on the MUC1-C oncoprotein in classic, variant and non-neuroendocrine small cell lung cancer. Mol. Cancer Res. 2022, 20, 1379–1390. [Google Scholar] [CrossRef]
  101. Morimoto, Y.; Fushimi, A.; Yamashita, N.; Hagiwara, M.; Bhattacharya, A.; Cheng, J.; Frost, T.C.; Ahmad, R.; Daimon, T.; Huang, L.; et al. Addiction of Merkel cell carcinoma to MUC1-C identifies a potential new target for treatment. Oncogene 2022, 41, 3511–3523. [Google Scholar] [CrossRef] [PubMed]
  102. Panchamoorthy, G.; Jin, C.; Raina, D.; Bharti, A.; Yamamoto, M.; Adeebge, D.; Zhao, Q.; Bronson, R.; Jiang, S.; Li, L.; et al. Targeting the human MUC1-C oncoprotein with an antibody-drug conjugate. J. Clin. Investig. 2018, 3, e99880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The activation of MUC1-C by loss of homeostasis contributes to wound healing and progression to cancer. The transmembrane MUC1-C subunit is expressed at the apical borders of polarized epithelial cells where it is poised to respond to stress. The activation of MUC1-C in response to loss of homeostasis induces the Yamanaka pluripotency factors, EMT and epigenetic reprogramming. MUC1-C also contributes to inflammatory, proliferative and remodeling responses associated with wound repair. These responses are, in principle, reversible with healing; however, prolonged activation of MUC1-C in settings of chronic inflammation with the remodeling of chromatin drive progression to cancer. Figure modified from [43].
Figure 1. The activation of MUC1-C by loss of homeostasis contributes to wound healing and progression to cancer. The transmembrane MUC1-C subunit is expressed at the apical borders of polarized epithelial cells where it is poised to respond to stress. The activation of MUC1-C in response to loss of homeostasis induces the Yamanaka pluripotency factors, EMT and epigenetic reprogramming. MUC1-C also contributes to inflammatory, proliferative and remodeling responses associated with wound repair. These responses are, in principle, reversible with healing; however, prolonged activation of MUC1-C in settings of chronic inflammation with the remodeling of chromatin drive progression to cancer. Figure modified from [43].
Ijms 24 03719 g001
Figure 3. MUC1-C activates the esBAF chromatin remodeling complex in driving the NEPC CSC state. In parallel with the induction of the BRN2 and NE phenotype, MUC1-C interacts with E2F1 and induces the expression of the esBAF subunits. MUC1-C forms a nuclear complex with BRG1 and ARID1A, which activates the NOTCH1 gene, NOTCH1 expression and NOTCH1 target genes. MUC1-C-induced activation of esBAF also promotes NANOG expression and self-renewal capacity. These findings demonstrate that MUC1-C integrates the MYC and E2F1 pathways in driving NEPC dedifferentiation. Figure modified from [73].
Figure 3. MUC1-C activates the esBAF chromatin remodeling complex in driving the NEPC CSC state. In parallel with the induction of the BRN2 and NE phenotype, MUC1-C interacts with E2F1 and induces the expression of the esBAF subunits. MUC1-C forms a nuclear complex with BRG1 and ARID1A, which activates the NOTCH1 gene, NOTCH1 expression and NOTCH1 target genes. MUC1-C-induced activation of esBAF also promotes NANOG expression and self-renewal capacity. These findings demonstrate that MUC1-C integrates the MYC and E2F1 pathways in driving NEPC dedifferentiation. Figure modified from [73].
Ijms 24 03719 g003
Figure 4. MUC1-C→E2F1 signaling integrates the activation of the esBAF and PBAF chromatin remodeling complexes. MUC1-C forms nuclear complexes with E2F1 that activate the expression of the esBAF and PBAF subunits. MUC1-C associates with PBRM1 and NRF2 in increasing the chromatin accessibility of NRF2 target genes, including SLC7A11, G6PD and PGD, that regulate redox balance. MUC1-C-induced activation of PBRM1/PBAF also contributes to the expression of the OSKM + NANOG pluripotency factors and integration with the ARID1A/esBAF complex that drives EMT, NOTCH1 and the NEPC CSC state. Figure modified from [84].
Figure 4. MUC1-C→E2F1 signaling integrates the activation of the esBAF and PBAF chromatin remodeling complexes. MUC1-C forms nuclear complexes with E2F1 that activate the expression of the esBAF and PBAF subunits. MUC1-C associates with PBRM1 and NRF2 in increasing the chromatin accessibility of NRF2 target genes, including SLC7A11, G6PD and PGD, that regulate redox balance. MUC1-C-induced activation of PBRM1/PBAF also contributes to the expression of the OSKM + NANOG pluripotency factors and integration with the ARID1A/esBAF complex that drives EMT, NOTCH1 and the NEPC CSC state. Figure modified from [84].
Ijms 24 03719 g004
Figure 5. MUC1-C integrates the induction of the esBAF and PBAF chromatin remodeling complexes with the chronic activation of the IFNG pathway and immunosuppression. MUC1-C activates the IFNGR1 gene by forming a complex with JUN and ARID1A that increases chromatin accessibility, H3K4 trimethylation and IFNGR1 expression. MUC1-C thereby contributes to upregulation of STAT1 and IRF1, and in turn interacts with IRF1 and PBRM1 to drive the expression of (i) IDO1, WARS and PTGES that metabolically suppress the TME, and (ii) ISG15 and SERPINB9, which inhibit T cell function. Consistent with the induction of these immunosuppressive effectors, MUC1 associates with immune cell-depleted cold TMEs. Figure modified from [92].
Figure 5. MUC1-C integrates the induction of the esBAF and PBAF chromatin remodeling complexes with the chronic activation of the IFNG pathway and immunosuppression. MUC1-C activates the IFNGR1 gene by forming a complex with JUN and ARID1A that increases chromatin accessibility, H3K4 trimethylation and IFNGR1 expression. MUC1-C thereby contributes to upregulation of STAT1 and IRF1, and in turn interacts with IRF1 and PBRM1 to drive the expression of (i) IDO1, WARS and PTGES that metabolically suppress the TME, and (ii) ISG15 and SERPINB9, which inhibit T cell function. Consistent with the induction of these immunosuppressive effectors, MUC1 associates with immune cell-depleted cold TMEs. Figure modified from [92].
Ijms 24 03719 g005
Table 1. Common MUC1-C Dependencies in NE Carcinomas.
Table 1. Common MUC1-C Dependencies in NE Carcinomas.
DependencyNEPCPDAC-NESCLCMCC
MYC FamilyMYCMYCMYCMYCL
Pluripotency FactorsOSKM
+
NANOG
OSKMND *OSKM
+
NANOG
Neural TFsBRN2
ASCL1
BRN2
ASCL1
NEUROD1
BRN2
ASCL1
BRN2
NEUROD1
ATOH1
Stemness FactorsNOTCH1
BMI1
NOTCH1
NOTCH2
NOTCH1
BMI1
BMI1
Tumorsphere
Formation
++++
Tumorigenicity++++
* ND: Not detected.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kufe, D. Dependence on MUC1-C in Progression of Neuroendocrine Prostate Cancer. Int. J. Mol. Sci. 2023, 24, 3719. https://doi.org/10.3390/ijms24043719

AMA Style

Kufe D. Dependence on MUC1-C in Progression of Neuroendocrine Prostate Cancer. International Journal of Molecular Sciences. 2023; 24(4):3719. https://doi.org/10.3390/ijms24043719

Chicago/Turabian Style

Kufe, Donald. 2023. "Dependence on MUC1-C in Progression of Neuroendocrine Prostate Cancer" International Journal of Molecular Sciences 24, no. 4: 3719. https://doi.org/10.3390/ijms24043719

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