Next Article in Journal
Unraveling the Transcriptional Dynamics of NASH Pathogenesis Affecting Atherosclerosis
Previous Article in Journal
Malectin Domain Protein Kinase (MDPK) Promotes Rice Resistance to Sheath Blight via IDD12, IDD13, and IDD14
Previous Article in Special Issue
BET Proteins as Attractive Targets for Cancer Therapeutics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Crucial Roles of Bmi-1 in Cancer: Implications in Pathogenesis, Metastasis, Drug Resistance, and Targeted Therapies

1
State Key Laboratory of Silkworm Genome Biology, Southwest University, Chongqing 400716, China
2
Cancer Center, Medical Research Institute, Southwest University, Chongqing 400716, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(15), 8231; https://doi.org/10.3390/ijms23158231
Submission received: 14 June 2022 / Revised: 22 July 2022 / Accepted: 23 July 2022 / Published: 26 July 2022
(This article belongs to the Special Issue Transcription in Cancer Initiation and Progression)

Abstract

:
B-cell-specific Moloney murine leukemia virus integration region 1 (Bmi-1, also known as RNF51 or PCGF4) is one of the important members of the PcG gene family, and is involved in regulating cell proliferation, differentiation and senescence, and maintaining the self-renewal of stem cells. Many studies in recent years have emphasized the role of Bmi-1 in the occurrence and development of tumors. In fact, Bmi-1 has multiple functions in cancer biology and is closely related to many classical molecules, including Akt, c-MYC, Pten, etc. This review summarizes the regulatory mechanisms of Bmi-1 in multiple pathways, and the interaction of Bmi-1 with noncoding RNAs. In particular, we focus on the pathological processes of Bmi-1 in cancer, and explore the clinical relevance of Bmi-1 in cancer biomarkers and prognosis, as well as its implications for chemoresistance and radioresistance. In conclusion, we summarize the role of Bmi-1 in tumor progression, reveal the pathophysiological process and molecular mechanism of Bmi-1 in tumors, and provide useful information for tumor diagnosis, treatment, and prognosis.

1. Introduction

Polycomb group (PcG) proteins are a group of transcriptional inhibitors that regulate targeted genes at the chromatin level. They also play a critical role in embryonic development, cell proliferation, and tumorigenesis [1,2]. PcG proteins act on the development of organisms by forming two multimeric protein complexes: the polycomb repressive complex 1 (PRC1) and the polycomb repressive complex 2 (PRC2)) [3]. B-cell-specific Moloney murine leukemia virus integration region 1 (Bmi-1, also known as RNF51 or PCGF4) is one of the core members of the PRC1 complex. This complex, acting through chromatin remodeling, is an essential epigenetic repressor involved in embryonic development and self-renewal in somatic stem cells [4,5,6]. Therefore, Bmi-1 also is an oncogene, and its abnormal expression is associated with tumorigenesis and drug resistance in many cancers, including bladder cancer cells, B-cell lymphoma cells, melanoma cells, and others. Noncoding RNAs (ncRNAs) are a class of RNAs produced in noncoding regions that are not translated into proteins but are widely expressed in organisms [7]. With the development of gene sequencing technology, increasing evidence suggests that ncRNAs have crucial biological functions and play core roles in regulating gene expression [8]. In recent years, it has been identified that microRNAs (miRNAs) play important roles in various biological processes by inhibiting the translation of mRNAs and inducing their degradation, which makes them potential molecular targets for cancer therapy [9,10,11,12,13]. Interestingly, several recent studies have found that Bmi-1 is targeted and regulated by many miRNAs in cancer. Furthermore, the unique biological functions and cancer-promoting effects of Bmi-1 have also attracted increased attention. Therefore, in this review, we first describe the normal function of Bmi-1; then, we review the relevant signaling pathways regulated by Bmi-1 in normal and carcinogenic conditions and its regulatory network with miRNAs, to provide a reference for a more comprehensive understanding of the role of Bmi-1 in cancer.

2. Molecular Features and Characteristics of Bmi-1

The structure of Bmi-1 in humans and mice is very similar, and the homologies of DNA and amino acid in Bmi-1 are 86% and 98%, respectively. The Bmi-1 gene in humans, composed of 10 exons and 9 introns, is localized on the short arm of chromosome 10 (10p11.23). The Bmi-1 protein is 37 kD in size, consists of 326 amino acids, and has a highly conserved structure. There are several important structures in the amino acid sequence of Bmi-1 protein. The N-terminal is a RING finger domain, which is composed of a zinc finger and C3hC4 sequence. Bmi-1 can regulate transcription, affect cell proliferation, and participate in the formation of the malignant tumors by RING finger binding to other critical proteins. The helix–turn–helix structure (HTH structure) is located in the center of the Bmi-1 protein, and mediates the transcriptional inhibition of Bmi-1 by binding with DNA. At the carboxyl terminus, the PEST structure of the PEST region situated at the C terminal contains many serine, glutamic acid, threonine, and proline residues [14], which are associated with intracellular turnover of Bmi-1 protein. Two nuclear localization signals (NLS1, NLS2) are contained in Bmi-1, of which NLS2 is required for Bmi-1 localization in the nucleus. Figure 1 illustrates the structure of the Bmi-1 gene and Bmi-1 protein, the illustration is redrawn (with modifications) from Sahasrabuddhe (2016) [15].

3. Upstream Regulatory Mechanisms of Bmi-1

Bmi-1 is expressed in almost every human tissue and in many cancers, and serves as a biomarker for some cancers. Therefore, it is crucial to explore the mechanisms that regulate Bmi-1 mRNA and Bmi-1 protein levels. Over the past decade, researchers have made some progress in the regulation of Bmi-1 at the transcriptional, post-transcriptional, and post-translational levels. It is currently known that Bmi-1 is included in the positive or negative transcriptional regulation of many transcription factors. Positively regulated transcription factors are Sp1, Twist1, FoxM1, ZEB1, E2F1, SALL4, MYC-N, c-MYC, and HDACs, while negatively regulated transcription factors are Mel18, Nanog, and KLF4 [15].
Post-translational modifications (PTMs) of proteins are an important part of proteomics. Proteins change their spatial conformation, activity, stability, and interaction performance with other molecules through PTMs, thereby participating in the regulation of various life activities in the body. Most proteins can undergo PTMs, and there are more than 200 known covalent modifications of proteins, including phosphorylation, nitrosylation, nitration, ubiquitination, and sumoylation, among others. PTMs are associated with many important diseases, including cancer. There have been some studies on the post-translational modification of Bmi-1.Bmi-1 can be activated by phosphorylation. For example, the state of Bmi-1 protein binding to chromosomes is related to its phosphorylation; in G1/S phase, non-phosphorylated Bmi-1 specifically binds to chromosomes, while in G2/M phase, phosphorylated Bmi-1 does not bind to chromosomes [16]. Bmi-1 represses the expression of the Ink4a-Arf locus, which encodes two tumor suppressor genes, p16Ink4a and p19Arf. Bmi-1 is phosphorylated (and inactivated) at Ser 316 by Akt serine/threonine kinase, also known as protein kinase B (PKB), to promote p16Ink4a and p19Arf expression and inhibit hematopoietic stem cell self-renewal [17]. Similarly, the Wnt5a-ROR1 complex can phosphorylate Bmi-1 through the Akt pathway to promote tumorigenesis [18]. It has also been reported that activation or overexpression of MAPKAP kinase 3 leads to the phosphorylation of Bmi-1 and other PcG members, and promotes their dissociation from chromatin [19]. Sumoylation is a reversible post-translational modification which has emerged as a crucial molecular regulatory mechanism, involved in the regulation of replication, cell-cycle progression, protein transport and the DNA damage response. The study found that DNA damage can induce the sumoylation of lysine 88 in Bmi-1, and knockout of chromobox 4 (CBX4), a member of the polycomb group family of epigenetic regulatory factors, can eliminate this phenomenon, which means that CBX4 can mediate the sumoylation of Bmi-1 [20]. Moreover, p53/p21 can inversely regulate Bmi-1 expression by affecting ubiquitin/proteasome activity [21]. This also implies that Bmi-1 may be ubiquitinated. Furthermore, Ser255 was found to be the site of O-GlcNAcylation of Bmi-1 in prostate cancer, and O-GlcNAcylation of Bmi-1 promoted the stability of Bmi-1 protein and its oncogenic activity [22].
MicroRNAs (miRNAs or miRs) are a class of small of ncRNAs that contain about 19–24 nt. Although miRNAs do not encode proteins, they regulate the expression levels of some genes at the post-transcriptional stage. MiRNAs negatively regulate the expression of target genes by complete or incomplete complementary binding to the 3′-UTR of target mRNAs, promoting the degradation or translational repression of target mRNAs. There is growing evidence that miRNAs are involved in essential biological processes, including development, differentiation, proliferation, apoptosis, invasion, and metastasis. The 3′UTR length of Bmi-1 is 2090 nt. The miRNAs mainly target two spacers of the 3′UTR of Bmi-1. The first target position is 469–725, which includes miR-128, miR-221, miR-183, and miR-200b/c; the other position is 1442–1758, including miR-203, miR-218. We summarize the interaction between Bmi-1 and miRNAs in human tumors in Table 1, to provide information for further research on cancer treatment and clinical applications.

4. Bmi-1-Targeted Processes in Cancer

Bmi-1 protein belongs to the polycomb family of transcriptional repressors, and binds to polycomb response elements in the genome to silence the transcription of specific target genes. The genes to be transcriptionally regulated by Bmi-1 include INK4a/ARF, Pten, homeobox protein HoxC13, human telomerase reverse transcriptase (hTERT), etc. [76,77,78,79]. Figure 2 illustrates the downstream genes directly regulated by Bmi-1. In addition, Bmi-1 is involved in many classical signaling pathways, for example mTOR, NF-κB, PI3K/Akt, and other signaling pathways. The function of Bmi-1 in tumors has been identified in various pathological processes, including abnormal cell proliferation, evasion of apoptosis, migration of cancer cells, and stemness maintenance of cancer stem cells (CSCs). Therefore, we provide a corresponding review of the key target genes and signaling pathways behind these processes.

4.1. Bmi-1 in Cancer Proliferation

Hyperproliferation is a typical feature of cancer progression, manifested by altered expression and activity of cell cycle-related proteins. The most classic downstream target of Bmi-1 is INK4a/ARF. The tumor suppressor gene INK4a/ARF can encode two regulatory genes: p16INK4a and p14ARF (p19ARF in mice). Their normal and orderly expression is the key to maintaining the balance of the cell cycle. On the one hand, P16INK4a arrests cells in the G0/G1 phase through cyclinD-CDK4/6-pRb-E2F. Mechanistically, when Bmi-1 is deficient, p16Ink4a is up-regulated, which prevents the binding of CDK4/6 to cyclinD, leading to phosphorylation of Rb. Phosphorylated Rb binds to E2F transcription factor to inhibit E2F transcription-factor-mediated transcription, which arrests the cell cycle in the G0/G1 phase [76]. On the other hand, p14ARF regulates the cell cycle through the MDM2/p53 pathway [15,84,85,86,87]. In detail, the p14ARF (p19ARF) protein can antagonize the ubiquitin-protein ligase MDM2 and p53-dependent transcription, thereby stabilizing tumor protein p53, causing cell apoptosis [15,84,85,86,87]. Bmi-1-shRNA reduces the expression of cyclin D1 and increases the expression of cyclin dependent kinase inhibitor p21 and p27 through the p16Ink4a independent pathway, to arrest lung adenocarcinoma cells in the G0/G1 phase [88]. USP22, an oncogene, could actively and effectively participate in the regulation of the cell cycle through the INK4a/ARF signaling pathway mediated by Bmi-1 in human colorectal cancer cells [89]. Similar phenomena have also been detected in gallbladder carcinoma, HeLa cells, lung cancer, esophageal carcinoma cell, etc. [90,91,92,93]. Bmi-1 can also activate the PI3K/mTOR/4EBP1 signaling pathway in ovarian cancer cells to regulate cell proliferation [94]. N-acetylglucosamine transferase (OGT) is the only known enzyme to catalyze the O-GlcNAcylation in humans. O-GlcNAcylation can promote the stability and oncogenic activity of Bmi-1 [22]. MiR-485-5p can regulate the O-GlcNAcylation level and the stability of Bmi-1 by inhibiting OGT, and then inhibit the proliferation of colorectal cancer cells [28]. Nevertheless, some studies have shown that Bmi-1 does not affect the cell cycle of lung cancer cells and the expression of p16/p19, PTEN, Akt, or p-Akt [95]. This may be related to cell type.

4.2. Bmi-1 in Cancer Apoptosis

Strong anti-apoptotic ability is one of the characteristics of cancer cells. Bmi-1 directly and indirectly regulates cell apoptosis via various pathways. The first pathway is the p14ARF/MDM2/p53 signal pathway mentioned above. Inhibition of Bmi-1 can promote the expression of p14ARF (p19ARF), and p14ARF (p19ARF) can antagonize the ubiquitin-protein ligase MDM2 to stabilize p53 and cause apoptosis. This phenomenon has been reported in many studies [29,96,97]. A second pathway involves the inhibition of Bmi-1, causing abnormal mitochondrial function and increasing the level of ROS, leading to apoptosis. For example, a recent study found that sodium butyrate induced ROS-mediated apoptosis by inhibiting the expression of Bmi-1 through miR-139-5p [13]. The third pathway is the Bmi-1 regulation of cell apoptosis by the NF-κB pathway. Expression of Bmi-1 could protect glioma cells from apoptosis by activating the NF-κB pathway [98]. Consistently, overexpression of Bmi-1 can increase cisplatin-induced apoptosis resistance in osteosarcoma by activating the NF-κB signal pathway [99]. In addition, the loss of Bmi-1 leads to impaired repair of DNA double-strand breaks by homologous recombination, ultimately leading to apoptosis [96]. Inhibition of Bmi-1 reduces the ubiquitination of myeloid cell leukemia sequence 1 (Mcl-1) by downregulating deubiquitinating enzyme (DUB3), resulting in cancer cell apoptosis [99].

4.3. Bmi-1 in Cancer Autophagy

Autophagy is a conserved self-degradation system that is critical for maintaining cellular homeostasis during stress conditions. Autophagy plays a dichotomous role in cancer by suppressing benign tumor growth but promoting advanced cancer growth [100]. Bmi-1 also affects cellular autophagy generally through PI3K/Akt and AMPK signaling pathways. Knockdown of Bmi-1 induced autophagy in ovarian cancer cells via ATP depletion [101]. Sodium butyrate induced AMPK/mTOR pathway-dependent autophagy via the miR-139-5p/Bmi-1 axis in human bladder cancer cells [13]. Knockdown of Bmi-1 in breast cancer cells also induced autophagy [102,103]. Additionally, betulinic acid induced autophagy and apoptosis in bladder cancer cells through the Bmi-1/ROS/AMPK/mTOR/ULK1 axis [86]. Similarly, Bmi-1 overexpression promoted migration and proliferation of cardiac fibroblasts by regulating the Pten/PI3K/Akt/mTOR signaling pathway [104]. In addition, the expression of Bmi-1 targets the inhibition of cyclinG2 expression, and cyclinG2 acts by disrupting the phosphatase 2A complex, which activates the PKCζ/AMPK/JNK/ERK pathway involved in autophagy [105].

4.4. Bmi-1 in Cancer EMT

Cancer cells have the characteristics of invasion into surrounding tissues and metastasis to distant tissues, and they continue to grow in these organs, destroying the function of the organs, and finally causing the death of the patient. The expression of Bmi-1 is closely related to tumor migration and invasion [93,95,106,107,108,109]. It has been reported that down-regulation of Bmi-1 reduces the migration and invasion of endometrial cancer cells in vivo and in vitro, by increasing the expression levels of E-cadherin and keratin, and down-regulating N-Cadherin, vimentin, and SLUG [106]. Likewise, Bmi-1 induces miR-27a and miR-155 to negatively regulate the expression of raf kinase inhibitory protein (RKIP), which promotes the migration and invasion of gastric cancer cells [107]. Additionally, Bmi-1 promotes the migration and invasion of Hepatocellular carcinoma by inhibiting phosphatase and tensin homolog (PTEN) and activating the PI3K/Akt pathway, and increasing the expression of MMP2, MMP9 and VEGF [110]. Similar phenomena have also been found in breast cancer, colon cancer, esophageal cancer cells [107]. Moreover, Bmi-1 increases luciferase activity of the MMP9 promoter-driven reporter gene containing an NF-κB binding site, which promotes MMP9 transcription levels in glioma cells [111]. Bmi-1 may also regulate the TLR4/MD2/MyD88 complex-mediated NF-κB signaling pathway to participate in colorectal cancer cell EMT [112]. Furthermore, miR-218, miR-330-3p, and miR-498 may regulate the migration and invasion of cancer cells through the Bmi-1/Akt axis [112].

4.5. Bmi-1 in Cancer DNA Damage Response

An underlying feature of cancer is genomic instability, which is associated with the accumulation of DNA damage. Bmi-1 plays an important role in the regulation of DNA damage response (DDR) [113,114]. The lack of Bmi-1 in cells could cause mitochondrial dysfunction, and at the same time, the DDR pathway could be initiated with the increase of ROS [115]. Bmi-1 enhances the repair of damaged DNA through epigenetic mechanisms to reduce the genotoxic effects of ionizing radiation (IR) [116]. Bmi-1 is a component of PRC1, and affects the expression of many genes through histone H2A ubiquitination. Mechanically, Bmi-1 binds to ring2/ring1b subunit to form a functional E3 ubiquitin ligase, and inhibits the expression of multiple gene through monoubiquitination of histone H2A in lysine 119 [117]. Through Bmi-1/RIN1b E3 ubiquitin ligase, Bmi-1 promotes histone H2A and γH2AX ubiquitination for the repair of double-stranded DNA breaks by stimulating homologous recombination and non-homologous end connection [117,118]. In addition, the knockout of Bmi-1 further aggravated the phosphorylation of checkpoint kinase 2 (Chk2) and H2AX induced by cisplatin treatment [119].

4.6. Bmi-1 in Cancer Inflammation

Inflammation is often associated with the progression of cancer. Cells responsible for cancer-related inflammation are genetically stable and therefore do not rapidly emerge drug resistance. Therefore, targeting inflammation is an attractive strategy for both cancer prevention and cancer therapy [120]. NF-κB transcription factors are major mediators of inflammatory processes and key players in innate and adaptive immune responses. Therefore, NF-κB pathway has been considered as the target of anti-inflammatory drugs. Recent studies have found that Bmi-1 is involved in the regulation of NF-κB signaling pathway. For example, Bmi-1 promotes tumor proliferation, metastasis and drug resistance by activating NF-κB signaling pathway [121]. Similarly, Bmi-1 promotes glioma invasion by activating NF-κB/MMP3 or NF-κB/MMP9 signaling pathways [111,122]. Moreover, Bmi-1 promote colorectal cancer migration and EMT in an inflammatory microenvironment by regulating TLR4/MD2/MyD88 complex-mediated NF-κB signaling pathway [112]. Furthermore, the expression of Bmi-1 can activate the NF-κB signaling pathway and increase the expression of vascular endothelial growth factor C (VEGF-C)to promote glioma angiogenesis [123]. Additionally, toll-like receptor 4 (TLR4) promotes inflammation by inhibiting Bmi-1 to activate the NOD-like receptors (NLRs) family member (NLRP3) pathway [124].

4.7. Bmi-1 in Cancer Stem Cells

Stem cells have self-renewal ability, and produce at least one highly differentiated cell. Cancer stem cells refer to cancer cells with stem cell properties such as self-renewal and multicellular differentiation. At present, many studies have proved that CSCs are present in a variety of tumors, such as leukemia, breast cancer, lung cancer, glioblastoma, colon cancer, liver cancer, and others [125]. Increasing evidence has indicated that Bmi-1 plays a critical role in the self-renewal and differentiation of cancer stem cells [6,126].
Revealing the molecular mechanism of the self-renewal of neural stem cells (NSCs) is one of the main goals for understanding the homeostasis of the adult brain. Neural stem cells and progenitor cells (NSPC) strictly regulate self-renewal potential to maintain homeostasis in the brain. The researchers found that the absence of Bmi-1 led to the decline of self-renewal ability in NSCs [127,128]. Molofsky et al. further found that Bmi-1 promoted the self-renewal of NSCs by inhibiting INK4a/ARF1. One study found that Bmi-1-knockout mice lacked stem cells and developed defects in the central nervous system, but no significant effect on the proliferation of progenitor cells was recorded. The results showed that Bmi-1 is a key factor for maintaining the stability of NSCs, but had little influence on differentiated cells [129,130]. Furthermore, Protein S (PROS1) can affect the self-renewal of adult hippocampal NSPC by down-regulating Bmi-1 [131]. In summary, Bmi-1 is a vital factor for the self-renewal of NSCs.
Bmi-1 also plays a pivotal role in hematopoietic stem cells (HSCs). Park et al. reported that Bmi-1 is strongly expressed in fetal mouse and adult human HSCs. Thus, Bmi-1 knockout resulted in the reduction of proliferation of mouse bone marrow, resulting in the death of most of the studied mice before adulthood [5]. Lessard et al. [132] found that the number of stem cells in peripheral leukemia cells from Bmi-1 wild-type mice was significantly higher than that in Bmi-1 knockout mice. Subsequently, they found a decrease of self-renewal ability of HSCs and poor hematopoiesis in Bmi-1−/− mice. Mice were then injected with fetal liver HSCs (Bmi-1+/+ or Bmi-1−/−) after high doses of lethal radiation. The results showed that the hematopoietic ability of bone marrow was dependent on the expression of Bmi-1.
In breast cancer stem cells (BCSCs), Bmi-1 is a target of the miR-200 family and miR-128 family. Furthermore, Bmi-1 is also regulated by certain signaling pathways, such as the Hedgehog and Wnt pathways. The Hedgehog signaling pathway promotes self-renewal in mammary stem cells and BCSCs by up-regulating Bmi-1 [4]. Bmi-1 can activate the Wnt signaling pathway by inhibiting Dickkopfs (DKK), a Wnt inhibitor gene. DKK1 resulted in the up-regulation of c-MYC, which further contributed to the transcriptional self-activation of Bmi-1 [133].
Bmi-1 is a regulator of self-renewal in prostate stem cells and a marker in intestinal stem cells [134,135]. Bmi-1 is also expressed in mesenchymal stem cells isolated from the umbilical cord [136]. Therefore, further regulatory mechanisms of Bmi-1-promoting CSC need to be revealed, and its role in tumor promotion requires exploration, to provide new therapeutic strategies for anti-tumor therapy.

4.8. Bmi-1 in Tumor Microenvironment

The tumor microenvironment (TME) comprises various cell types (endothelial cells, fibroblasts, immune cells, etc.) and extra-cellular components (cytokines, growth factors, hormones, extracellular matrix, etc.) that surround tumor cells and are nourished by a vascular network. Tumor cells interact with the surrounding cells through the circulatory and lymphatic systems to influence the progression of cancer and therapeutic efficacy of treatments [137,138]. The expression of Bmi-1 can also affect the tumor microenvironment (TME). Selective Bmi-1 inhibitor PTC-209 increases the expression of DKK1 by down-regulating Bmi-1, impairing in vitro osteoclast formation and destroying the tumor microenvironment [139]. There is a report that Bmi-1 is upregulated in multiple myeloma-associated macrophages (MM-MΦs) and that Bmi-1 modulates MM-MΦ’s pro-myeloma functions. Bmi-1 inhibitors could not only target multiple myeloma (MM) cells, but also eliminate MM-MΦs in the treatment of myeloma [140]. Another report suggests that tumor-associated macrophages (TAMs) may cause increased Bmi-1 expression through miR-30e* suppression, leading to gastrointestinal cancer progression [141]. At present, there are few reports on the effect of Bmi-1 on the tumor microenvironment. Therefore, the role of Bmi-1 in the tumor microenvironment needs to be further explored.

5. Clinical Characteristics and Cancer Therapy of Bmi-1

5.1. Protein Expression and Clinical Characteristics of Bmi-1

In the past two decades, many researchers have discussed the relationship between the expression level of Bmi-1 and the clinical characteristics of different cancers. Although a few results are contradictory, most of the results prove the importance of Bmi-1 in the occurrence and development of cancer. In order to describe it more simply and intuitively, we summarized the expression level and clinical characteristics of Bmi-1 in Table 2.

5.2. Bmi-1 in Chemoresistance and Cancer Therapy

Radiation and chemotherapy are common treatments for cancer. Drug resistance is one of the main obstacles to effective anticancer treatment. Many clinical data have reported that Bmi-1 is related to the drug resistance of tumors [114,164,165,166,167]. Silencing Bmi-1 could enhance camptothecin-induced DNA double-strand breaks and promote camptothecin-induced apoptosis. Conversely, increasing Bmi-1 can significantly reduce camptothecin-induced apoptosis [168]. Direct inhibition of Bmi-1 abrogates head and neck cancer stem-cell self-renewal and increases tumor sensitivity to cisplatin [169]. Hypoxic exposure regulates PI3K/Akt signal and EMT through activation of the HIF-1α/Bmi-1 signal to induce LSCs drug resistance [170]. Similarly, Bmi-1 activates the PI3K/Akt pathway and NF–κB pathway to promote cell growth and resistance to cisplatin treatment [170]. Furthermore, the low expression of Survivin is considered to be one of the factors for the success of chemotherapy; Bmi-1 promotes the drug resistance of B-cell lymphoma cells through the regulation of Survivin [171]. Although radiation therapy is one of the main methods of cancer treatment, tumors often acquire radioresistance, which leads to radiotherapy failure. Bmi-1 depletion makes radiation-resistant ESCC cells sensitive to radiotherapy by inducing apoptosis, senescence, ROS production, and oxidase gene expression (Lpo, Noxo1 and Alox15), reducing DNA repair capabilities [172]. Interestingly, the inhibition of Bmi-1 activates immune responses in tumor cells and recruits CD8+ T cells. Mechanistically, knockdown of Bmi-1 not only eliminated CSCs, but also sensitized tumor cells to anti-PD-1 antibodies by recruiting CD8+ T cells [173]. In conclusion, Bmi-1 has development potential as a target for the treatment of cancer. Perhaps in future the development of small-molecule compounds that effectively target Bmi-1 will be an option for anticancer drug development.
At present, certain small molecule inhibitors of Bmi-1 have been studied, and the most commonly used small molecule inhibitor of Bmi-1 is PTC-209. An article published in 2013 reported that researchers had identified a low molecular compound PTC-209 for the first time. They also found that PTC-209 reduced the tumorigenic activity and volume of xenograft tumors and the number of functional colorectal cancer-initiating cells in mice, after short-term treatment [174]. Subsequently, the drug has also been studied in other cancer cells and has demonstrated its anti-tumor effect, including for head neck squamous cell carcinoma (HNSCC) [175], prostate cancer [176], myeloma [177], glioblastoma [178], and cervical cancer cells [179]. In 2019, researchers developed an orally active nanoparticle PTC-209 vector, which can significantly inhibit colony formation and migration of colon cancer cells, and inhibit tumor progression and metastasis in orthotopic tumor-bearing mice by reversing stemness [180]. An orally active and selective Bmi-1 inhibitor, PTC-596, downregulated bcl2 family apoptosis regulator Mcl-1 and induces p53-independent mitochondrial apoptosis in AML progenitor cells [181]. The drug has also been explored for treatment of other cancers, for example, mantle cell lymphoma [182], cancer stem-like cells [99], multiple myeloma [183], and glioblastoma multiforme [184]. PTC-596 has been used in clinical phase 1 for conditions including diffuse intrinsic pontine glioma and leiomyosarcoma [185,186]. Moreover, a novel inhibitor, RU-A1, was developed based on the RNA three-dimensional (3D) structure of Bmi-1 and showed stronger potency than PTC-209 in targeting CSCs [187]. PTC-028, an orally bioavailable compound, mediates hyperphosphorylation of Bmi-1 accompanied by a transient decrease in ATP and disruption of mitochondrial redox balance, enhancing caspase-dependent apoptosis. In an orthotopic mouse model of ovarian cancer, oral administration of PTC-028 as a single agent demonstrated significant antitumor activity comparable to standard cisplatin/paclitaxel therapy [188]. Researchers have also discovered another small molecule compound QW24, which can inhibit the expression of Bmi-1 through the lysosomal autophagy pathway, and can inhibit the growth of mouse xenograft tumors and liver metastasis, and prolong the survival period of mice [189]. Unfortunately, there have been no other reports of the anticancer effects of QW24, so its anti-cancer effect needs to be further verified. Table 3 summarizes the inhibitors of Bmi-1, the ways they regulate Bmi-1, and the regulatory mechanisms in tumors.

6. Future Research and Conclusions

PcG family member Bmi-1 plays a vital role in the proliferation, apoptosis, metastasis, and chemical sensitivity of cancer cells. The discovery of the biogenesis and function of non-coding RNA has improved our understanding of the complexity of the human genome. With the deepening of research, it has been found that non-coding RNA is involved in the regulation of the cell cycle, proliferation, and differentiation, especially in the occurrence and development of cancer. Many miRNAs inhibit Bmi-1 expression by targeting the 3′UTR [22]. That means that these miRNAs can also affect cancer proliferation, migration, invasion, apoptosis, and drug sensitivity [22]. Thus, the interaction between miRNA and Bmi-1 plays an important role in the occurrence and development of cancer. This also provides new strategies for cancer treatment. However, at present, the research on the interaction between miRNA and Bmi-1 remains in the initial stages, and there are many problems still to be solved.
As a proto-oncogene, Bmi-1 has been confirmed to be highly expressed in a variety of tumors. It is related to the clinical stage, pathological classification, and lymph node metastasis, and can be used as one of the predictors for prognosis and recurrence in tumor patients. With the development of molecular biological technology and the application of gene chip technology, development is expected of a microfluidic multi-indicator joint inspection chip, a circulating cancer-cell capture chip, and other devices for automatic detection of Bmi-1 expression [208,209,210]. In addition, Bmi-1 is involved in the occurrence and development of many tumors, and the targeted therapy of cancer stem cells is an important measure for future development. However, there is currently no drug that can be used clinically to target Bmi-1 specifically. Therefore, the discovery of new small-molecule compounds that specifically inhibit Bmi-1 will be the focus of future research.

Author Contributions

L.Y. and H.C. contributed to the conception and design of the study. J.X. collected the related papers and was a major contributor in writing the manuscript. L.L. generated the table and prepared the figures. P.S. revised the article. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Fundamental Research Funds for the Central Universities (XDJK2020B006).

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

4EBP1Eukaryotic Translation Initiation Factor 4E Binding Protein 1
AKTProtein kinase B
AMLAcute myeloid leukemia
AMPKAMP-activated protein kinase
ATPAdenosine triphosphate
BCSCsBreast cancer stem cells
Bmi-1B Lymphoma Mo-MLV Insertion Region 1 Homolog
CBX4Chromobox 4
Chk2Checkpoint Kinase 2
c-MYCMyc proto-oncogene protein
CSCsCancer stem cells
DDRDNA damage response
DKK1Dickkopf WNT Signaling Pathway Inhibitor 1
DUB3Deubiquitinating Protein 3
E2F1E2F Transcription Factor 1
EMTEpithelial-mesenchymal-like transformation
FoxM1Forkhead Box M1
H2AXH2A.X Variant Histone
HDACsHistone Deacetylase
HoxC13Homeobox C13
HSCsHematopoietic stem cells
hTERTHuman telomerase reverse transcriptase
HTH structureHelix-turn-helix structure
IRIonizing radiation
KLF4Kruppel-like Factor 4
Mcl-1Myeloid cell leukemia sequence 1
MD-2Myeloid Differentiation Protein-2
MDM2E3 ubiquitin-protein ligase Mdm2
Mel18Polycomb Group RING Finger 2
miRNAsMicroRNAs
MMMultiple myeloma
MM-MΦsMyeloma-associated macrophages
MMP2Matrix Metallopeptidase2
MMP9Matrix Metallopeptidase 9
mTORmammalian target of rapamycin
MYC-NMYCN Proto-Oncogene, BHLH Transcription Factor
MyD88MYD88 Innate Immune Signal Transduction Adaptor
NanogHomeobox protein NANOG
ncRNAsNon-coding RNA
NF-κBNuclear factor kappa-B
NLRP3NACHT, LRR and PYD domains-containing protein 3
NLRP3NOD-like receptors (NLRs) family member
NLSNuclear localization signal
NoxaPhorbol-12-Myristate-13-Acetate-Induced Protein 1
NSCsNeural stem cells
NSPCNeural stem cells and progenitor cells
P21Cyclin-dependent kinase inhibitor 1A
P53Tumor Protein P53
PcGPolycomb group
PRC1polycomb repressive complex1
PRC2polycomb repressive complex 2
PROS1Protein S
PTENPhosphatidylinositol3,4,5-trisphosphate 3-phosphatase and dual-specificity protein phosphatase PTEN
PTMsPost-translational modifications
RKIPRaf kinase inhibitory protein
ROSProto-oncogene tyrosine-protein kinase ROS
SALL4Spalt Like Transcription Factor 4
SlugSnail Family Transcriptional Repressor 2
SnailZinc finger protein SNAI1
SP1Transcription factor Sp1
Stat3Signal transducer and activator of transcription 3
TAMsTumor-associated macrophages
TLR4Toll-like receptor 4
TMETumor microenvironment
Twist1Twist Family BHLH Transcription Factor 1
USP22Ubiquitin Specific Peptidase 22
VEGFVascular Endothelial Growth Factor A
VEGF-CVascular Endothelial Growth Factor C
WWOXWW Domain Containing Oxidoreductase
ZEB1Zinc Finger E-Box Binding Homeobox 1

References

  1. Wang, W.; Qin, J.J.; Voruganti, S.; Nag, S.; Zhou, J.; Zhang, R. Polycomb Group (PcG) Proteins and Human Cancers: Multifaceted Functions and Therapeutic Implications. Med. Res. Rev. 2015, 35, 1220–1267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kanno, R.; Janakiraman, H.; Kanno, M. Epigenetic regulator polycomb group protein complexes control cell fate and cancer. Cancer Sci. 2008, 99, 1077–1084. [Google Scholar] [CrossRef] [PubMed]
  3. Schwartz, Y.B.; Pirrotta, V. Polycomb complexes and epigenetic states. Curr. Opin. Cell Biol. 2008, 20, 266–273. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, S.L.; Dontu, G.; Mantle, I.D.; Patel, S.; Ahn, N.S.; Jackson, K.W.; Suri, P.; Wicha, M.S. Hedgehog signaling and Bmi-1 regulate self-renewal of normal and malignant human mammary stem cells. Cancer Res. 2006, 66, 6063–6071. [Google Scholar] [CrossRef] [Green Version]
  5. Park, I.K.; Qian, D.; Kiel, M.; Becker, M.W.; Pihalja, M.; Weissman, I.L.; Morrison, S.J.; Clarke, M.F. Bmi-1 is required for maintenance of adult self-renewing haematopoietic stem cells. Nature 2003, 423, 302–305. [Google Scholar] [CrossRef] [PubMed]
  6. Jiang, L.; Li, J.; Song, L. Bmi-1, stem cells and cancer. Acta Biochim. Biophys. Sin. 2009, 41, 527–534. [Google Scholar] [CrossRef] [Green Version]
  7. Gomes, A.Q.; Nolasco, S.; Soares, H. Non-coding RNAs: Multi-tasking molecules in the cell. Int. J. Mol. Sci. 2013, 14, 16010–16039. [Google Scholar] [CrossRef]
  8. Wei, J.W.; Huang, K.; Yang, C.; Kang, C.S. Non-coding RNAs as regulators in epigenetics. Oncol. Rep. 2017, 37, 3–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Xu, Z.J.; Yan, Y.L.; Qian, L.; Gong, Z.C. Long non-coding RNAs act as regulators of cell autophagy in diseases (Review). Oncol. Rep. 2017, 37, 1359–1366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Beermann, J.; Piccoli, M.T.; Viereck, J.; Thum, T. Non-coding RNAs in Development and Disease: Background, Mechanisms, and Therapeutic Approaches. Physiol. Rev. 2016, 96, 1297–1325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. de Almeida, R.A.; Fraczek, M.G.; Parker, S.; Delneri, D.; O’Keefe, R.T. Non-coding RNAs and disease: The classical ncRNAs make a comeback. Biochem. Soc. Trans. 2016, 44, 1073–1078. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, X.; Wang, C.; Zhang, X.; Hua, R.; Gan, L.; Huang, M.; Zhao, L.; Ni, S.; Guo, W. Bmi-1 regulates stem cell-like properties of gastric cancer cells via modulating miRNAs. J. Hematol. Oncol. 2016, 9, 90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, F.; Wu, H.; Fan, M.; Yu, R.; Zhang, Y.; Liu, J.; Zhou, X.; Cai, Y.; Huang, S.; Hu, Z.; et al. Sodium butyrate inhibits migration and induces AMPK-mTOR pathway-dependent autophagy and ROS-mediated apoptosis via the miR-139-5p/Bmi-1 axis in human bladder cancer cells. FASEB J. 2020, 34, 4266–4282. [Google Scholar] [CrossRef] [Green Version]
  14. Alkema, M.J.; Wiegant, J.; Raap, A.K.; Berns, A.; Vanlohuizen, M. Characterization And Chromosomal Localization Of the Human Protooncogene Bmi-1. Hum. Mol. Genet. 1993, 2, 1597–1603. [Google Scholar] [CrossRef] [PubMed]
  15. Sahasrabuddhe, A.A. BMI1: A Biomarker of Hematologic Malignancies. Biomark. Cancer 2016, 8, 65–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Voncken, J.W.; Schweizer, D.; Aagaard, L.; Sattler, L.; Jantsch, M.F.; van Lohuizen, M. Chromatin-association of the Polycomb group protein BMI1 is cell cycle-regulated and correlates with its phosphorylation status. J. Cell Sci. 1999, 112 Pt 24, 4627–4639. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, Y.; Liu, F.; Yu, H.; Zhao, X.; Sashida, G.; Deblasio, A.; Harr, M.; She, Q.B.; Chen, Z.; Lin, H.K.; et al. Akt phosphorylates the transcriptional repressor bmi1 to block its effects on the tumor-suppressing ink4a-arf locus. Sci. Signal. 2012, 5, ra77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Karvonen, H.; Barker, H.; Kaleva, L.; Niininen, W.; Ungureanu, D. Molecular Mechanisms Associated with ROR1-Mediated Drug Resistance: Crosstalk with Hippo-YAP/TAZ and BMI-1 Pathways. Cells 2019, 8, 812. [Google Scholar] [CrossRef] [Green Version]
  19. Voncken, J.W.; Niessen, H.; Neufeld, B.; Rennefahrt, U.; Dahlmans, V.; Kubben, N.; Holzer, B.; Ludwig, S.; Rapp, U.R. MAPKAP kinase 3pK phosphorylates and regulates chromatin association of the polycomb group protein Bmi1. J. Biol. Chem. 2005, 280, 5178–5187. [Google Scholar] [CrossRef] [Green Version]
  20. Ismail, I.H.; Gagné, J.P.; Caron, M.C.; McDonald, D.; Xu, Z.; Masson, J.Y.; Poirier, G.G.; Hendzel, M.J. CBX4-mediated SUMO modification regulates BMI1 recruitment at sites of DNA damage. Nucleic Acids Res. 2012, 40, 5497–5510. [Google Scholar] [CrossRef] [Green Version]
  21. Zhang, Z.; Oh, M.; Sasaki, J.I.; Nör, J.E. Inverse and reciprocal regulation of p53/p21 and Bmi-1 modulates vasculogenic differentiation of dental pulp stem cells. Cell Death Dis. 2021, 12, 644. [Google Scholar] [CrossRef] [PubMed]
  22. Li, Y.; Wang, L.; Liu, J.; Zhang, P.; An, M.; Han, C.; Li, Y.; Guan, X.; Zhang, K. O-GlcNAcylation modulates Bmi-1 protein stability and potential oncogenic function in prostate cancer. Oncogene 2017, 36, 6293–6305. [Google Scholar] [CrossRef] [PubMed]
  23. You, D.; Wang, D.; Liu, P.; Chu, Y.; Zhang, X.; Ding, X.; Li, X.; Mao, T.; Jing, X.; Tian, Z.; et al. MicroRNA-498 inhibits the proliferation, migration and invasion of gastric cancer through targeting BMI-1 and suppressing AKT pathway. Hum. Cell 2020, 33, 366–376. [Google Scholar] [CrossRef]
  24. Yu, W.W.; Jiang, H.; Zhang, C.T.; Peng, Y. The SNAIL/miR-128 axis regulated growth, invasion, metastasis, and epithelial-to-mesenchymal transition of gastric cancer. Oncotarget 2017, 8, 39280–39295. [Google Scholar] [CrossRef] [Green Version]
  25. Wu, C.; Zheng, X.; Li, X.; Fesler, A.; Hu, W.; Chen, L.; Xu, B.; Wang, Q.; Tong, A.; Burke, S.; et al. Reduction of gastric cancer proliferation and invasion by miR-15a mediated suppression of Bmi-1 translation. Oncotarget 2016, 7, 14522–14536. [Google Scholar] [CrossRef] [Green Version]
  26. Xu, L.; Li, Y.; Yan, D.; He, J.; Liu, D. MicroRNA-183 inhibits gastric cancer proliferation and invasion via directly targeting Bmi-1. Oncol. Lett. 2014, 8, 2345–2351. [Google Scholar] [CrossRef] [Green Version]
  27. Wu, Y.; Tian, S.; Chen, Y.; Ji, M.; Qu, Y.; Hou, P. miR-218 inhibits gastric tumorigenesis through regulating Bmi-1/Akt signaling pathway. Pathol. Res. Pract. 2019, 215, 243–250. [Google Scholar] [CrossRef]
  28. Chai, Y.; Du, Y.; Zhang, S.; Xiao, J.; Luo, Z.; He, F.; Huang, K. MicroRNA-485-5p reduces O-GlcNAcylation of Bmi-1 and inhibits colorectal cancer proliferation. Exp. Cell Res. 2018, 368, 111–118. [Google Scholar] [CrossRef] [PubMed]
  29. He, X.; Dong, Y.; Wu, C.W.; Zhao, Z.; Ng, S.S.; Chan, F.K.; Sung, J.J.; Yu, J. MicroRNA-218 inhibits cell cycle progression and promotes apoptosis in colon cancer by downregulating BMI1 polycomb ring finger oncogene. Mol. Med. 2013, 18, 1491–1498. [Google Scholar] [CrossRef]
  30. Xu, Z.; Tao, J.; Chen, P.; Chen, L.; Sharma, S.; Wang, G.; Dong, Q. Sodium Butyrate Inhibits Colorectal Cancer Cell Migration by Downregulating Bmi-1 Through Enhanced miR-200c Expression. Mol. Nutr. Food Res. 2018, 62, e1700844. [Google Scholar] [CrossRef]
  31. Zhou, L.; Zhang, W.G.; Wang, D.S.; Tao, K.S.; Song, W.J.; Dou, K.F. MicroRNA-183 is involved in cell proliferation, survival and poor prognosis in pancreatic ductal adenocarcinoma by regulating Bmi-1. Oncol. Rep. 2014, 32, 1734–1740. [Google Scholar] [CrossRef] [PubMed]
  32. Guo, S.; Xu, X.; Tang, Y.; Zhang, C.; Li, J.; Ouyang, Y.; Ju, J.; Bie, P.; Wang, H. miR-15a inhibits cell proliferation and epithelial to mesenchymal transition in pancreatic ductal adenocarcinoma by down-regulating Bmi-1 expression. Cancer. Lett. 2014, 344, 40–46. [Google Scholar] [CrossRef] [PubMed]
  33. Yu, X.; Jiang, X.; Li, H.; Guo, L.; Jiang, W.; Lu, S.H. miR-203 inhibits the proliferation and self-renewal of esophageal cancer stem-like cells by suppressing stem renewal factor Bmi-1. Stem Cells Dev 2014, 23, 576–585. [Google Scholar] [CrossRef]
  34. Kim, J.S.; Choi, D.W.; Kim, C.S.; Yu, S.K.; Kim, H.J.; Go, D.S.; Lee, S.A.; Moon, S.M.; Kim, S.G.; Chun, H.S.; et al. MicroRNA-203 Induces Apoptosis by Targeting Bmi-1 in YD-38 Oral Cancer Cells. Anticancer Res. 2018, 38, 3477–3485. [Google Scholar] [CrossRef]
  35. Wu, J.; Jiang, Z.M.; Xie, Y.; He, X.Q.; Lin, X.H.; Hu, J.L.; Peng, Y.Z. miR-218 suppresses the growth of hepatocellular carcinoma by inhibiting the expression of proto-oncogene Bmi-1. J. BUON 2018, 23, 604–610. [Google Scholar]
  36. Yang, F.; Lv, L.Z.; Cai, Q.C.; Jiang, Y. Potential roles of EZH2, Bmi-1 and miR-203 in cell proliferation and invasion in hepatocellular carcinoma cell line Hep3B. World J. Gastroenterol. 2015, 21, 13268–13276. [Google Scholar] [CrossRef]
  37. Shao, Y.; Zhang, D.; Li, X.; Yang, J.; Chen, L.; Ning, Z.; Xu, Y.; Deng, G.; Tao, M.; Zhu, Y.; et al. MicroRNA-203 Increases Cell Radiosensitivity via Directly Targeting Bmi-1 in Hepatocellular Carcinoma. Mol. Pharm. 2018, 15, 3205–3215. [Google Scholar] [CrossRef] [PubMed]
  38. Fu, W.M.; Tang, L.P.; Zhu, X.; Lu, Y.F.; Zhang, Y.L.; Lee, W.Y.; Wang, H.; Yu, Y.; Liang, W.C.; Ko, C.H.; et al. MiR-218-targeting-Bmi-1 mediates the suppressive effect of 1,6,7-trihydroxyxanthone on liver cancer cells. Apoptosis 2015, 20, 75–82. [Google Scholar] [CrossRef] [PubMed]
  39. Tu, K.; Li, C.; Zheng, X.; Yang, W.; Yao, Y.; Liu, Q. Prognostic significance of miR-218 in human hepatocellular carcinoma and its role in cell growth. Oncol. Rep. 2014, 32, 1571–1577. [Google Scholar] [CrossRef] [Green Version]
  40. Qi, X.; Li, J.; Zhou, C.; Lv, C.; Tian, M. MicroRNA-320a inhibits cell proliferation, migration and invasion by targeting BMI-1 in nasopharyngeal carcinoma. FEBS Lett. 2014, 588, 3732–3738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Liu, G.F.; Zhang, S.H.; Li, X.F.; Cao, L.Y.; Fu, Z.Z.; Yu, S.N. Overexpression of microRNA-132 enhances the radiosensitivity of cervical cancer cells by down-regulating Bmi-1. Oncotarget 2017, 8, 80757–80769. [Google Scholar] [CrossRef]
  42. Rong, X.; Gao, W.; Yang, X.; Guo, J. Downregulation of hsa_circ_0007534 restricts the proliferation and invasion of cervical cancer through regulating miR-498/BMI-1 signaling. Life Sci. 2019, 235, 116785. [Google Scholar] [CrossRef]
  43. Li, Q.; Wang, W.; Zhang, M.; Sun, W.; Shi, W.; Li, F. Circular RNA circ-0016068 Promotes the Growth, Migration, and Invasion of Prostate Cancer Cells by Regulating the miR-330-3p/BMI-1 Axis as a Competing Endogenous RNA. Front. Cell Dev. Biol. 2020, 8, 827. [Google Scholar] [CrossRef] [PubMed]
  44. Tang, Y.; Liu, J.; Li, X.; Wang, W. Exosomal circRNA HIPK3 knockdown inhibited cell proliferation and metastasis in prostate cancer by regulating miR-212/BMI-1 pathway. J. Biosci. 2021, 46, 69. [Google Scholar] [CrossRef]
  45. Khan, A.P.; Poisson, L.M.; Bhat, V.B.; Fermin, D.; Zhao, R.; Kalyana-Sundaram, S.; Michailidis, G.; Nesvizhskii, A.I.; Omenn, G.S.; Chinnaiyan, A.M.; et al. Quantitative proteomic profiling of prostate cancer reveals a role for miR-128 in prostate cancer. Mol. Cell Proteom. 2010, 9, 298–312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Jin, M.; Zhang, T.; Liu, C.; Badeaux, M.A.; Liu, B.; Liu, R.; Jeter, C.; Chen, X.; Vlassov, A.V.; Tang, D.G. miRNA-128 suppresses prostate cancer by inhibiting BMI-1 to inhibit tumor-initiating cells. Cancer Res. 2014, 74, 4183–4195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Feng, B.; Wang, R.; Chen, L.B. Review of miR-200b and cancer chemosensitivity. Biomed. Pharm. 2012, 66, 397–402. [Google Scholar] [CrossRef] [PubMed]
  48. Yu, J.; Lu, Y.; Cui, D.; Li, E.; Zhu, Y.; Zhao, Y.; Zhao, F.; Xia, S. miR-200b suppresses cell proliferation, migration and enhances chemosensitivity in prostate cancer by regulating Bmi-1. Oncol. Rep. 2014, 31, 910–918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. 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 2015, 80, 276–283. [Google Scholar] [CrossRef]
  50. Ambs, S.; Prueitt, R.L.; Yi, M.; Hudson, R.S.; Howe, T.M.; Petrocca, F.; Wallace, T.A.; Liu, C.G.; Volinia, S.; Calin, G.A.; et al. Genomic profiling of microRNA and messenger RNA reveals deregulated microRNA expression in prostate cancer. Cancer Res. 2008, 68, 6162–6170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Li, F.; Liang, A.; Lv, Y.; Liu, G.; Jiang, A.; Liu, P. MicroRNA-200c Inhibits Epithelial-Mesenchymal Transition by Targeting the BMI-1 Gene Through the Phospho-AKT Pathway in Endometrial Carcinoma Cells In Vitro. Med. Sci. Monit. 2017, 23, 5139–5149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Dong, P.; Kaneuchi, M.; Watari, H.; Hamada, J.; Sudo, S.; Ju, J.; Sakuragi, N. MicroRNA-194 inhibits epithelial to mesenchymal transition of endometrial cancer cells by targeting oncogene BMI-1. Mol. Cancer 2011, 10, 99. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, L.; Liu, J.L.; Yu, L.; Liu, X.X.; Wu, H.M.; Lei, F.Y.; Wu, S.; Wang, X. Downregulated miR-495 [Corrected] Inhibits the G1-S Phase Transition by Targeting Bmi-1 in Breast Cancer. Medicine 2015, 94, e718. [Google Scholar] [CrossRef]
  54. Zhu, Y.; Yu, F.; Jiao, Y.; Feng, J.; Tang, W.; Yao, H.; Gong, C.; Chen, J.; Su, F.; Zhang, Y.; et al. Reduced miR-128 in breast tumor-initiating cells induces chemotherapeutic resistance via Bmi-1 and ABCC5. Clin. Cancer Res. 2011, 17, 7105–7115. [Google Scholar] [CrossRef] [Green Version]
  55. Bhattacharya, R.; Nicoloso, M.; Arvizo, R.; Wang, E.; Cortez, A.; Rossi, S.; Calin, G.A.; Mukherjee, P. MiR-15a and MiR-16 control Bmi-1 expression in ovarian cancer. Cancer Res. 2009, 69, 9090–9095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zhang, X.L.; Sun, B.L.; Tian, S.X.; Li, L.; Zhao, Y.C.; Shi, P.P. MicroRNA-132 reverses cisplatin resistance and metastasis in ovarian cancer by the targeted regulation on Bmi-1. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 3635–3644. [Google Scholar] [CrossRef]
  57. Li, B.; Chen, H.; Wu, N.; Zhang, W.J.; Shang, L.X. Deregulation of miR-128 in ovarian cancer promotes cisplatin resistance. Int. J. Gynecol. Cancer 2014, 24, 1381–1388. [Google Scholar] [CrossRef]
  58. Godlewski, J.; Nowicki, M.O.; Bronisz, A.; Williams, S.; Otsuki, A.; Nuovo, G.; Raychaudhury, A.; Newton, H.B.; Chiocca, E.A.; Lawler, S. Targeting of the Bmi-1 oncogene/stem cell renewal factor by microRNA-128 inhibits glioma proliferation and self-renewal. Cancer Res. 2008, 68, 9125–9130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Zhang, X.; Wei, C.; Li, J.; Liu, J.; Qu, J. MicroRNA-361-5p inhibits epithelial-to-mesenchymal transition of glioma cells through targeting Twist1. Oncol. Rep. 2017, 37, 1849–1856. [Google Scholar] [CrossRef] [Green Version]
  60. Tu, Y.; Gao, X.; Li, G.; Fu, H.; Cui, D.; Liu, H.; Jin, W.; Zhang, Y. MicroRNA-218 inhibits glioma invasion, migration, proliferation, and cancer stem-like cell self-renewal by targeting the polycomb group gene Bmi1. Cancer Res. 2013, 73, 6046–6055. [Google Scholar] [CrossRef] [Green Version]
  61. Cui, J.G.; Zhao, Y.; Sethi, P.; Li, Y.Y.; Mahta, A.; Culicchia, F.; Lukiw, W.J. Micro-RNA-128 (miRNA-128) down-regulation in glioblastoma targets ARP5 (ANGPTL6), Bmi-1 and E2F-3a, key regulators of brain cell proliferation. J. Neurooncol. 2010, 98, 297–304. [Google Scholar] [CrossRef] [PubMed]
  62. Ye, L.; Yu, G.; Wang, C.; Du, B.; Sun, D.; Liu, J.; Qi, T.; Yu, X.; Wei, W.; Cheng, J.; et al. MicroRNA-128a, BMI1 polycomb ring finger oncogene, and reactive oxygen species inhibit the growth of U-87 MG glioblastoma cells following exposure to X-ray radiation. Mol. Med. Rep. 2015, 12, 6247–6254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Chen, F.; Chen, L.; He, H.; Huang, W.; Zhang, R.; Li, P.; Meng, Y.; Jiang, X. Up-regulation of microRNA-16 in Glioblastoma Inhibits the Function of Endothelial Cells and Tumor Angiogenesis by Targeting Bmi-1. Anticancer. Agents Med. Chem. 2016, 16, 609–620. [Google Scholar] [CrossRef] [PubMed]
  64. Venkataraman, S.; Alimova, I.; Fan, R.; Harris, P.; Foreman, N.; Vibhakar, R. MicroRNA 128a increases intracellular ROS level by targeting Bmi-1 and inhibits medulloblastoma cancer cell growth by promoting senescence. PLoS ONE 2010, 5, e10748. [Google Scholar] [CrossRef]
  65. Lai, G.H.; Huang, A.L.; Zhao, Z.; Lu, X.H.; Zu, W.X. [MicroRNA-218 promotes osteosarcoma cell apoptosis by down-regulating oncogene B lymphoma mouse Moloney leukemia virus insertion region 1]. Nan Fang Yi Ke Da Xue Xue Bao 2018, 38, 505–510. [Google Scholar] [CrossRef] [PubMed]
  66. Peng, H.; Liang, D.; Li, B.; Liang, C.; Huang, W.; Lin, H. MicroRNA-320a protects against osteoarthritis cartilage degeneration by regulating the expressions of BMI-1 and RUNX2 in chondrocytes. Pharmazie 2017, 72, 223–226. [Google Scholar] [CrossRef]
  67. Liu, L.; Qiu, M.; Tan, G.; Liang, Z.; Qin, Y.; Chen, L.; Chen, H.; Liu, J. miR-200c inhibits invasion, migration and proliferation of bladder cancer cells through down-regulation of BMI-1 and E2F3. J. Transl. Med. 2014, 12, 305. [Google Scholar] [CrossRef] [Green Version]
  68. Cheng, Y.; Yang, X.; Deng, X.; Zhang, X.; Li, P.; Tao, J.; Lu, Q. MicroRNA-218 inhibits bladder cancer cell proliferation, migration, and invasion by targeting BMI-1. Tumour Biol. 2015, 36, 8015–8023. [Google Scholar] [CrossRef]
  69. Zhang, L.; Wang, C.Z.; Ma, M.; Shao, G.F. MiR-15 suppressed the progression of bladder cancer by targeting BMI1 oncogene via PI3K/AKT signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 8813–8822. [Google Scholar] [CrossRef]
  70. Liu, J.F.; He, P.; Pan, D.F. [MiR-218 Targeting Bmi-1 Inhibits Proliferation of Acute Promyelocytic Leukemia Cells]. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2020, 28, 815–820. [Google Scholar] [CrossRef]
  71. Xu, L.; Sun, H.B.; Xu, Z.N.; Han, X.L.; Yin, Y.Y.; Zheng, Y.; Zhao, Y.; Wang, Z.X. MicroRNA-218 regulates the epithelial-to-mesenchymal transition and the PI3K/Akt signaling pathway to suppress lung adenocarcinoma progression by directly targeting BMI-1. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 7978–7988. [Google Scholar] [CrossRef]
  72. Wu, S.Q.; Niu, W.Y.; Li, Y.P.; Huang, H.B.; Zhan, R. miR-203 inhibits cell growth and regulates G1/S transition by targeting Bmi-1 in myeloma cells. Mol. Med. Rep. 2016, 14, 4795–4801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Qiu, M.; Liang, Z.; Chen, L.; Tan, G.; Liu, L.; Wang, K.; Chen, H.; Liu, J. MicroRNA-200c suppresses cell growth and metastasis by targeting Bmi-1 and E2F3 in renal cancer cells. Exp. Ther. Med. 2017, 13, 1329–1336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Liu, S.; Yang, Y.; Chen, L.; Liu, D.; Dong, H. MicroRNA-154 functions as a tumor suppressor in non-small cell lung cancer through directly targeting B-cell-specific Moloney murine leukemia virus insertion site 1. Oncol. Lett. 2018, 15, 10098–10104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Liu, S.; Tetzlaff, M.T.; Cui, R.; Xu, X. miR-200c inhibits melanoma progression and drug resistance through down-regulation of BMI-1. Am. J. Pathol. 2012, 181, 1823–1835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Bassiouni, M.; Dos Santos, A.; Avci, H.X.; Löwenheim, H.; Müller, M. Bmi1 Loss in the Organ of Corti Results in p16ink4a Upregulation and Reduced Cell Proliferation of Otic Progenitors In Vitro. PLoS ONE 2016, 11, e0164579. [Google Scholar] [CrossRef] [Green Version]
  77. Cao, R.; Tsukada, Y.; Zhang, Y. Role of Bmi-1 and Ring1A in H2A ubiquitylation and Hox gene silencing. Mol. Cell 2005, 20, 845–854. [Google Scholar] [CrossRef]
  78. Gao, F.; Huang, W.; Zhang, Y.; Tang, S.; Zheng, L.; Ma, F.; Wang, Y.; Tang, H.; Li, X. Hes1 promotes cell proliferation and migration by activating Bmi-1 and PTEN/Akt/GSK3β pathway in human colon cancer. Oncotarget 2015, 6, 38667–38680. [Google Scholar] [CrossRef] [Green Version]
  79. Dimri, G.P.; Martinez, J.L.; Jacobs, J.J.; Keblusek, P.; Itahana, K.; Van Lohuizen, M.; Campisi, J.; Wazer, D.E.; Band, V. The Bmi-1 oncogene induces telomerase activity and immortalizes human mammary epithelial cells. Cancer Res. 2002, 62, 4736–4745. [Google Scholar]
  80. Yamashita, M.; Kuwahara, M.; Suzuki, A.; Hirahara, K.; Shinnaksu, R.; Hosokawa, H.; Hasegawa, A.; Motohashi, S.; Iwama, A.; Nakayama, T. Bmi1 regulates memory CD4 T cell survival via repression of the Noxa gene. J. Exp. Med. 2008, 205, 1109–1120. [Google Scholar] [CrossRef]
  81. Kimura, M.; Takenobu, H.; Akita, N.; Nakazawa, A.; Ochiai, H.; Shimozato, O.; Fujimura, Y.; Koseki, H.; Yoshino, I.; Kimura, H.; et al. Bmi1 regulates cell fate via tumor suppressor WWOX repression in small-cell lung cancer cells. Cancer Sci. 2011, 102, 983–990. [Google Scholar] [CrossRef]
  82. Kameyama, A.; Nishijima, R.; Yamakoshi, K. Bmi-1 regulates mucin levels and mucin O-glycosylation in the submandibular gland of mice. PLoS ONE 2021, 16, e0245607. [Google Scholar] [CrossRef]
  83. Yu, F.; Zhou, C.; Zeng, H.; Liu, Y.; Li, S. BMI1 activates WNT signaling in colon cancer by negatively regulating the WNT antagonist IDAX. Biochem. Biophys. Res. Commun. 2018, 496, 468–474. [Google Scholar] [CrossRef]
  84. Siddique, H.R.; Saleem, M. Role of BMI1, a stem cell factor, in cancer recurrence and chemoresistance: Preclinical and clinical evidences. Stem Cells 2012, 30, 372–378. [Google Scholar] [CrossRef] [PubMed]
  85. Park, I.K.; Morrison, S.J.; Clarke, M.F. Bmi1, stem cells, and senescence regulation. J. Clin. Investig. 2004, 113, 175–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Ren, H.; Du, P.; Ge, Z.; Jin, Y.; Ding, D.; Liu, X.; Zou, Q. TWIST1 and BMI1 in Cancer Metastasis and Chemoresistance. J. Cancer 2016, 7, 1074–1080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Ma, D.Q.; Zhang, Y.H.; Ding, D.P.; Li, J.; Chen, L.L.; Tian, Y.Y.; Ao, K.J. Effect of Bmi-1-mediated NF-κB signaling pathway on the stem-like properties of CD133+ human liver cancer cells. Cancer Biomark. 2018, 22, 575–585. [Google Scholar] [CrossRef]
  88. Zheng, X.; Wang, Y.; Liu, B.; Liu, C.; Liu, D.; Zhu, J.; Yang, C.; Yan, J.; Liao, X.; Meng, X.; et al. Bmi-1-shRNA inhibits the proliferation of lung adenocarcinoma cells by blocking the G1/S phase through decreasing cyclin D1 and increasing p21/p27 levels. Nucleic Acid Ther. 2014, 24, 210–216. [Google Scholar] [CrossRef] [Green Version]
  89. Liu, Y.L.; Jiang, S.X.; Yang, Y.M.; Xu, H.; Liu, J.L.; Wang, X.S. USP22 acts as an oncogene by the activation of BMI-1-mediated INK4a/ARF pathway and Akt pathway. Cell Biochem. Biophys. 2012, 62, 229–235. [Google Scholar] [CrossRef]
  90. Jiao, K.; Jiang, W.; Zhao, C.; Su, D.; Zhang, H. Bmi-1 in gallbladder carcinoma: Clinicopathology and mechanism of regulation of human gallbladder carcinoma proliferation. Oncol. Lett. 2019, 18, 1365–1371. [Google Scholar] [CrossRef] [Green Version]
  91. Chen, F.; Li, Y.; Wang, L.; Hu, L. Knockdown of BMI-1 causes cell-cycle arrest and derepresses p16INK4a, HOXA9 and HOXC13 mRNA expression in HeLa cells. Med. Oncol. 2011, 28, 1201–1209. [Google Scholar] [CrossRef]
  92. Wang, J.F.; Liu, Y.; Liu, W.J.; He, S.Y. Expression of Bmi-1 gene in esophageal carcinoma cell EC9706 and its effect on cell cycle, apoptosis and migration. Chin. J. Cancer 2010, 29, 689–696. [Google Scholar] [CrossRef]
  93. Hu, C.; Zhang, Q.; Tang, Q.; Zhou, H.; Liu, W.; Huang, J.; Liu, Y.; Wang, Q.; Zhang, J.; Zhou, M.; et al. CBX4 promotes the proliferation and metastasis via regulating BMI-1 in lung cancer. J. Cell Mol. Med. 2020, 24, 618–631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kim, B.R.; Kwon, Y.; Rho, S.B. BMI-1 interacts with sMEK1 and inactivates sMEK1-induced apoptotic cell death. Oncol. Rep. 2017, 37, 579–586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Xiong, D.; Ye, Y.; Fu, Y.; Wang, J.; Kuang, B.; Wang, H.; Wang, X.; Zu, L.; Xiao, G.; Hao, M.; et al. Bmi-1 expression modulates non-small cell lung cancer progression. Cancer Biol. Ther. 2015, 16, 756–763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ginjala, V.; Nacerddine, K.; Kulkarni, A.; Oza, J.; Hill, S.J.; Yao, M.; Citterio, E.; van Lohuizen, M.; Ganesan, S. BMI1 is recruited to DNA breaks and contributes to DNA damage-induced H2A ubiquitination and repair. Mol. Cell Biol. 2011, 31, 1972–1982. [Google Scholar] [CrossRef] [Green Version]
  97. Wu, K.; Woo, S.M.; Seo, S.U.; Kwon, T.K. Inhibition of BMI-1 Induces Apoptosis through Downregulation of DUB3-Mediated Mcl-1 Stabilization. Int. J. Mol. Sci. 2021, 22, 10107. [Google Scholar] [CrossRef]
  98. Li, J.; Gong, L.Y.; Song, L.B.; Jiang, L.L.; Liu, L.P.; Wu, J.; Yuan, J.; Cai, J.C.; He, M.; Wang, L.; et al. Oncoprotein Bmi-1 renders apoptotic resistance to glioma cells through activation of the IKK-nuclear factor-kappaB Pathway. Am. J. Pathol. 2010, 176, 699–709. [Google Scholar] [CrossRef] [Green Version]
  99. Liu, J.; Luo, B.; Zhao, M. Bmi-1-targeting suppresses osteosarcoma aggressiveness through the NF-κB signaling pathway. Mol Med. Rep. 2017, 16, 7949–7958. [Google Scholar] [CrossRef] [Green Version]
  100. Levy, J.M.M.; Towers, C.G.; Thorburn, A. Targeting autophagy in cancer. Nat. Rev. Cancer 2017, 17, 528–542. [Google Scholar] [CrossRef]
  101. Dey, A.; Mustafi, S.B.; Saha, S.; Kumar Dhar Dwivedi, S.; Mukherjee, P.; Bhattacharya, R. Inhibition of BMI1 induces autophagy-mediated necroptosis. Autophagy 2016, 12, 659–670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Chen, S.; Li, H.; Chen, S.; Wang, B.; Zhang, K. BMI1 promotes the proliferation and inhibits autophagy of breast cancer cells by activating COPZ1. Clin. Transl. Oncol. 2022; Online ahead of print. [Google Scholar] [CrossRef]
  103. Griffith, J.; Andrade, D.; Mehta, M.; Berry, W.; Benbrook, D.M.; Aravindan, N.; Herman, T.S.; Ramesh, R.; Munshi, A. Silencing BMI1 radiosensitizes human breast cancer cells by inducing DNA damage and autophagy. Oncol. Rep. 2017, 37, 2382–2390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yang, W.; Wu, Z.; Yang, K.; Han, Y.; Chen, Y.; Zhao, W.; Huang, F.; Jin, Y.; Jin, W. BMI1 promotes cardiac fibrosis in ischemia-induced heart failure via the PTEN-PI3K/Akt-mTOR signaling pathway. Am. J. Physiol. Heart Circ. Physiol. 2019, 316, H61–H69. [Google Scholar] [CrossRef] [Green Version]
  105. Mourgues, L.; Imbert, V.; Nebout, M.; Colosetti, P.; Neffati, Z.; Lagadec, P.; Verhoeyen, E.; Peng, C.; Duprez, E.; Legros, L.; et al. The BMI1 polycomb protein represses cyclin G2-induced autophagy to support proliferation in chronic myeloid leukemia cells. Leukemia 2015, 29, 1993–2002. [Google Scholar] [CrossRef]
  106. Yu, J.; Chen, L.; Bao, Z.; Liu, Y.; Liu, G.; Li, F.; Li, L. BMI-1 promotes invasion and metastasis in endometrial adenocarcinoma and is a poor prognostic factor. Oncol. Rep. 2020, 43, 1630–1640. [Google Scholar] [CrossRef] [PubMed]
  107. Li, Y.; Tian, Z.; Tan, Y.; Lian, G.; Chen, S.; Chen, S.; Li, J.; Li, X.; Huang, K.; Chen, Y. Bmi-1-induced miR-27a and miR-155 promote tumor metastasis and chemoresistance by targeting RKIP in gastric cancer. Mol. Cancer 2020, 19, 109. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, X.; Tian, T.; Sun, W.; Liu, C.; Fang, X. Bmi-1 overexpression as an efficient prognostic marker in patients with nonsmall cell lung cancer. Medicine 2017, 96, e7346. [Google Scholar] [CrossRef]
  109. Pourjafar, M.; Samadi, P.; Karami, M.; Najafi, R. Assessment of clinicopathological and prognostic relevance of BMI-1 in patients with colorectal cancer: A meta-analysis. Biotechnol. Appl. Biochem. 2021, 68, 1313–1322. [Google Scholar] [CrossRef]
  110. Li, X.; Yang, Z.; Song, W.; Zhou, L.; Li, Q.; Tao, K.; Zhou, J.; Wang, X.; Zheng, Z.; You, N.; et al. Overexpression of Bmi-1 contributes to the invasion and metastasis of hepatocellular carcinoma by increasing the expression of matrix metalloproteinase (MMP)-2, MMP-9 and vascular endothelial growth factor via the PTEN/PI3K/Akt pathway. Int. J. Oncol. 2013, 43, 793–802. [Google Scholar] [CrossRef] [Green Version]
  111. Jiang, L.; Wu, J.; Yang, Y.; Liu, L.; Song, L.; Li, J.; Li, M. Bmi-1 promotes the aggressiveness of glioma via activating the NF-kappaB/MMP-9 signaling pathway. BMC Cancer 2012, 12, 406. [Google Scholar] [CrossRef] [Green Version]
  112. Ye, K.; Chen, Q.W.; Sun, Y.F.; Lin, J.A.; Xu, J.H. Loss of BMI-1 dampens migration and EMT of colorectal cancer in inflammatory microenvironment through TLR4/MD-2/MyD88-mediated NF-κB signaling. J. Cell Biochem. 2018, 119, 1922–1930. [Google Scholar] [CrossRef] [PubMed]
  113. Chagraoui, J.; Hébert, J.; Girard, S.; Sauvageau, G. An anticlastogenic function for the Polycomb Group gene Bmi1. Proc. Natl. Acad. Sci. USA 2011, 108, 5284–5289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Gieni, R.S.; Ismail, I.H.; Campbell, S.; Hendzel, M.J. Polycomb group proteins in the DNA damage response: A link between radiation resistance and “stemness”. Cell Cycle 2011, 10, 883–894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Liu, J.; Cao, L.; Chen, J.; Song, S.; Lee, I.H.; Quijano, C.; Liu, H.; Keyvanfar, K.; Chen, H.; Cao, L.Y.; et al. Bmi1 regulates mitochondrial function and the DNA damage response pathway. Nature 2009, 459, 387–392. [Google Scholar] [CrossRef]
  116. Dong, Q.; Oh, J.E.; Chen, W.; Kim, R.; Kim, R.H.; Shin, K.H.; McBride, W.H.; Park, N.H.; Kang, M.K. Radioprotective effects of Bmi-1 involve epigenetic silencing of oxidase genes and enhanced DNA repair in normal human keratinocytes. J. Investig. Dermatol. 2011, 131, 1216–1225. [Google Scholar] [CrossRef] [Green Version]
  117. Lin, X.; Ojo, D.; Wei, F.; Wong, N.; Gu, Y.; Tang, D. A Novel Aspect of Tumorigenesis-BMI1 Functions in Regulating DNA Damage Response. Biomolecules 2015, 5, 3396–3415. [Google Scholar] [CrossRef] [Green Version]
  118. Ismail, I.H.; Andrin, C.; McDonald, D.; Hendzel, M.J. BMI1-mediated histone ubiquitylation promotes DNA double-strand break repair. J. Cell Biol. 2010, 191, 45–60. [Google Scholar] [CrossRef]
  119. Wang, E.; Bhattacharyya, S.; Szabolcs, A.; Rodriguez-Aguayo, C.; Jennings, N.B.; Lopez-Berestein, G.; Mukherjee, P.; Sood, A.K.; Bhattacharya, R. Enhancing chemotherapy response with Bmi-1 silencing in ovarian cancer. PLoS ONE 2011, 6, e17918. [Google Scholar] [CrossRef] [Green Version]
  120. Singh, N.; Baby, D.; Rajguru, J.P.; Patil, P.B.; Thakkannavar, S.S.; Pujari, V.B. Inflammation and cancer. Ann. Afr. Med. 2019, 18, 121–126. [Google Scholar] [CrossRef]
  121. Vlachostergios, P.J.; Papandreou, C.N. The Bmi-1/NF-κB/VEGF story: Another hint for proteasome involvement in glioma angiogenesis? J. Cell Commun. Signal. 2013, 7, 235–237. [Google Scholar] [CrossRef] [Green Version]
  122. Sun, P.; Mu, Y.; Zhang, S. A novel NF-κB/MMP-3 signal pathway involves in the aggressivity of glioma promoted by Bmi-1. Tumour. Biol. 2014, 35, 12721–12727. [Google Scholar] [CrossRef]
  123. Jiang, L.; Song, L.; Wu, J.; Yang, Y.; Zhu, X.; Hu, B.; Cheng, S.Y.; Li, M. Bmi-1 promotes glioma angiogenesis by activating NF-κB signaling. PLoS ONE 2013, 8, e55527. [Google Scholar] [CrossRef] [Green Version]
  124. Qin, Z.Y.; Gu, X.; Chen, Y.L.; Liu, J.B.; Hou, C.X.; Lin, S.Y.; Hao, N.N.; Liang, Y.; Chen, W.; Meng, H.Y. Toll-like receptor 4 activates the NLRP3 inflammasome pathway and periodontal inflammaging by inhibiting Bmi-1 expression. Int. J. Mol. Med. 2021, 47, 137–150. [Google Scholar] [CrossRef]
  125. Spillane, J.B.; Henderson, M.A. Cancer stem cells: A review. ANZ J. Surg. 2007, 77, 464–468. [Google Scholar] [CrossRef]
  126. Raaphorst, F.M. Self-renewal of hematopoietic and leukemic stem cells: A central role for the Polycomb-group gene Bmi-1. Trends Immunol. 2003, 24, 522–524. [Google Scholar] [CrossRef]
  127. Bruggeman, S.W.; Valk-Lingbeek, M.E.; van der Stoop, P.P.; Jacobs, J.J.; Kieboom, K.; Tanger, E.; Hulsman, D.; Leung, C.; Arsenijevic, Y.; Marino, S.; et al. Ink4a and Arf differentially affect cell proliferation and neural stem cell self-renewal in Bmi1-deficient mice. Genes Dev. 2005, 19, 1438–1443. [Google Scholar] [CrossRef] [Green Version]
  128. Zencak, D.; Lingbeek, M.; Kostic, C.; Tekaya, M.; Tanger, E.; Hornfeld, D.; Jaquet, M.; Munier, F.L.; Schorderet, D.F.; van Lohuizen, M.; et al. Bmi1 loss produces an increase in astroglial cells and a decrease in neural stem cell population and proliferation. J. Neurosci. Off. J. Soc. Neurosci. 2005, 25, 5774–5783. [Google Scholar] [CrossRef]
  129. Gargiulo, G.; Cesaroni, M.; Serresi, M.; de Vries, N.; Hulsman, D.; Bruggeman, S.W.; Lancini, C.; van Lohuizen, M. In Vivo RNAi Screen for BMI1 Targets Identifies TGF-beta/BMP-ER Stress Pathways as Key Regulators of Neural- and Malignant Glioma-Stem Cell Homeostasis. Cancer Cell 2013, 23, 660–676. [Google Scholar] [CrossRef] [Green Version]
  130. Molofsky, A.V.; He, S.; Bydon, M.; Morrison, S.J.; Pardal, R. Bmi-1 promotes neural stem cell self-renewal and neural development but not mouse growth and survival by repressing the p16Ink4a and p19Arf senescence pathways. Genes Dev. 2005, 19, 1432–1437. [Google Scholar] [CrossRef] [Green Version]
  131. Zelentsova-Levytskyi, K.; Talmi, Z.; Abboud-Jarrous, G.; Capucha, T.; Sapir, T.; Burstyn-Cohen, T. Protein S Negatively Regulates Neural Stem Cell Self-Renewal through Bmi-1 Signaling. Front. Mol. Neurosci. 2017, 10, 124. [Google Scholar] [CrossRef]
  132. Lessard, J.; Sauvageau, G. Bmi-1 determines the proliferative capacity of normal and leukaemic stem cells. Nature 2003, 423, 255–260. [Google Scholar] [CrossRef] [PubMed]
  133. Cho, J.H.; Dimri, M.; Dimri, G.P. A Positive Feedback Loop Regulates the Expression of Polycomb Group Protein BMI1 via WNT Signaling Pathway. J. Biol. Chem. 2013, 288, 3406–3418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Lukacs, R.U.; Memarzadeh, S.; Wu, H.; Witte, O.N. Bmi-1 Is a Crucial Regulator of Prostate Stem Cell Self-Renewal and Malignant Transformation. Cell Stem Cell 2010, 7, 682–693. [Google Scholar] [CrossRef] [Green Version]
  135. Yan, K.S.; Chia, L.A.; Li, X.N.; Ootani, A.; Su, J.; Lee, J.Y.; Su, N.; Luo, Y.L.; Heilshorn, S.C.; Amieva, M.R.; et al. The intestinal stem cell markers Bmi1 and Lgr5 identify two functionally distinct populations. Proc. Natl. Acad. Sci. USA 2012, 109, 466–471. [Google Scholar] [CrossRef] [Green Version]
  136. Qiao, C.; Xu, W.R.; Zhu, W.; Hu, J.B.; Qian, H.; Yin, Q.; Jiang, R.Q.; Yan, Y.M.; Mao, F.; Yang, H.; et al. Human mesenchymal stem cells isolated from the umbilical cord. Cell Biol. Int. 2008, 32, 8–15. [Google Scholar] [CrossRef] [PubMed]
  137. Wu, T.; Dai, Y. Tumor microenvironment and therapeutic response. Cancer Lett. 2017, 387, 61–68. [Google Scholar] [CrossRef]
  138. Arneth, B. Tumor Microenvironment. Medicina 2019, 56, 15. [Google Scholar] [CrossRef] [Green Version]
  139. Bolomsky, A.; Schlangen, K.; Schreiner, W.; Zojer, N.; Ludwig, H. Targeting of BMI-1 with PTC-209 shows potent anti-myeloma activity and impairs the tumour microenvironment. J. Hematol. Oncol. 2016, 9, 17. [Google Scholar] [CrossRef] [Green Version]
  140. Zhang, D.; Huang, J.; Wang, F.; Ding, H.; Cui, Y.; Yang, Y.; Xu, J.; Luo, H.; Gao, Y.; Pan, L.; et al. BMI1 regulates multiple myeloma-associated macrophage’s pro-myeloma functions. Cell Death Dis. 2021, 12, 495. [Google Scholar] [CrossRef]
  141. Sugihara, H.; Ishimoto, T.; Watanabe, M.; Sawayama, H.; Iwatsuki, M.; Baba, Y.; Komohara, Y.; Takeya, M.; Baba, H. Identification of miR-30e* regulation of Bmi1 expression mediated by tumor-associated macrophages in gastrointestinal cancer. PLoS ONE 2013, 8, e81839. [Google Scholar] [CrossRef]
  142. Zhao, J.; Luo, X.D.; Da, C.L.; Xin, Y. Clinicopathological significance of B-cell-specific Moloney murine leukemia virus insertion site 1 expression in gastric carcinoma and its precancerous lesion. World J. Gastroenterol. 2009, 15, 2145–2150. [Google Scholar] [CrossRef]
  143. Lee, H.; Yoon, S.O.; Jeong, W.Y.; Kim, H.K.; Kim, A.; Kim, B.H. Immunohistochemical analysis of polycomb group protein expression in advanced gastric cancer. Hum. Pathol. 2012, 43, 1704–1710. [Google Scholar] [CrossRef]
  144. Lu, Y.W.; Li, J.; Guo, W.J. Expression and clinicopathological significance of Mel-18 and Bmi-1 mRNA in gastric carcinoma. J. Exp. Clin. Cancer Res. 2010, 29, 143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Horie, K.; Iseki, C.; Kikuchi, M.; Miyakawa, K.; Yoshizaki, M.; Yoshioka, H.; Watanabe, J. Bmi-1 Immunohistochemical Expression in Endometrial Carcinoma is Correlated with Prognostic Activity. Medicina 2020, 56, 72. [Google Scholar] [CrossRef] [Green Version]
  146. Liu, J.; Liu, K.; Jiang, X.; Wang, X.; Chen, Y.; Cui, X.; Pang, L.; Li, S.; Liu, C.; Zou, H.; et al. Clinicopathological significance of Bmi-1 overexpression in esophageal cancer: A meta-analysis. Biomark. Med. 2018, 12, 71–81. [Google Scholar] [CrossRef]
  147. Tong, Y.Q.; Liu, B.; Zheng, H.Y.; He, Y.J.; Gu, J.; Li, F.; Li, Y. Overexpression of BMI-1 is associated with poor prognosis in cervical cancer. Asia Pac. J. Clin. Oncol. 2012, 8, e55–e62. [Google Scholar] [CrossRef]
  148. Swelem, R.S.; Elneely, D.A.; Shehata, A.A.R. The Study of SALL4 Gene and BMI-1 Gene Expression in Acute Myeloid Leukemia Patients. Lab. Med. 2020, 51, 265–270. [Google Scholar] [CrossRef]
  149. Zhang, Y.; Zhang, Y.L.; Chen, H.M.; Pu, H.W.; Ma, W.J.; Li, X.M.; Ma, H.; Chen, X. Expression of Bmi-1 and PAI-1 in esophageal squamous cell carcinoma. World J. Gastroenterol. 2014, 20, 5533–5539. [Google Scholar] [CrossRef] [PubMed]
  150. Liu, W.L.; Guo, X.Z.; Zhang, L.J.; Wang, J.Y.; Zhang, G.; Guan, S.; Chen, Y.M.; Kong, Q.L.; Xu, L.H.; Li, M.Z.; et al. Prognostic relevance of Bmi-1 expression and autoantibodies in esophageal squamous cell carcinoma. BMC Cancer 2010, 10, 467. [Google Scholar] [CrossRef] [Green Version]
  151. Li, D.W.; Tang, H.M.; Fan, J.W.; Yan, D.W.; Zhou, C.Z.; Li, S.X.; Wang, X.L.; Peng, Z.H. Expression level of Bmi-1 oncoprotein is associated with progression and prognosis in colon cancer. J. Cancer Res. Clin. Oncol. 2010, 136, 997–1006. [Google Scholar] [CrossRef]
  152. Zhang, X.; Wang, C.X.; Zhu, C.B.; Zhang, J.; Kan, S.F.; Du, L.T.; Li, W.; Wang, L.L.; Wang, S. Overexpression of Bmi-1 in uterine cervical cancer: Correlation with clinicopathology and prognosis. Int. J. Gynecol. Cancer 2010, 20, 1597–1603. [Google Scholar] [PubMed]
  153. Qin, Z.K.; Yang, J.A.; Ye, Y.L.; Zhang, X.; Xu, L.H.; Zhou, F.J.; Han, H.; Liu, Z.W.; Song, L.B.; Zeng, M.S. Expression of Bmi-1 is a prognostic marker in bladder cancer. BMC Cancer 2009, 9, 61. [Google Scholar] [CrossRef] [Green Version]
  154. Yang, G.F.; He, W.P.; Cai, M.Y.; He, L.R.; Luo, J.H.; Deng, H.X.; Guan, X.Y.; Zeng, M.S.; Zeng, Y.X.; Xie, D. Intensive expression of Bmi-1 is a new independent predictor of poor outcome in patients with ovarian carcinoma. BMC Cancer 2010, 10, 133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Abd El Hafez, A.; El-Hadaad, H.A. Immunohistochemical expression and prognostic relevance of Bmi-1, a stem cell factor, in epithelial ovarian cancer. Ann. Diagn. Pathol. 2014, 18, 58–62. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, X.; Wang, C.; Wang, L.; Du, L.; Wang, S.; Zheng, G.; Li, W.; Zhuang, X.; Zhang, X.; Dong, Z. Detection of circulating Bmi-1 mRNA in plasma and its potential diagnostic and prognostic value for uterine cervical cancer. Int. J. Cancer 2012, 131, 165–172. [Google Scholar] [CrossRef]
  157. Yi, C.; Li, B.B.; Zhou, C.X. Bmi-1 expression predicts prognosis in salivary adenoid cystic carcinoma and correlates with epithelial-mesenchymal transition-related factors. Ann. Diagn. Pathol. 2016, 22, 38–44. [Google Scholar] [CrossRef]
  158. Allegra, E.; Puzzo, L.; Zuccalà, V.; Trapasso, S.; Vasquez, E.; Garozzo, A.; Caltabiano, R. Nuclear BMI-1 expression in laryngeal carcinoma correlates with lymph node pathological status. World J. Surg. Oncol. 2012, 10, 206. [Google Scholar] [CrossRef] [Green Version]
  159. Song, W.; Tao, K.; Li, H.; Jin, C.; Song, Z.; Li, J.; Shi, H.; Li, X.; Dang, Z.; Dou, K. Bmi-1 is related to proliferation, survival and poor prognosis in pancreatic cancer. Cancer Sci. 2010, 101, 1754–1760. [Google Scholar] [CrossRef]
  160. Häyry, V.; Mäkinen, L.K.; Atula, T.; Sariola, H.; Mäkitie, A.; Leivo, I.; Keski-Säntti, H.; Lundin, J.; Haglund, C.; Hagström, J. Bmi-1 expression predicts prognosis in squamous cell carcinoma of the tongue. Br. J. Cancer 2010, 102, 892–897. [Google Scholar] [CrossRef]
  161. Lozneanu, L.; Balan, R.A.; Păvăleanu, I.; Giuşcă, S.E.; Căruntu, I.D.; Amalinei, C. BMI-1 Expression Heterogeneity in Endometriosis-Related and Non-Endometriotic Ovarian Carcinoma. Int. J. Mol. Sci 2021, 22, 6082. [Google Scholar] [CrossRef]
  162. Cui, H.; Hu, B.; Li, T.; Ma, J.; Alam, G.; Gunning, W.T.; Ding, H.F. Bmi-1 is essential for the tumorigenicity of neuroblastoma cells. Am. J. Pathol. 2007, 170, 1370–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Farivar, S.; Zati Keikha, R.; Shiari, R.; Jadali, F. Expression of bmi-1 in pediatric brain tumors as a new independent prognostic marker of patient survival. Biomed. Res. Int. 2013, 2013, 192548. [Google Scholar] [CrossRef] [Green Version]
  164. Huber, G.F.; Albinger-Hegyi, A.; Soltermann, A.; Roessle, M.; Graf, N.; Haerle, S.K.; Holzmann, D.; Moch, H.; Hegyi, I. Expression patterns of Bmi-1 and p16 significantly correlate with overall, disease-specific, and recurrence-free survival in oropharyngeal squamous cell carcinoma. Cancer 2011, 117, 4659–4670. [Google Scholar] [CrossRef]
  165. Mihic-Probst, D.; Kuster, A.; Kilgus, S.; Bode-Lesniewska, B.; Ingold-Heppner, B.; Leung, C.; Storz, M.; Seifert, B.; Marino, S.; Schraml, P.; et al. Consistent expression of the stem cell renewal factor BMI-1 in primary and metastatic melanoma. Int. J. Cancer 2007, 121, 1764–1770. [Google Scholar] [CrossRef]
  166. Yu, D.; Liu, Y.; Yang, J.; Jin, C.; Zhao, X.; Cheng, J.; Liu, X.; Qi, X. Clinical implications of BMI-1 in cancer stem cells of laryngeal carcinoma. Cell Biochem. Biophys. 2015, 71, 261–269. [Google Scholar] [CrossRef] [PubMed]
  167. Chen, Y.; Lian, G.; Ou, G.; Yang, K.; Chen, J.; Li, H.; Chen, S.; Li, J.; Zeng, L.; Huang, K. Inverse association between Bmi-1 and RKIP affecting clinical outcome of gastric cancer and revealing the potential molecular mechanisms underlying tumor metastasis and chemotherapy resistance. Gastric Cancer 2016, 19, 392–402. [Google Scholar] [CrossRef] [Green Version]
  168. Dong, S.C.; Jing, R.N.; Pei, H.; Liu, F.; Zhao, B.X.; Meng, X.X. [Effect of Bmi-1 Expression on Chemotherapy Sensitivity in THP-1 Cells]. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2021, 29, 363–368. [Google Scholar] [CrossRef]
  169. Herzog, A.E.; Warner, K.A.; Zhang, Z.; Bellile, E.; Bhagat, M.A.; Castilho, R.M.; Wolf, G.T.; Polverini, P.J.; Pearson, A.T.; Nör, J.E. The IL-6R and Bmi-1 axis controls self-renewal and chemoresistance of head and neck cancer stem cells. Cell Death Dis. 2021, 12, 988. [Google Scholar] [CrossRef]
  170. Jiang, M.; He, G.; Li, J.; Li, J.; Guo, X.; Gao, J. Hypoxic exposure activates the B cell-specific Moloney murine leukaemia virus integration site 1/PI3K/Akt axis and promotes EMT in leukaemia stem cells. Oncol. Lett. 2021, 21, 98. [Google Scholar] [CrossRef]
  171. Bhattacharyya, J.; Mihara, K.; Ohtsubo, M.; Yasunaga, S.; Takei, Y.; Yanagihara, K.; Sakai, A.; Hoshi, M.; Takihara, Y.; Kimura, A. Overexpression of BMI-1 correlates with drug resistance in B-cell lymphoma cells through the stabilization of survivin expression. Cancer Sci. 2012, 103, 34–41. [Google Scholar] [CrossRef]
  172. Wang, G.; Liu, L.; Sharma, S.; Liu, H.; Yang, W.; Sun, X.; Dong, Q. Bmi-1 confers adaptive radioresistance to KYSE-150R esophageal carcinoma cells. Biochem. Biophys. Res. Commun. 2012, 425, 309–314. [Google Scholar] [CrossRef] [PubMed]
  173. Jia, L.; Zhang, W.; Wang, C.Y. BMI1 Inhibition Eliminates Residual Cancer Stem Cells after PD1 Blockade and Activates Antitumor Immunity to Prevent Metastasis and Relapse. Cell Stem Cell 2020, 27, 238–253.e236. [Google Scholar] [CrossRef] [PubMed]
  174. Kreso, A.; van Galen, P.; Pedley, N.M.; Lima-Fernandes, E.; Frelin, C.; Davis, T.; Cao, L.; Baiazitov, R.; Du, W.; Sydorenko, N.; et al. Self-renewal as a therapeutic target in human colorectal cancer. Nat. Med. 2014, 20, 29–36. [Google Scholar] [CrossRef]
  175. Wang, Q.; Li, Z.; Wu, Y.; Huang, R.; Zhu, Y.; Zhang, W.; Wang, Y.; Cheng, J. Pharmacological inhibition of Bmi1 by PTC-209 impaired tumor growth in head neck squamous cell carcinoma. Cancer Cell Int. 2017, 17, 107. [Google Scholar] [CrossRef] [Green Version]
  176. Yoo, Y.A.; Vatapalli, R.; Lysy, B.; Mok, H.; Desouki, M.M.; Abdulkadir, S.A. The Role of Castration-Resistant Bmi1+Sox2+ Cells in Driving Recurrence in Prostate Cancer. J. Natl. Cancer Inst. 2019, 111, 311–321. [Google Scholar] [CrossRef]
  177. Alzrigat, M.; Párraga, A.A.; Majumder, M.M.; Ma, A.; Jin, J.; Österborg, A.; Nahi, H.; Nilsson, K.; Heckman, C.A.; Öberg, F.; et al. The polycomb group protein BMI-1 inhibitor PTC-209 is a potent anti-myeloma agent alone or in combination with epigenetic inhibitors targeting EZH2 and the BET bromodomains. Oncotarget 2017, 8, 103731–103743. [Google Scholar] [CrossRef] [Green Version]
  178. Kong, Y.; Ai, C.; Dong, F.; Xia, X.; Zhao, X.; Yang, C.; Kang, C.; Zhou, Y.; Zhao, Q.; Sun, X.; et al. Targeting of BMI-1 with PTC-209 inhibits glioblastoma development. Cell Cycle 2018, 17, 1199–1211. [Google Scholar] [CrossRef] [Green Version]
  179. Li, J.; Vangundy, Z.; Poi, M. PTC209, a Specific Inhibitor of BMI1, Promotes Cell Cycle Arrest and Apoptosis in Cervical Cancer Cell Lines. Anticancer Res. 2020, 40, 133–141. [Google Scholar] [CrossRef] [PubMed]
  180. Xu, J.; Zhang, Y.; Xu, J.; Wang, M.; Liu, G.; Wang, J.; Zhao, X.; Qi, Y.; Shi, J.; Cheng, K.; et al. Reversing tumor stemness via orally targeted nanoparticles achieves efficient colon cancer treatment. Biomaterials 2019, 216, 119247. [Google Scholar] [CrossRef]
  181. Nishida, Y.; Maeda, A.; Kim, M.J.; Cao, L.; Kubota, Y.; Ishizawa, J.; AlRawi, A.; Kato, Y.; Iwama, A.; Fujisawa, M.; et al. The novel BMI-1 inhibitor PTC596 downregulates MCL-1 and induces p53-independent mitochondrial apoptosis in acute myeloid leukemia progenitor cells. Blood Cancer J. 2017, 7, e527. [Google Scholar] [CrossRef]
  182. Maeda, A.; Nishida, Y.; Weetall, M.; Cao, L.; Branstrom, A.; Ishizawa, J.; Nii, T.; Schober, W.D.; Abe, Y.; Matsue, K.; et al. Targeting of BMI-1 expression by the novel small molecule PTC596 in mantle cell lymphoma. Oncotarget 2018, 9, 28547–28560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Bolomsky, A.; Muller, J.; Stangelberger, K.; Lejeune, M.; Duray, E.; Breid, H.; Vrancken, L.; Pfeiffer, C.; Hübl, W.; Willheim, M.; et al. The anti-mitotic agents PTC-028 and PTC596 display potent activity in pre-clinical models of multiple myeloma but challenge the role of BMI-1 as an essential tumour gene. Br. J. Haematol. 2020, 190, 877–890. [Google Scholar] [CrossRef] [PubMed]
  184. Flamier, A.; Abdouh, M.; Hamam, R.; Barabino, A.; Patel, N.; Gao, A.; Hanna, R.; Bernier, G. Off-target effect of the BMI1 inhibitor PTC596 drives epithelial-mesenchymal transition in glioblastoma multiforme. NPJ Precis. Oncol. 2020, 4, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Jernigan, F.; Branstrom, A.; Baird, J.D.; Cao, L.; Dali, M.; Furia, B.; Kim, M.J.; O’Keefe, K.; Kong, R.; Laskin, O.L.; et al. Preclinical and Early Clinical Development of PTC596, a Novel Small-Molecule Tubulin-Binding Agent. Mol. Cancer Ther. 2021, 20, 1846–1857. [Google Scholar] [CrossRef] [PubMed]
  186. Shapiro, G.I.; O’Mara, E.; Laskin, O.L.; Gao, L.; Baird, J.D.; Spiegel, R.J.; Kaushik, D.; Weetall, M.; Colacino, J.; O’Keefe, K.; et al. Pharmacokinetics and Safety of PTC596, a Novel Tubulin-Binding Agent, in Subjects With Advanced Solid Tumors. Clin. Pharmacol. Drug Dev. 2021, 10, 940–949. [Google Scholar] [CrossRef]
  187. Bartucci, M.; Hussein, M.S.; Huselid, E.; Flaherty, K.; Patrizii, M.; Laddha, S.V.; Kui, C.; Bigos, R.A.; Gilleran, J.A.; El Ansary, M.M.S.; et al. Synthesis and Characterization of Novel BMI1 Inhibitors Targeting Cellular Self-Renewal in Hepatocellular Carcinoma. Target. Oncol. 2017, 12, 449–462. [Google Scholar] [CrossRef] [PubMed]
  188. Dey, A.; Xiong, X.; Crim, A.; Dwivedi, S.K.D.; Mustafi, S.B.; Mukherjee, P.; Cao, L.; Sydorenko, N.; Baiazitov, R.; Moon, Y.C.; et al. Evaluating the Mechanism and Therapeutic Potential of PTC-028, a Novel Inhibitor of BMI-1 Function in Ovarian Cancer. Mol. Cancer Ther. 2018, 17, 39–49. [Google Scholar] [CrossRef] [Green Version]
  189. Wang, J.; Xing, Y.; Wang, Y.; He, Y.; Wang, L.; Peng, S.; Yang, L.; Xie, J.; Li, X.; Qiu, W.; et al. A novel BMI-1 inhibitor QW24 for the treatment of stem-like colorectal cancer. J. Exp. Clin. Cancer Res. 2019, 38, 422. [Google Scholar] [CrossRef]
  190. Shields, C.E.; Potlapalli, S.; Cuya-Smith, S.M.; Chappell, S.K.; Chen, D.; Martinez, D.; Pogoriler, J.; Rathi, K.S.; Patel, S.A.; Oristian, K.M.; et al. Epigenetic regulator BMI1 promotes alveolar rhabdomyosarcoma proliferation and constitutes a novel therapeutic target. Mol. Oncol. 2021, 15, 2156–2171. [Google Scholar] [CrossRef]
  191. Sulaiman, S.; Arafat, K.; Iratni, R.; Attoub, S. PTC-209 Anti-Cancer Effects Involved the Inhibition of STAT3 Phosphorylation. Front. Pharmacol. 2019, 10, 1199. [Google Scholar] [CrossRef]
  192. Shan, W.; Zhou, L.; Liu, L.; Lin, D.; Yu, Q. Polycomb group protein Bmi1 is required for the neuronal differentiation of mouse induced pluripotent stem cells. Exp. Ther. Med. 2021, 21, 619. [Google Scholar] [CrossRef] [PubMed]
  193. Wang, X.; Li, K.; Cheng, M.; Wang, G.; Han, H.; Chen, F.; Liao, W.; Chen, Z.; Chen, J.; Bao, Y.; et al. Bmi1 Severs as a Potential Tumor-Initiating Cell Marker and Therapeutic Target in Esophageal Squamous Cell Carcinoma. Stem Cells Int. 2020, 2020, 8877577. [Google Scholar] [CrossRef] [PubMed]
  194. Barbosa, K.; Deshpande, A.; Chen, B.R.; Ghosh, A.; Sun, Y.; Dutta, S.; Weetall, M.; Dixon, J.; Armstrong, S.A.; Bohlander, S.K.; et al. Acute myeloid leukemia driven by the CALM-AF10 fusion gene is dependent on BMI1. Exp. Hematol. 2019, 74, 42–51.e43. [Google Scholar] [CrossRef] [PubMed]
  195. Ohtaka, M.; Itoh, M.; Tohda, S. BMI1 Inhibitors Down-regulate NOTCH Signaling and Suppress Proliferation of Acute Leukemia Cells. Anticancer Res. 2017, 37, 6047–6053. [Google Scholar] [CrossRef] [PubMed]
  196. Nishida, Y.; Maeda, A.; Chachad, D.; Ishizawa, J.; Qiu, Y.H.; Kornblau, S.M.; Kimura, S.; Andreeff, M.; Kojima, K. Preclinical activity of the novel B-cell-specific Moloney murine leukemia virus integration site 1 inhibitor PTC-209 in acute myeloid leukemia: Implications for leukemia therapy. Cancer Sci. 2015, 106, 1705–1713. [Google Scholar] [CrossRef] [PubMed]
  197. Dimri, M.; Kang, M.; Dimri, G.P. A miR-200c/141-BMI1 autoregulatory loop regulates oncogenic activity of BMI1 in cancer cells. Oncotarget 2016, 7, 36220–36234. [Google Scholar] [CrossRef] [Green Version]
  198. Mayr, C.; Wagner, A.; Loeffelberger, M.; Bruckner, D.; Jakab, M.; Berr, F.; Di Fazio, P.; Ocker, M.; Neureiter, D.; Pichler, M.; et al. The BMI1 inhibitor PTC-209 is a potential compound to halt cellular growth in biliary tract cancer cells. Oncotarget 2016, 7, 745–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Nagai, Y.; Mimura, N.; Rizq, O.; Isshiki, Y.; Oshima, M.; Rizk, M.; Saraya, A.; Koide, S.; Nakajima-Takagi, Y.; Miyota, M.; et al. The combination of the tubulin binding small molecule PTC596 and proteasome inhibitors suppresses the growth of myeloma cells. Sci Rep. 2021, 11, 2074. [Google Scholar] [CrossRef]
  200. Eberle-Singh, J.A.; Sagalovskiy, I.; Maurer, H.C.; Sastra, S.A.; Olive, K.P. Effective Delivery of a Microtubule Polymerization Inhibitor Synergizes with Standard Regimens in Models of Pancreatic Ductal Adenocarcinoma. Clin. Cancer Res. 2019, 25, 5548–5560. [Google Scholar] [CrossRef]
  201. Kumar, S.S.; Sengupta, S.; Zhu, X.; Mishra, D.K.; Phoenix, T.; Dyer, L.; Fuller, C.; Stevenson, C.B.; Dewire, M.; Fouladi, M. Diffuse Intrinsic Pontine Glioma Cells Are Vulnerable to Mitotic Abnormalities Associated with BMI-1 Modulation. Mol. Cancer Res. 2020, 18, 1711–1723. [Google Scholar] [CrossRef] [PubMed]
  202. Balakrishnan, I.; Danis, E.; Pierce, A.; Madhavan, K.; Wang, D.; Dahl, N.; Sanford, B.; Birks, D.K.; Davidson, N.; Metselaar, D.S.; et al. Senescence Induced by BMI1 Inhibition Is a Therapeutic Vulnerability in H3K27M-Mutant DIPG. Cell Rep. 2020, 33, 108286. [Google Scholar] [CrossRef]
  203. Bakhshinyan, D.; Venugopal, C.; Adile, A.A.; Garg, N.; Manoranjan, B.; Hallett, R.; Wang, X.; Mahendram, S.; Vora, P.; Vijayakumar, T.; et al. BMI1 is a therapeutic target in recurrent medulloblastoma. Oncogene 2019, 38, 1702–1716. [Google Scholar] [CrossRef] [PubMed]
  204. Buechel, M.; Dey, A.; Dwivedi, S.K.D.; Crim, A.; Ding, K.; Zhang, R.; Mukherjee, P.; Moore, K.N.; Cao, L.; Branstrom, A.; et al. Inhibition of BMI1, a Therapeutic Approach in Endometrial Cancer. Mol. Cancer Ther. 2018, 17, 2136–2143. [Google Scholar] [CrossRef] [Green Version]
  205. Yang, L.F.; Xing, Y.; Xiao, J.X.; Xie, J.; Gao, W.; Xie, J.; Wang, L.T.; Wang, J.; Liu, M.; Yi, Z.; et al. Synthesis of Cyanoenone-Modified Diterpenoid Analogs as Novel Bmi-1-Mediated Antitumor Agents. ACS Med. Chem. Lett. 2018, 9, 1105–1110. [Google Scholar] [CrossRef]
  206. Wu, J.; Hu, D.; Yang, G.; Zhou, J.; Yang, C.; Gao, Y.; Zhu, Z. Down-regulation of BMI-1 cooperates with artemisinin on growth inhibition of nasopharyngeal carcinoma cells. J. Cell Biochem. 2011, 112, 1938–1948. [Google Scholar] [CrossRef]
  207. Wang, Y.Y.; Yang, Y.X.; Zhao, R.; Pan, S.T.; Zhe, H.; He, Z.X.; Duan, W.; Zhang, X.; Yang, T.; Qiu, J.X.; et al. Bardoxolone methyl induces apoptosis and autophagy and inhibits epithelial-to-mesenchymal transition and stemness in esophageal squamous cancer cells. Drug Des. Dev. Ther. 2015, 9, 993–1026. [Google Scholar] [CrossRef] [Green Version]
  208. Zhao, Z.; Yang, Y.; Zeng, Y.; He, M. A microfluidic ExoSearch chip for multiplexed exosome detection towards blood-based ovarian cancer diagnosis. Lab Chip 2016, 16, 489–496. [Google Scholar] [CrossRef] [Green Version]
  209. Glia, A.; Deliorman, M.; Sukumar, P.; Janahi, F.K.; Samara, B.; Brimmo, A.T.; Qasaimeh, M.A. Herringbone Microfluidic Probe for Multiplexed Affinity-Capture of Prostate Circulating Tumor Cells. Adv. Mater. Technol. 2021, 6, 2100053. [Google Scholar] [CrossRef]
  210. Kure, K.; Hosoya, M.; Ueyama, T.; Fukaya, M.; Komiyama, H. Using the polymeric circulating tumor cell chip to capture circulating tumor cells in blood samples of patients with colorectal cancer. Oncol. Lett. 2020, 19, 2286–2294. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The structure of gene and protein of Bmi-1. The Bmi-1 gene contains 10 exons and 9 introns. The amino acid sequence of Bmi-1 protein contains a RING finger domain, a helix–turn–helix, two nuclear localization signals (NLS), and a PEST region.
Figure 1. The structure of gene and protein of Bmi-1. The Bmi-1 gene contains 10 exons and 9 introns. The amino acid sequence of Bmi-1 protein contains a RING finger domain, a helix–turn–helix, two nuclear localization signals (NLS), and a PEST region.
Ijms 23 08231 g001
Figure 2. Downstream genes directly regulated by Bmi-1. Bmi-1 can enhance H2A ubiquitination and bind to the promoter of HoxC13 to reduce the expression of HoxC13 in HeLa cells [77]. The most classic downstream target of Bmi-1 is INK4a/ARF. The tumor suppressor gene INK4a/ARF can encode two regulatory genes: p16INK4a and p14ARF (p19ARF in mice). P16INK4a arrests cells in G0/G1 phase through cyclinD-CDK4/6-pRb-E2F. p14ARF regulates the cell apoptosis through MDM2-p53 pathway [76]. Bmi-1 can regulate the development of colon cancer by negatively regulating the expression level of PTEN and then activating the Akt/GSK3β axis [78]. Bmi-1 regulates memory CD4 T cell survival via repression of the proapoptotic BH3-only protein Noxa gene [80]. Bmi-1 regulates cell fate via transcriptional repression tumor suppressor WW Domain Containing Oxidoreductase (WWOX) in small-cell lung cancer cells [81]. Bmi-1 suppresses the expression of Smgc gene and Gcnt3 gene [82]. Bmi-1 activates Wnt signaling in colon cancer by negatively regulating the Wnt antagonist IDAX [83]. Bmi-1 can also transcriptionally positively regulate human telomerase reverse transcriptase (hTERT) to block senescence in human mammary epithelial cells [79].
Figure 2. Downstream genes directly regulated by Bmi-1. Bmi-1 can enhance H2A ubiquitination and bind to the promoter of HoxC13 to reduce the expression of HoxC13 in HeLa cells [77]. The most classic downstream target of Bmi-1 is INK4a/ARF. The tumor suppressor gene INK4a/ARF can encode two regulatory genes: p16INK4a and p14ARF (p19ARF in mice). P16INK4a arrests cells in G0/G1 phase through cyclinD-CDK4/6-pRb-E2F. p14ARF regulates the cell apoptosis through MDM2-p53 pathway [76]. Bmi-1 can regulate the development of colon cancer by negatively regulating the expression level of PTEN and then activating the Akt/GSK3β axis [78]. Bmi-1 regulates memory CD4 T cell survival via repression of the proapoptotic BH3-only protein Noxa gene [80]. Bmi-1 regulates cell fate via transcriptional repression tumor suppressor WW Domain Containing Oxidoreductase (WWOX) in small-cell lung cancer cells [81]. Bmi-1 suppresses the expression of Smgc gene and Gcnt3 gene [82]. Bmi-1 activates Wnt signaling in colon cancer by negatively regulating the Wnt antagonist IDAX [83]. Bmi-1 can also transcriptionally positively regulate human telomerase reverse transcriptase (hTERT) to block senescence in human mammary epithelial cells [79].
Ijms 23 08231 g002
Table 1. MicroRNAs that inhibit Bmi-1 in tumors.
Table 1. MicroRNAs that inhibit Bmi-1 in tumors.
miRNAsEfficacy for CancerCancer TypeReferences
miR-34a, miR-15a, miR-218, miR-183, miR-498, miR-128Inhibits proliferation; inhibits metastasis; decreases chemoresistanceGastric cancer[23,24,25,26,27]
miR-218, miR-200c, miR-485-5pInhibits proliferation; inhibits migration; increases apoptosisColorectal cancer[28,29,30]
miR-15a, miR-183Inhibits proliferation and EMTPancreatic ductal adenocarcinoma[31,32]
miR-203Inhibits self-renewalEsophageal cancer[33]
miR-203Promotes apoptosisOral cancer[34]
miR-218, miR-203Inhibits proliferation and invasion; introduces apoptosis; decreases radiosensitivity and chemosensitivityHepatocellular carcinoma[35,36,37,38,39]
miR-218Increases chemosensitivityLiver cancer[38]
miR-320aInhibits proliferation and migrationNosopharyngeal carcinoma[40]
miR-132, miR-498Increases radiosensitivity; inhibits proliferation and invasionCervical cancer[41,42]
miR-128, miR-200b, miR-221, miR-30d, miR-15a, miR-330-3p, miR-212Inhibits proliferation and migration; increases chemosensitivityProstate cancer[43,44,45,46,47,48,49,50]
miR-200c, miR-194Inhibits proliferation; inhibits EMTEndometrial carcinoma[51,52]
miR-128, miR-495Inhibits proliferation; introduces apoptosis; increases chemosensitivity and DNA damageBreast cancer[53,54]
miR-132, miR-15a, miR-16, miR-128Inhibits metastasis; increases chemosensitivityOvarian cancer[55,56,57]
miR-361-5p, miR-218, miR-128Inhibits proliferation; inhibits EMT Glioma[58,59,60]
miR-128, miR-16, miR-128aInhibits proliferation and angiogenesis; introduces radiosensitivityGlioblastoma[61,62,63]
miR-128aInhibits ROSMedulloblastoma[64]
miR-218Inhibits proliferation, inhibits migration, inhibits apoptosisOsteosarcoma[65,66]
miR-200c, miR-139-5p, miR-218, miR-15Inhibits proliferation; inhibits metastasis; inhibits apoptosis; inhibits autophagy Bladder cancer[13,67,68,69]
miR-218Inhibits proliferationAcute Promyelocytic Leukemia [70]
miR-218Inhibits proliferation and metastasis Lung adenocarcinoma[71]
miR-203Inhibits proliferationMyeloma [72]
miR-200CInhibits proliferation and metastasisRenal cancer[73]
miR-154Inhibits proliferation and migration Non-small cell lung cancer[74]
miR-200cInhibits proliferation and migration; increases chemosensitivityMelanoma[75]
Table 2. Protein expression and clinical characteristics of Bmi-1.
Table 2. Protein expression and clinical characteristics of Bmi-1.
Cancer TypemRNA/ProteinHigh/Low ExpressionPositive PercentageClinical CharacteristicsRemarksRef.
Gastric cancerProteinHighGC162 (52.5%)Associated with Lauren’s and Borrmann’s classification and clinical stageMainly in nucleus[142]
ProteinHighGC 178 (70.8%)Associated with sex, gross type, and histologic type Mainly in nucleus[143]
mRNAHigh71Associated with tumor size, depth of invasion, lymph node metastasis, and clinical stageNot involved[144]
Nonsmall cell lung cancerProtein and mRNAHigh Associated with tumor size, poor differentiation, and distant metastasisMainly in nucleus[108]
Endometrial CarcinomaProteinHigh48A significant positive relationship between Bmi-1 and Ki-67, cyclin A, or p53Mainly in nucleus[145]
Esophageal cancerProteinHigh1523Associated with differentiation, tumor/node/metastasis stage, depth of invasion, and lymph node metastasis Mainly in nucleus[146]
Cervical cancerProtein and mRNAHigh302 (55.3%)Correlated with clinical stage, lymph node metastasis, vascular invasion, and human papillomavirus (HPV) infectionMainly in nucleus[147]
Acute myeloid leukemia (AML)mRNAHigh60Showed a strong association with failure to achieve complete remission (CR) or with relapse Not involved[148]
Esophageal squamous cell carcinoma (ESCC)ProteinHigh80 (78.7%)Correlated with depth of invasion and lymph node metastasis, but not with patient age, tumor size, or nationality Not involved[149]
ESCCProtein and mRNAHigh171 (64.3%)Correlated with stage and pN classification.Mainly in nucleus[150]
Endometrial adenocarcinomaProteinHigh60 Correlated with FIGO stage, myometrial invasion, and lymph node metastasis both the nucleus and cytoplasm[106]
Colon cancerProtein and mRNAHigh203 (66.5%)Correlated with clinical stage, depth of invasion, nodal involvement, distant metastasis, and Ki67 levelMainly in nucleus[151]
Uterine cervical cancerProteinHigh152 Correlated with tumor size, clinical stage, and regional lymph nodes metastasisMainly in nucleus[152]
Bladder cancerProtein High137 Correlated withhistopathological classification, clinical stage, recurrence, and patient survivalMainly in nucleus[153]
Ovarian carcinomaProteinLow179Correlated with tumors’ histological type, grade, pT/pN/pM status, and FIGO stageboth the nucleus and cytoplasm[154]
Epithelial ovarian cancerProteinHigh40 (72.5%)Associated with advanced International Federation of Gynecology and Obstetrics stages, bilaterality, and higher Gynecologic Oncology Group grades and carcinomas of serous histologyMainly in nucleus[155]
Uterine cervical cancermRNAHigh109Correlated with clinical stage and lymph nodes metastasis Not involved[156]
Salivary adenoid cystic carcinomaProteinHigh10Associated with tumor metastasis, Snail, Slug, and E-cadherin, serves as a highrisk for AdCCNot involved[157]
Laryngeal carcinomaProteinHigh64 (84.4%)Correlated with distant metastasis, N pathological status, T classificationB oth the nucleus and cytoplasm[158]
Pancreatic cancerProteinHigh72 (48.61%)Correlated with the presence of lymph node metastases and negatively correlated with patient survival ratesMainly in nucleus[159]
Squamous cell carcinoma of the tongueProteinHigh73 (82%)Correlated with recurrence in nucleus[160]
ovarian CarcinomamRNAHigh47(72.34%)Correlated with tumor grade both the nucleus and cytoplasm[161]
NeuroblastomaProteinHigh45Correlated with MYCNin nucleus[162]
Pediatric brain tumorsmRNAHigh56Expression of Bmi-1 showed significant differences between high-grade tumors and low-grade tumorsNot involved[163]
Table 3. Inhibitors of Bmi-1.
Table 3. Inhibitors of Bmi-1.
Inhibitor NameRegulatory for Bmi-1Tumor TypeRegulatory MechanismReferences
PTC-209Reduces transcript levelsCervical cancerPromotes cell G0/G1 arrest and apoptosis[179]
Colon cancerDeveloped an orally active, easily synthesized PTC209 nanomedicine[180]
Alveolar rhabdomyosarcomaActivates the Hippo pathway[190]
GlioblastomaInhibits glioblastoma cell proliferation and migration[178]
Ovarian cancerInduces autophagy through ATP depletion[101]
Lung cancer cells, breast cancer cells and colon cancer cellsInhibits STAT3 Phosphorylation[191]
Pluripotent stem cellsReduces the expression of neuronal markers, such as Nestin[192]
Prostate cancerEfficiently targets Bmi-1 and Sox2[176]
ESCCInhibits ESCC progression when combined with cisplatin[193]
Acute myeloid leukemiaInhibits proliferation and induce apoptosis[194]
Acute Leukemia CellsDown-regulates the expression of Notch signaling proteins Notch1, Hes1, and MYC[195]
Acute myeloid leukemiaReduces protein level of Bmi-1 and its downstream target mono-ubiquitinated histone H2A and induces apoptosis[196]
Breast cancerTranscriptionally upregulates expression of miR-200c/141 cluster[197]
Biliary tract cancer cellsCauses down-regulation of cell cycle-promoting genes, DNA synthesis gene and DNA repair gene[198]
Chronic myeloid leukemia cellsTriggers CCNG2 expression[105]
MMDown-regulates the expression of Bmi-1 protein and the associated repressive histone mark H2AK119ub[177]
HNSCCInhibits proliferation, migration and invasiveness, increases cell apoptosis and chemosensitivity[175]
PTC596Reduces protein levels of BMI-1MyelomaInduces cell cycle arrest at G2/M phase followed by apoptotic cell death[199]
AML Downregulates Mcl-1 and induces p53-independent mitochondrial apoptosis[181]
GlioblastomaTargets both Bmi-1 and EZH2, prevents GBM colony growth and CSC self-renewal[184]
Mantle cell lymphomaInduces mitochondrial apoptosis, loss of mitochondrial membrane potential, C-caspase-3, Bax activation, and phosphatidylserine externalization[182]
Cancer stem-like cellsInduces apoptosis through DUB3-mediated Mcl-1 degradation[99]
Pancreatic ductal adenocarcinoma (PDA)Induces mitotic arrest and apoptosis[200]
Diffuse intrinsic pontine glioma (DIPG)Decreases tumor volume and growth kinetics, increases intertumoral apoptosis, and sustains animal survival benefit.[201]
RU-A1Bind to the Bmi-1 mRNAHepatocellular carcinomaImpairs cell viability, reduces cell migration, enhances HCC cell sensitivity to 5-fluorouracil (5-FU) in vitro[187]
PTC-028Posttranslational modificationmultiple myelomaImpairs MYC and Akt signalling activity; induces cell cycle arrest at G2/M phase and apoptotic [183]
Diffuse intrinsic pontine glioma (DIPG)Decreases the expression of E2F1, KRAS, Nestin, SOX2 while increases the expression of p21 and differentiation markers (GFAP)[202]
HyperphosphorylationOvarian cancerDecreases ATP and a compromised mitochondrial redox balance potentiate caspase-dependent apoptosis[188]
Medulloblastoma (MB)Abolishes the self-renewal capacity of MB stem cells, reduces tumor initiation ability of recurrent MB cells[203]
Endometrial cancerReduces cell invasive capacity and enhances caspase-dependent apoptosis[204]
Alveolar rhabdomyosarcomaInhibits proliferation and causes tumor growth delay in vivo[190]
QW24Autophagy-lysosome degradation pathwayStem-like colorectal cancerInhibits self-renewal of colorectal cancer-initiating cells (CICs)[189]
SH498 Colorectal cancerReduces PRC1 complex activity by down-regulating Bmi-1 and ub-H2A[205]
ArtemisininProtein and transcript levelsNasopharyngeal carcinoma Induces G1 cell cycle arrest via the Bmi-1-p16/CDK4 axis[206]
PRT4165Down-regulating Bmi-1/RING1A self-ubiquitinationAcute leukemiaIncreases cell apoptosis[195]
CDDO-Me ESCCInduces autophagy via suppression of PI3K/Akt/mTOR signaling pathway[207]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, J.; Li, L.; Shi, P.; Cui, H.; Yang, L. The Crucial Roles of Bmi-1 in Cancer: Implications in Pathogenesis, Metastasis, Drug Resistance, and Targeted Therapies. Int. J. Mol. Sci. 2022, 23, 8231. https://doi.org/10.3390/ijms23158231

AMA Style

Xu J, Li L, Shi P, Cui H, Yang L. The Crucial Roles of Bmi-1 in Cancer: Implications in Pathogenesis, Metastasis, Drug Resistance, and Targeted Therapies. International Journal of Molecular Sciences. 2022; 23(15):8231. https://doi.org/10.3390/ijms23158231

Chicago/Turabian Style

Xu, Jie, Lin Li, Pengfei Shi, Hongjuan Cui, and Liqun Yang. 2022. "The Crucial Roles of Bmi-1 in Cancer: Implications in Pathogenesis, Metastasis, Drug Resistance, and Targeted Therapies" International Journal of Molecular Sciences 23, no. 15: 8231. https://doi.org/10.3390/ijms23158231

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