Next Article in Journal
Repurposing Riociguat to Target a Novel Paracrine Nitric Oxide-TRPC6 Pathway to Prevent Podocyte Injury
Next Article in Special Issue
Molecular Mechanisms of lncRNAs in the Dependent Regulation of Cancer and Their Potential Therapeutic Use
Previous Article in Journal
High-Throughput and Accurate Determination of Transgene Copy Number and Zygosity in Transgenic Maize: From DNA Extraction to Data Analysis
Previous Article in Special Issue
Evidence That Regulation of Pri-miRNA/miRNA Expression Is Not a General Rule of miPEPs Function in Humans
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Roles of the Colon Cancer Associated Transcript 2 (CCAT2) Long Non-Coding RNA in Cancer: A Comprehensive Characterization of the Tumorigenic and Molecular Functions

1
Research Center for Functional Genomics Biomedicine and Translational Medicine, “Iuliu Hatieganu” University of Medicine and Pharmacy, 400012 Cluj-Napoca, Romania
2
Department of Translational Molecular Pathology, Division of Pathology, The University of Texas MD Anderson Cancer Center, Houston, TX 77030, USA
*
Author to whom correspondence should be addressed.
Authors with equal contribution.
Int. J. Mol. Sci. 2021, 22(22), 12491; https://doi.org/10.3390/ijms222212491
Submission received: 29 October 2021 / Revised: 17 November 2021 / Accepted: 17 November 2021 / Published: 19 November 2021

Abstract

:
Colon cancer-associated transcript 2 (CCAT2) is an intensively studied lncRNA with important regulatory roles in cancer. As such, cumulative studies indicate that CCAT2 displays a high functional versatility due to its direct interaction with multiple RNA binding proteins, transcription factors, and other species of non-coding RNA, especially microRNA. The definitory mechanisms of CCAT2 are its role as a regulator of the TCF7L2 transcription factor, enhancer of MYC expression, and activator of the WNT/β-catenin pathway, as well as a role in promoting and maintaining chromosome instability through the BOP1–AURKB pathway. Additionally, we highlight how the encompassing rs6983267 SNP has been shown to confer CCAT2 with allele-specific functional and structural particularities, such as the allelic-specific reprogramming of glutamine metabolism. Additionally, we emphasize CCAT2’s role as a competitive endogenous RNA (ceRNA) for multiple tumor suppressor miRNAs, such as miR-4496, miR-493, miR-424, miR-216b, miR-23b, miR-34a, miR-145, miR-200b, and miR-143 and the pro-tumorigenic role of the altered regulatory axis. Additionally, due to its upregulation in tumor tissues, wide distribution across cancer types, and presence in serum samples, we outline CCAT2’s potential as a biomarker and disease indicator and its implications for the development of resistance against current cancer therapy regiments and metastasis.

1. Introduction

While the existence of non-coding species of RNA has been acknowledged for several decades already, the advent of modern sequencing techniques has allowed their detailed characterization and classification. As such, ncRNAs have been mainly categorized based on size, structure, and specific cellular function [1]. Smaller species of ncRNA are mainly represented by microRNA (miRNA), which are considered to range between 19 and 25 nucleotides and are mainly involved in regulating gene expression post-transcriptionally, targeting specific mRNAs based on the presence of complementary seed sequences [2,3,4,5]. On the other hand, long non-coding RNAs are a more recent class of non-coding RNAs, longer than 200 nucleotides, that are involved in a wide array of biological processes, ranging from early phase physiological developmental to intricate disease-associated and molecular cancer mechanisms [6,7]. Unlike miRNAs, lncRNAs are subjected to polyadenylation and splicing, and they mostly display preferential nuclear localization and highly cell-specific expression patterns. As such, lncRNAs have been widely investigated in different developmental [8,9] and pathological processes [10,11], which indicated their high functional variability. As a relevant example, the lncRNA HOTAIR, initially highlighted as a trans epigenetic regulator [12] of the key developmental HOX genes through its direct interaction with the Polycomb Repressive Complex 2 (PRC2), was proven several years later to be a bona fide tumor-promoting lncRNA through the alteration of its regulatory mechanism in the case of metastatic breast cancer. In addition, the lncRNA NEAT2 (also known as MALAT-1) was initially indicated as a prognostic indicator for early-stage non-small-cell lung cancer [13]. Later investigations confirmed its regulatory involvement in mRNA metabolism based on the association with the SC35 [14] and serine/arginine (SR) splicing factors [15].
Colon cancer-associated transcripts 1 and 2 (CCAT1 and CCAT2) are two of the more recently identified transcripts with reoccurring implications in different types of cancers. First described in colon cancer, CCAT1 and CCAT2 both originate from the 8q24.21 chromosomal region, in close proximity with the MYC gene (500 kb and 300 kb upstream, respectively) and encompassing the cancer risk-associated rs6983267 single nucleotide polymorphism (SNP) (Figure 1) [16,17,18].
CCAT1 was the first one to be identified of the two, highlighted as a potential pre-malignant indicator for colorectal cancer (CRC), being identified as overexpressed in pre-tumoral tissues (adenomatous polyps and tumor-proximal colonic epithelium) but also in later stage tumor tissue [18]. Further investigations indicated that the CCAT1 gene encodes two isoforms for the lncRNA, a long one (CCAT1-L), with preferential nuclear localization, and a short (CCAT1-S) isoform, mainly localized in the cytoplasm, which is presumed to be a processing result of the long variant [17,18,19]. Functionally, CCAT1 is a transcriptional regulator for MYC and acts as a long-distance enhancer element via its interaction with the transcription factor and chromatin organizer CTCF, thus promoting and stabilizing chromatin looping formation at the MYC locus [20]. CCAT1 is elevated in various types of cancers, including gastrointestinal cancers, hepatocellular carcinoma, lung cancer, melanoma, and multiple myeloma, where it was correlated with the activation of essential cancer biologic processes including proliferation, invasion, metastasis, and treatment resistance [18,19,21]. Other identified cancer-specific mechanisms involve the formation of a complex with the TP63 and SOX2 with transcription factors. The triplex can then bind both the CCAT1 promoter, determining its transcription, and the super-enhancer domains of epidermal growth factor receptor (EGFR), thus activating downstream MAPK signaling cascade with the effect of promoting small cell carcinoma tumorigenesis [22,23]. CCAT1 shows many functional similarities with its sister transcript, as we highlight further throughout this review.
Similarly, CCAT2 was first identified in colon cancer during investigations aiming to identify the functional elements associated with the rs6983267 SNP [16], a locus on the 8q24 “gene desert” often associated with increased predisposition to the development of colon, ovarian, and prostate cancers [24]. Referring specifically to the encompassing rs6983267 SNP, the genetic variants identified in multiple populations consist of the three GG, TT, and GT alleles, the latter being considered the wild type. The presence of a specific allele appears to have functional effects, which are discussed further in the latter chapters of this manuscript. Since then, CCAT2 has been investigated in multiple oncologic malignancies and is emerging as a relevant lncRNA. Therefore, the implications of CCAT2 in different cancer types, emphasizing all the influential cancer-specific mechanisms and highlighting its potential biomarker and therapeutic use, are within the scope of this review.

2. Cancer-Associated Regulatory Activity of CCAT2

2.1. MYC and WNT Pathway Activation

The first indication of CCAT2 being a tumor-associated transcript came from the study of its encompassed SNP, rs6983267. Ling et al. highlighted that the novel transcript, originating from the 8q24.21 chromosomal region, was significantly enriched in microsatellite-stable colon cancer samples. Initial in vitro and in vivo investigations confirmed that overexpressing CCAT2 increased both the proliferation and metastatic potential of colon cancer cells. Based on its genomic proximity to MYC, the group investigated whether CCAT2 had any regulatory effect on its expression [16]. Multiple validation studies confirmed a positive correlation between CCAT2 and MYC, both at the RNA and protein level, also indicative of a similar, although unconfirmed, enhancer activity similar to CCAT1 and a previous indication of chromatin looping associated with the rs6983267 locus [25]. Further investigations of the regulatory mechanisms indicated that the TCF7L2 transcription factor interacts directly with CCAT2, an event confirmed by immunoprecipitation and supported by its nuclear co-localization. Previous studies highlighted the regulatory activity of TCF7L2 on the expression of MYC via the presence of multiple TCF7L2 binding sites on the MYC promoter. Moreover, TCF7L2 is a known activator of the pro-tumorigenic Wnt/β-catenin, one of the ubiquitous pathways constitutively overactivated and considered a critical driver of CRC. On the other hand, MYC is one of the downstream targets of the Wnt/β-catenin pathway, highlighting the dual potentiating role of CCAT2 in this tumor-promoting feedback loop. Moreover, downstream of the Wnt/β-catenin pathway, the Wnt-induced-secreted-protein-1 (WISP1) was also significantly upregulated in CCAT2 active cancers. In the downstream pathway, high WISP1 levels stimulate the expression of its effectors VEGF-A, MMP2, and ICAM-1, with important roles in angiogenesis, epithelial-to-mesenchymal transition (EMT) and metastasis [26]. This interplay between CCAT2, MYC, and WNT molecules has been explored and confirmed in multiple cancers in recent years and is responsible for the main pro-tumorigenic roles of CCAT2. Similar studies confirmed the presence of the mechanism in the case of thyroid [27], ovarian [28], clear cell renal cell carcinoma [29], and breast cancers [30].

2.2. Allele-Specific Metabolic Reprogramming

As previously mentioned, the allelic variant defined by the rs6983267 SNP has functional effects on the secondary structure and the enrichment levels of CCAT2. The first functional differences were first reported by Ling et al. in 2013 [16]. Following the characterization of the allelic distribution of rs6983267 in the studied cohorts, the group wanted to investigate whether the rs6983267 status affected CCAT2 expression in patient samples. While there were no noticeable differences in patient samples, cell lines with heterogeneous rs6983267 genotype expressed the GG variant at a significantly higher level. Furthermore, the group indicated that the variation of G and T alleles influenced the secondary structure of the lncRNA [16].
A more intricate characterization of the functional consequences of the allele distribution came several years later when Redis et al. indicated that CCAT2 promotes an allelic-specific metabolic reprogramming of glutamine metabolism in CRC [31]. Specifically, the interaction between the G-allele CCAT2 and two subunits (CFIm25 and CFIm68) of the cleavage factor I (CFIm) promoted the alternative splicing of the glutaminase (GLS) enzyme [31]. The upregulation of the glutaminase enzyme acted as a metabolic switch resulting in an acceleration in glutamine metabolism, thus promoting proliferative and metastatic programs, confirmed both in vitro and in vivo. Additionally, the CCAT2-G-CFIm-GLS regulatory axis was identified in more than half (61%) of the investigated clinical samples, highlighting the potential clinical utility of this allelic-dependent process.

2.3. Chromosomal Instability

The first indication of CCAT2s as a promoter of chromosomal instability was highlighted in 2013 when Ling. et al. reported that CRC cell lines that overexpressed CCAT2 displayed a higher percentage of abnormal metaphases and aberrant centrosomes numbers. Additionally, the group reported a positive correlation score between CIN score and CCAT2 expression levels in 218 breast cancer patients with lymph node-negative disease [16]. Further details on the related mechanism were reported several years later in the case of microsatellite stable CRC, with the extending implications of the development and sustainability of cancer-associated chromosomal instability in the therapeutic resistance to 5-fluorouracil and oxaliplatin. Investigations confirmed that the mechanism is dependent on CCAT2’s interaction with the BOP1 ribosomal biogenesis factor, which were both upregulated as a result of CCAT2’s upregulation of MYC and stabilized following a direct interaction. Upregulation of BOP1 increased the active form of Aurora kinase B, causing chromosomal mis-segregation errors. Inhibiting this regulatory axis, both CCAT2 or BOP1 knockdown decreased the invasive phenotype of the cells both in vitro and in vivo. Additionally, higher expression of CCAT2 and BOP1 correlated with shorter survival times in CRC patients [32].

2.4. RNA Editing and CCAT2-Associated RNA–DNA Differences at the rs6983267 Locus

The concept of RNA editing consists of a conserved process of post-transcriptional modifications in the RNA transcripts that result in varying degrees of sequence modifications when compared with the originating DNA sequence [33]. Both coding and non-coding RNAs are subjected to different types of RNA editing, resulting in increased transcriptional diversity and various functional effects, dependent on the type of RNA modified and the nature of the modification. The general pattern (also known as canonical editing) of modifications consists of adenine to inosine (A to I) and cytosine to uracil (C to U) and is dictated by adenosine (ADAR) and cytidine (APOBEC) deaminases, respectively. Adenine to inosine is the most commonly occurring ADAR adenosine deaminase, mainly represented by ADAR1 and ADARB1, catalyzing the RNA editing predominantly in RNA duplexes formed in either the UTR regions of mRNAs, introns, and repetitive sequences, such as Alu repeats [34].
CCAT2 is subjected to a particular non-ADAR, non-APOBEC type of editing, specifically at the rs6983267 locus. Shah et al. first indicated the existence of a DNA-to-RNA allelic imbalance after observing discrepancies between the genomic DNA (gDNA) and the reverse-transcribed CCAT2 RNA sequences (cDNA) originating from the same myelodysplastic syndrome patients [35]. After the exclusion of any sequencing or technical errors, the group confirmed that the event, named RE (RNA Editing), was occurring specifically at the rs6983267 locus and consisted in the variation between the G and T alleles of CCAT2 not falling into the previously described canonical RNA editing patterns. Additionally, other non-physiological allele variations, such as C or A, were not identified. Further investigation indicated that the editing event was disease-specific, being identified at a much higher rate in the blood marrow and peripheral blood of myelodysplastic syndrome (MDS) and myeloproliferative syndrome (MPS) patients when compared with healthy controls. Further in vivo investigations confirmed the similar occurrence rate of RE in CCAT2 transgenic models, which was also correlated with increased splenomegaly and blood marrow hypercellularity in individuals in which RE was identified (RE+). As for the in vivo functional implications of rs6983267-RE, bone marrow-derived cells from RE+ mice were found to have significant alterations in immune-related pathways, especially genes involved in B-cell receptor, IL22, and IL4 signaling. Further immune profiling of RE+ hematopoietic stem and progenitor cells highlighted D22, CD19, and H2-Ob as the main dysregulated genes in both cell types. All things considered, while the exact RNA editing mechanism to which CCAT2 is subjected has not yet been uncovered, there is significant evidence pointing towards its role in the modulation of immune signaling in MDS and MPS [35].

2.5. ceRNA Activity of CCAT2 and miRNA Sponging

A commonly described role of lncRNAs in different pathologies is their complementarity-based interaction with different species of RNA, especially miRNAs [36,37], thus creating complex regulatory networks that involve both the interacting miRNAs and their mRNA targets [38]. Also called “miRNA sponging”, the competitive endogenous RNA mechanism (ceRNA) implies the competitive binding via a complementary sequence of a single or multiple miRNAs to the lncRNA rather than its mRNA target, thus mitigating the regulatory function of that miRNA [39]. The miRNA binding sites are named MRE (miRNA response elements) and consist of short sequences (under 10 bp) in the lncRNA transcript [40].
Recent studies have also highlighted CCAT2 as having multiple miRNA regulatory targets (Figure 2, Table 1), acting as one of the main pro-neoplastic drivers in various human cancers [39,41,42,43]. For example, CCAT2 promotes its neoplastic regulatory functions through extensive ceRNA networks involving multiple tumor suppressor miRNAs, including miR-424, miR-145, miR-23b-5p, and miR-143 [39,42,44,45,46].
In glial tumor cells, CCAT2’s role in angiogenesis and apoptosis was investigated through mechanistic studies that revealed that CCAT2 is encapsulated by tumor cells in exosomes, which are subsequently internalized by endothelial cells. Exosomes enriched in CCAT2 promoted the translation of proangiogenic genes (VEGF, TGF β, FGF, and KDR) and antiapoptotic gene BCL-2 by inhibiting the proapoptotic genes BAX and CASPASE-3. Following the CCAT2 activation, the endothelial cells enhance the angiogenesis pathway, which leads to downstream upregulation of VEGF, TGFβ, and FGF protein secretion. Moreover, Lang et al. showed that the inhibitory effect on apoptosis of endothelial cells is enhanced under hypoxic conditions. Endothelial cells pre-exposed to hypoxia and then transfected with CCAT2-enriched exosomes showed increased expression of BCL-2 and lower BAX and CASPASE-3 expression [71]. Additional studies revealed that part of the CCAT2 regulatory effect on the VEGF gene is modulated through direct induction of the PI3K/AKT signaling pathway and via sponging of the miR-424 [43]. The ceRNA regulatory network is based on both CCAT2 and the VEGFA mRNA sharing a miR-424 MRE [43]. This ceRNA relationship between CCAT2 and miR-424 was first described in epithelial ovarian cancer, where the tumorigenic effects of CCAT2 were found to be modulated via the inhibition of the tumor suppressor activity of the miR-424. Knockdown of CCAT2 reduced cellular proliferation, migration, and angiogenesis through upregulation of miR-424 and consequent inhibition of VEGF, confirming that CCAT2 is a crucial element of the processes via this regulatory axis [39].
A new regulatory axis between CCAT2/miR-424 and the protein checkpoint kinase 1 (CHK1) gene was described in glioma, with roles in the tumor chemoresistance [44]. Previous studies reported that miR-424 tumor suppressor activity relies on interfering with genes (including CHK1) that are essential to the G1/S cell cycle transition, thus limiting cellular proliferation and promoting apoptosis. In cervical cancer, miR-424 expression was inversely correlated with the CHK1 gene, and high Chk1 protein levels and reduced levels of miR-424 were associated with aggressive behavior and poor outcome [101].
EMT is responsible for tumor aggressive behavior and acquisition of the ability to develop distant metastasis. CCAT2 is an enhancer of EMT and cellular proliferation through downregulation of E-cadherin and upregulation of N-cadherin and vimentin [28,62]. Additionally, CCAT2 was found to modulate the EMT phenotype through cAMP-responsive element-binding protein 1 (CREB1) expression modulation [96]. CREB1 is an oncogene that was shown to promote tumor cell proliferation, EMT transition, and metastasis in gastric cancer, CRC, and prostate cancer [102,103]. CREB1 was shown to be a target of the tumor suppressor miR-493-5p. In prostate cancer cell lines, upregulation of miR-493-5p inhibited CREB1 expression and inhibited EMT via AKT/GSK-3β/Snail signaling [103]. Wang et al. investigated the regulatory loop between CCAT2/miR-493-5p/CREB1 in cervical cancer, revealing that CCAT2 sponges miR-493-5p leading to upregulation of CREB1, which promotes proliferation and EMT [96].
Bioinformatic sequence prediction analysis revealed the possible interaction between CCAT2 and miR-23b-5p based on a putative binding site in the 3′-untranslated region (3-UTR) of CCAT2. miR-23b-5p was described as acting as a tumor suppressor in a variety of cancers, most importantly by inhibiting the FOXC1 gene, a known oncogene with important roles in signal transduction and tumor progression [104,105]. The direct interaction between CCAT2 and miR-23b-5p was evidenced using the dual-luciferase reporter assay in non-small-cell lung cancer cell lines (NSCLC). The expression level of CCAT2 was further investigated in relation to the expression of miR-23b-5p target gene FOXC1. A direct correlation between CCAT2 expression level FOXC1 gene expression and an inverse correlation between CCAT2 and miR-23b-5p was described. These data support that CCAT2 coordinates a loop that enhances the expression of the FOXC1 gene by direct sponging of the tumor suppressor miR-23b-5p. Moreover, the increased levels of FOXC1 further promote cellular proliferation and migration [42].
In hepatocellular carcinoma (HCC) cell models, CCAT2 was reported to promote cellular migration, proliferation, and decreased apoptosis by interaction with another member of the FOX gene family, FOXM1 [65,66,106]. Referring to FOXM1, it was previously shown to be an important tumor promoter in HCC progression, and its upregulation is correlated with a poor prognosis [106]. Recent investigations suggest a connection between CCAT2 and FOXM1 via a regulatory loop involving miR-34a. The study highlighted that increased levels of FOXM1 were correlated with CCAT2 in HCC cell lines. Bioinformatic analysis indicated miR-34a as the regulatory link between CCAT2 and FOXM1, as both transcripts have the competing binding ability for miR-34a. The direct interaction between CCAT2 and miR-34a was evidenced by dual-luciferase reporter assay analysis using miR-34a and CCAT2-WT and CCAT2 miR-34a mutated binding site [66]. Thus, overexpression of CCAT2 might be responsible for the mitigation of the tumor suppression ability of miR-34a, which has been also shown to promote the activation of cellular senescence programs by downregulating the FOXM1 oncogene independent of CCAT2 [107].
A different study on the role of CCAT2 in HCC has highlighted its involvement in autophagy and metastasis programs via a bimodal cytoplasmic and nuclear mechanism [69]. The CCAT2 localization in the HCC cell cytoplasm promotes its ceRNA interaction with miR-4496, a previously described tumor suppressor [69]. On the other hand, CCAT2 interacts with the RNA binding protein (RBP) ELAVL1/HuR at the nuclear level, a key factor involved in mRNA stabilization, previously shown to promote proliferation, apoptosis, and differentiation in multiple types of cancer [108]. Shi et al. showed that ELAVL1 knockdown decreased Agt5 expression, this effect being restored through CCAT2 upregulation. These findings indicate that CCAT2 promotes the proliferative processes by targeting miR-4496 and the activation of autophagy via, and interacting with ELAVL1, and downregulating Agt5. [69].
One of the CCAT2s main regulatory ceRNA targets is miR-145, a tumor suppressor described in various solid human cancers, including colon, gastric, and prostate cancers [109,110]. In these cancers, miR-145 inhibits tumor proliferation by inhibiting the MYC and Friend leukemia virus integration 1 (FLI-1) genes [111,112]. The opposite regulatory effects between miR-145 and CCAT2 on the MYC pathway suggested a possible modulatory relationship between these two ncRNAs that was further investigated. Considering miR-145 biogenesis and using bioinformatic analysis, interaction with the pre-miR-145 was found. This relationship was confirmed on cellular models that support downregulation of the mature miR-145 by direct interaction between CCAT2 and pre-miR-145 at the nuclear level [49]. Additionally, a direct bidirectional modulatory relationship between miR-145 and the oncogenic miR-21 was shown, in which miR-145 negatively regulates miR-21 in colon cancer cell lines [113]. Therefore, by inhibiting miR-145 maturation, CCAT2 indirectly upregulates the miR-21, which promotes proliferation and is associated with a poor prognosis [49].
The CCAT2/miR-145 regulation axis was also described in HCC progression via modulation of the MDM2 gene [45]. MDM2 is an oncogene previously reported in epithelial cancers, high-grade tumors, being a signature biomarker in multiple soft tissues [114]. In HCC, the MDM2–p53 pathway is an essential pathway in tumorigenesis, being altered in 25% of cases [115]. In physiological conditions, p53 enhances miR-145 activity, which was shown to act on MDM2. In human epithelial cancers, when an alteration of the p53 gene is produced, the miR-145 is downregulated, which removes the inhibition on the MDM2 oncogene [116]. In HCC, miR-145 expression was inversely correlated with MDM2 and was shown to directly interact with it by dual-luciferase reporter assay [45].
Further analysis into the mechanisms by which CCAT2 promotes human esophageal carcinoma cell (ESCC) progression led to the identification by bioinformatics analysis of miR-200b as a possible target. The direct interaction between CCAT2 and miR-200b was confirmed by a dual-luciferase reporter assay [60]. Functionally, miR-200b is a tumor suppressor miRNA that was reported in ESCC to act as an inhibitor of cellular proliferation and invasion [117,118]. Moreover, miR-200b was shown to have a complementarity sequence with the insulin-like growth factor 2 mRNA-binding protein 2 (IGF2BP2) oncogene [60]. IGF2BP2 and thymidine kinase 1 (TK1) are two important oncogenes with aberrant expression in ESCC that are associated with advanced disease and poor prognosis. IGF2BP2 was shown to act as a reader on the N6-methyladenosine (m6a) to modulate tumor progression [119]. Wu et al. confirmed by in vitro experiments the indirect positive regulatory effect of CCAT2 on the IGF2BP2 through sponging of miR-200b. Moreover, the regulatory loop includes the modulation of TK1 mRNA methylation by the IGF2BP2, which stabilizes the TK1 and enhances its expression. Therefore, the tumor promoter effects of CCAT2 in ESCC are directed by the CCAT2/miR-200b sponging, which limits the miR-200b levels and inhibitory effect on IGF2BP2. Increased IGF2BP2 protein levels recognize TK1 m6a modification and maintain its stability, inducing cellular proliferation and invasion [60].
CCAT2 was highly expressed in endometrial cancer tissue and cell lines; its expression is correlated with increased cellular proliferation and metastasis. The antiapoptotic effect of CCAT2 is switched on by sponging off the tumor suppressor miR-216b. At normal levels, miR-216b is an inhibitor of the anti-apoptotic gene BCL-2. CCAT2 contains the binding site for miR-216b and was revealed to directly interact with the miRNA. The increased levels in CCAT2 via the inhibition of miR-216b enhanced the activity of BCL-2, which in turn activated the PTEN/PI3K/AKT and mTOR signaling pathways that promote cellular proliferation and malignant transformation [41].
As another ceRNA target of CCAT2, miR-143 is a tumor-suppressor miRNA with important roles in the inhibition of major oncogenes involved in proliferation and treatment resistance. miR-143 was investigated in colon, breast, and cervical cancers [46]. Bi et al. predicted by bioinformatic analysis that miR-143 was a possible target for CCAT2 [83]. The direct interaction between CCAT2 and miR-14 was revealed in osteosarcoma cell lines. Furthermore, miR-143 is known to interact with the FOSL2 [120]. FOSL2 is a known oncogene that promotes metastasis and angiogenesis in breast and lung cancer [121,122]. Therefore, a regulatory axis between FOSL2, CCAT2, and miR-143 was described, with CCAT2 upregulation modulating metastasis and proliferation through indirect upregulation of FOSL2 oncogene via sponging of the miR-143 [83].

2.6. Therapeutic Resistance

The expression level of CCAT2 was associated in several cases with chemotherapeutic and radiotherapeutic resistance in gliomas, CRC, thyroid, breast, and esophageal cancers [32,59,89,123,124]. The roles of lncRNA CCAT2 on modulating therapeutic resistance in these cancers are presented in Table 2.
The regulatory roles of CCAT2 in cancer therapeutic resistance are related to its direct inhibition of tumor suppressor miRNAs and through its effect on the DNA damage response pathway and chromosomal instability [32,44,59]. The regulatory axis between CCAT2/miR-424/CHK1 was proven in gliomas, showing a direct correlation between CCAT2 expression direct inhibition of miR-424 and upregulation of CHK1. Interestingly, CHK1 was described as an important modulator of chemotherapy and radiotherapy resistance in acute myeloid leukemia and glial tumors and via its effect in the DNA damage response [124,126,127]. The CCAT2 upregulation in glioblastoma cell lines was associated with upregulation of the CHK1 gene and increased resistance to teniposide, temozolomide, vincristine, and cisplatin [44].
The previously described regulatory axis between miR-145 and CCAT2 was further investigated in ESCC lines, focusing on the axis’s role in promoting radiotherapy resistance. miR-145/p53 pathway downregulation was shown to promote radioresistance by altering the normal response to DNA damage and inhibiting apoptosis [128]. Additionally, miR-145 was shown to inhibit the expression level of P70 ribosomal protein S6 kinase 1 (p70S6K1), an important effector on the mTOR pathway that is involved in radioresistance [129,130]. p70S6K1 upregulation induces downstream activation of the Akt/ERK/p70S6K1 signaling pathway, enhancing cellular viability and proliferation [59]. Moreover, the enhancing effect of CCAT2 on radioresistance is influenced through its negative regulatory roles on p53 and p21 proteins, which are essential for cell cycle arrest, DNA repair, and promotion of apoptosis [59]. Targeting CCAT2 in ESCC demands further investigation to better understand its implications on tumor progression and involvement in tumor radioresistance [58].

2.7. CCAT2 Enhancer Activity and Interactions with RBPs

CCAT2 was upregulated in pituitary adenomas in comparison with adjacent normal tissue and associated with an aggressive tumor phenotype. CCAT2 overexpression was partly explained by the increased E2F1 factor, which binds to the CCAT2 promoter region and facilitates CCAT2 transcription. Increased CCAT2 levels were shown to inhibit degradation of the Securin protein, which could explain the aggressive phenotype of the pituitary adenomas [131]. Additionally, in a different study, inhibiting CCAT2 reduced proliferation, invasion, and increased cellular apoptosis—effects correlated with reduced expression of TGF-β, MAD homolog 2, and α-SMA proteins. These results highlight the CCAT2 regulatory role on the TGF-β signaling pathway [92]. Moreover, CCAT2’s positive effect on cellular proliferation is induced by suppressing p15 protein, a tumor suppressor protein that arrests the cell cycle in G1 by interacting with the histone methyltransferase EZH2 [91].
A recent study by Zang et al. investigated the mechanisms by which CCAT2 promotes cellular proliferation and showed that it directly binds to YAP protein and inhibits its phosphorylation, which further promotes YAP nuclear translocation and activation of YAP’s downstream oncogenic targets CTGF, CYR62, and AMOTL2 [132].
Investigations into the CCAT2 modulatory role on HCC proliferation and metastasis revealed that its expression is closely correlated with the NDRG1 gene. CCAT2 promotes the expression of the NDRG1 gene by enhancing its promoter activity [68]. NDRG1 is an important oncogene in HCC and is involved in promoting proliferation, metastasis, and poor prognosis [133]. Similar to other cancers, upregulation of CCAT2 enhances the EMT and is associated with a poor prognosis. HepG2, SMMC772, and MHCC97H cell lines were used to investigate the promoter role of CCAT2 on the EMT showing that its expression is positively correlated with Vimentin and Snail2 and negatively correlated with E-cadherin [67].
In vitro downregulation of CCAT2 in pancreatic cell lines PANC-1, SW1990, and PC-3 hindered proliferative and invasive capabilities of the cells, presumably due to regulatory interaction with the mitogen-activated protein kinase (MAPK) pathway [70], an important pathway involved in the translation of the signals extracellular messengers to intracellular responses [23].
In neuroblastoma, a malignant embryonal solid tumor that most frequently affects children, CCAT2 was found at increased levels in tumor tissue compared with adjacent normal tissue, and its expression was associated with increased cellular proliferation and lower three-year survival. Chen et el. revealed that part of the CCAT2 effect on cellular proliferation is due to inhibition of tumor suppressor protein p53 and enhancement of the anti-apoptotic protein Bcl-2 [75].
CCAT2 upregulation promotes EMT by upregulating ZEB2 gene expression and decreasing the expression of the E-cadherin gene. CCAT2 modulates EMT by interacting epigenetically with EZH2, H327me3, and LSD1. Increased occupancy of these genes decreases their binding affinity across the promoters of E-cadherin and LATS2. These results were validated on in vivo models in which inhibition of CCAT2 was associated with increased expression of E-cadherin and LATS2 and slower tumor growth [62].
Moreover, CCAT2 expression in gastric cancer modulates the expression of the POU5F1B, a retrogene located adjacent to MYC that inhibits apoptosis and stimulates angiogenesis and metastasis [134,135]. CCAT2 silencing is increasing apoptosis and autophagy in vitro through a downregulation of PI3K and mTOR pathways, suggesting that CCAT2 has also a regulatory effect on the PI3K/mTOR signaling pathways [63,134].
Inhibition of CCAT2 in ovarian cancer cell lines reduced invasiveness and cellular progression and inhibited EMT by downregulating vimentin and N-cadherin and upregulating E-cadherin [28]. Interestingly, treatment with calcitriol (vitamin D) on ovarian cancer cell lines had the same inhibitory effect that was achieved by silencing CCAT2. Calcitriol stimulation highlighted an inhibitory effect on CCAT2’s regulatory activity by decreasing the affinity between the transcription factor TCF7L2 and the MYC promoter, hindering its expression [98] and therefore supporting a possible place for calcitriol as an inhibitor of CCAT2 in cancer.

3. CCAT2 as a Potential Biomarker

CCAT2 expression levels and their influence on cancer progression have been assessed in various cancer types and compared with survival data and prognostic factors, including proliferation, tumor size, migration, and metastasis (Table 1) [50,136]. Various studies analyzed the association between CCAT2 expression level in cancer, identifying a direct association between CCAT2 upregulation and disease progression, invasion, positive lymph nodes, and metastasis [17,137]. When comparing the tumor tissue with adjacent normal tissue in solid cancers, CCAT2 expression was upregulated in the tumor tissue, supporting its role as a potential biomarker [17,137]. A meta-analysis by Jing et al. indicated the correlations between CCAT2 expression level in tumor tissue and the clinical stage and outcome and identified a positive correlation between CCAT2 expression level and the risk of developing distant metastasis, positive lymph nodes, shorter progression-free survival, and poorer overall survival [138].
CCAT2 expression was reported as being elevated in the plasma of cancer patients with cervical cancer, gastric cancer, colon cancer, lung cancer, and multiple myeloma [64,80,86,95,139]. The serum expression levels of a 3 lncRNA panel, which includes CCAT2, were able to distinguish early-stage cervical cancer patients from normal controls, with an area under the curve of 0.894, 67.1% sensitivity, and 96.1% specificity [95]. Therefore, this model validates a simple non-invasive procedure that has the potential to use serum samples to better identify early-stage cervical cancer. In gastric cancer, CCAT2 was investigated as part of a 5 lncRNAs (TINCR, CCAT2, AOC4P, BANCR, and LINC00857) plasma panel. Their lncRNA-based index managed to outperform the CEA-based index (p < 0.001) when distinguishing gastric cancer patients from healthy patients. Additionally, the expression levels of the lncRNAs part of the index decreased post-surgically, suggesting their role in dynamic monitoring of the patients for relapse [64]. In multiple myeloma, CCAT2 was upregulated in both peripheral blood and bone marrow when comparing patients with healthy controls. The CCAT2 expression level was correlated with International Scoring System stages, kidney function, and light chain concentrations. Moreover, when CCAT2 expression level was associated with classical multiple myeloma biomarkers such as IgA, β2MG, and HGB, the area under the curve (AUC) was significantly improved up to 0.974 (95% CI 0.958~0.990; p < 0.001). In addition, integrating the CCAT2 expression with classical biomarkers improved the sensitivity and specificity of multiple myeloma diagnoses [86].
In CRC, CCAT2 was investigated as part of 10 lncRNAs panels including CCAT1, H19, HOTAIR, HULC, MALAT1, PCAT1, MEG3, PTENP1, and TUSC7 as potential biomarkers for early detection of CRC from stool samples. The diagnostic performance of the combined lncRNA model when distinguishing the CRC from non-malignant samples had an area under the ROC curve of 0.8554 in the training set and 0.8465 in the validation set [52]. Although more in-depth studies on the intricate regulatory functions of CCAT2 in cancer are needed, it has the potential to be used as a prognostic biomarker and as a possible therapeutic target in CRC [50,139,140]. On the other hand, when CCAT1, CCAT2, and MYC expression levels were assessed as risk factors for predicting early-stage CRC metastasis, the results were limited for two lncRNAs. Even though, the individual expression of CCAT2 was individually upregulated in CRC tissue when compared with the adjacent control, its expression did not improve a 19 gene classifier, ColoMet19, previously proposed by the same authors for early-stage CRC metastasis. Nevertheless, CCAT2 expression was higher in metastatic patients compared with metastatic negative patients. Even though these results support the differential expression of CCAT2 in CRC samples when comparing with normal tissue and in metastatic vs non-metastatic CRC, they show that when integrating the CCAT2 expression level into the global genomic alterations that occur in cancer, its role fails to significantly influence the previously described ColoMet19 predictor [54].

4. Conclusions

The role of CCAT2 in either initiating or promoting the oncological phenotype of multiple cancer types is still an expanding niche in the area of tumor-associated non-coding RNAs. Currently, the main highlighted and reoccurring regulatory interactions with tumor-promoting signaling pathways are Wnt/β-catenin and MYC. Additionally, novel insight regarding CCAT2’s implication in promoting cancer-associated genomic instability hints towards the multiple implications of these transcripts across the various types of cancers associated with chromosomal abnormalities. It is worth mentioning both the direct (ceRNA activity) and indirect (transcriptional regulation) interactions with other species of ncRNAs, especially miRNAs, which add a layer of complexity in the regulatory axis in which CCAT2 is involved. In most described studies, CCAT2 was highlighted as a transcript with both biomarker and prognostic potential, as its overexpression was correlated with poor prognosis, presence of metastases, and reduced progression-free and overall survival. Therefore, CCAT2 could be considered a valuable pan-cancer biomarker that is associated with a more aggressive disease course, one with potential as a therapeutic target of new treatments or for use in strategies for overcoming chemoresistance and radioresistance [136,137,138].
Although most studies offer similar results regarding CCAT2 expression in cancer tissue and adjacent control, with significant differences between the study samples, recent studies on Saudi and Iranian populations that investigated CCAT2 expression in CRC and lung cancer failed to reproduce the results. The divergence in results should be further investigated in similar populations or studied in a larger sample to understand whether the cause was due to study sample characteristics or due to differences in etiology or lncRNA signatures [51,79].

Author Contributions

Conceptualization R.P. and I.B.-N.; methodology G.A.C.; data curation R.P., R.D. and I.B.-N.; writing—original draft preparation R.P., R.D. and A.N.; writing—review and editing G.A.C. and I.B.-N.; visualization, A.N.; supervision, G.A.C. and I.B.-N. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was supported thorough projects CNFIS-FDI-2021-0595, “Development and support of the medical entrepreneurial culture among UMF Cluj-Napoca students through dissemination, training and mentoring”, CNFIS-FDI-2021-0594 “Increasing the capacity, competitiveness and visibility of The UMPh Iuliu Hațieganu research by training and supporting multidisciplinary teams in research excellence”, and from Competitiveness Operational Program, 2014–2020, entitled “Clinical and economical impact of personalized targeted anti-microRNA therapies in reconverting lung cancer chemoresistance—CANTEMIR”, grant no. 35/01.09.2016, MySMIS 103375.

Acknowledgments

The first author, R.P., is currently a Fulbright Visiting Researcher at MD Anderson Cancer Center in Calin’s laboratory and would like to acknowledge the support received from The Fulbright Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mattick, J.S.; Rinn, J.L. Discovery and annotation of long noncoding RNAs. Nat. Struct. Mol. Biol. 2015, 22, 5–7. [Google Scholar] [CrossRef]
  2. van Leeuwen, S.; Mikkers, H. Long non-coding RNAs: Guardians of development. Differ. Res. Biol. Divers 2010, 80, 175–183. [Google Scholar] [CrossRef]
  3. Mercer, T.R.; Dinger, M.E.; Mattick, J.S. Long non-coding RNAs: Insights into functions. Nat. Rev. Genet. 2009, 10, 155–159. [Google Scholar] [CrossRef] [PubMed]
  4. Deng, H.; Wang, J.M.; Li, M.; Tang, R.; Tang, K.; Su, Y.; Hou, Y.; Zhang, J. Long non-coding RNAs: New biomarkers for prognosis and diagnosis of colon cancer. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2017, 39, 1010428317706332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kondo, Y.; Shinjo, K.; Katsushima, K. Long non-coding RNAs as an epigenetic regulator in human cancers. Cancer Sci. 2017, 108, 1927–1933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Chandra Gupta, S.; Nandan Tripathi, Y. Potential of long non-coding RNAs in cancer patients: From biomarkers to therapeutic targets. Int. J. Cancer 2017, 140, 1955–1967. [Google Scholar] [CrossRef]
  7. Qiu, L.; Tang, Q.; Li, G.; Chen, K. Long non-coding RNAs as biomarkers and therapeutic targets: Recent insights into hepatocellular carcinoma. Life Sci. 2017, 191, 273–282. [Google Scholar] [CrossRef]
  8. Brown, C.J.; Hendrich, B.D.; Rupert, J.L.; Lafrenière, R.G.; Xing, Y.; Lawrence, J.; Willard, H.F. The human XIST gene: Analysis of a 17 kb inactive X-specific RNA that contains conserved repeats and is highly localized within the nucleus. Cell 1992, 71, 527–542. [Google Scholar] [CrossRef]
  9. Woo, C.J.; Kingston, R.E. HOTAIR lifts noncoding RNAs to new levels. Cell 2007, 129, 1257–1259. [Google Scholar] [CrossRef] [Green Version]
  10. Schmidt, L.H.; Spieker, T.; Koschmieder, S.; Schäffers, S.; Humberg, J.; Jungen, D.; Bulk, E.; Hascher, A.; Wittmer, D.; Marra, A.; et al. The long noncoding MALAT-1 RNA indicates a poor prognosis in non-small cell lung cancer and induces migration and tumor growth. J. Thorac. Oncol. Off. Publ. Int. Assoc. Study Lung Cancer 2011, 6, 1984–1992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Gupta, R.A.; Shah, N.; Wang, K.C.; Kim, J.; Horlings, H.M.; Wong, D.J.; Tsai, M.-C.; Hung, T.; Argani, P.; Rinn, J.L.; et al. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature 2010, 464, 1071–1076. [Google Scholar] [CrossRef]
  12. Rinn, J.L.; Kertesz, M.; Wang, J.K.; Squazzo, S.L.; Xu, X.; Brugmann, S.A.; Goodnough, L.H.; Helms, J.A.; Farnham, P.J.; Segal, E.; et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell 2007, 129, 1311–1323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Ji, P.; Diederichs, S.; Wang, W.; Böing, S.; Metzger, R.; Schneider, P.M.; Tidow, N.; Brandt, B.; Buerger, H.; Bulk, E.; et al. MALAT-1, a novel noncoding RNA, and thymosin beta4 predict metastasis and survival in early-stage non-small cell lung cancer. Oncogene 2003, 22, 8031–8041. [Google Scholar] [CrossRef] [Green Version]
  14. Hutchinson, J.N.; Ensminger, A.W.; Clemson, C.M.; Lynch, C.R.; Lawrence, J.B.; Chess, A. A screen for nuclear transcripts identifies two linked noncoding RNAs associated with SC35 splicing domains. BMC Genom. 2007, 8, 39. [Google Scholar] [CrossRef] [Green Version]
  15. Tripathi, V.; Ellis, J.D.; Shen, Z.; Song, D.Y.; Pan, Q.; Watt, A.T.; Freier, S.M.; Bennett, C.F.; Sharma, A.; Bubulya, P.A.; et al. The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Mol. Cell 2010, 39, 925–938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Ling, H.; Spizzo, R.; Atlasi, Y.; Nicoloso, M.; Shimizu, M.; Redis, R.S.; Nishida, N.; Gafa, R.; Song, J.; Guo, Z.; et al. CCAT2, a novel noncoding RNA mapping to 8q24, underlies metastatic progression and chromosomal instability in colon cancer. Genome Res. 2013, 23, 1446–1461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Ghafouri-Fard, S.; Taheri, M. Colon Cancer-Associated Transcripts 1 and 2: Roles and functions in human cancers. J. Cell Physiol. 2019, 234, 14581–14600. [Google Scholar] [CrossRef]
  18. Nissan, A.; Stojadinovic, A.; Mitrani-Rosenbaum, S.; Halle, D.; Grinbaum, R.; Roistacher, M.; Bochem, A.; Dayanc, B.E.; Ritter, G.; Gomceli, I.; et al. Colon cancer associated transcript-1: A novel RNA expressed in malignant and pre-malignant human tissues. Int. J. Cancer 2012, 130, 1598–1606. [Google Scholar] [CrossRef] [PubMed]
  19. Liu, Z.; Chen, Q.; Hann, S.S. The functions and oncogenic roles of CCAT1 in human cancer. Biomed. Pharmacother. 2019, 115, 108943. [Google Scholar] [CrossRef]
  20. Xiang, J.-F.; Yin, Q.-F.; Chen, T.; Zhang, Y.; Zhang, X.-O.; Wu, Z.; Zhang, S.; Wang, H.-B.; Ge, J.; Lu, X.; et al. Human colorectal cancer-specific CCAT1-L lncRNA regulates long-range chromatin interactions at the MYC locus. Cell Res. 2014, 24, 513–531. [Google Scholar] [CrossRef] [Green Version]
  21. Guo, X.; Hua, Y. CCAT1: An oncogenic long noncoding RNA in human cancers. J. Cancer Res. Clin. Oncol. 2017, 143, 555–562. [Google Scholar] [CrossRef]
  22. Jiang, Y.; Jiang, Y.-Y.; Xie, J.-J.; Mayakonda, A.; Hazawa, M.; Chen, L.; Xiao, J.-F.; Li, C.-Q.; Huang, M.-L.; Ding, L.-W.; et al. Co-activation of super-enhancer-driven CCAT1 by TP63 and SOX2 promotes squamous cancer progression. Nat. Commun. 2018, 9, 3619. [Google Scholar] [CrossRef]
  23. Braicu, C.; Buse, M.; Busuioc, C.; Drula, R.; Gulei, D.; Raduly, L.; Rusu, A.; Irimie, A.; Atanasov, A.G.; Slaby, O.; et al. A Comprehensive Review on MAPK: A Promising Therapeutic Target in Cancer. Cancers 2019, 11, 1618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Ghoussaini, M.; Song, H.; Koessler, T.; Al Olama, A.A.; Kote-Jarai, Z.; Driver, K.E.; Pooley, K.A.; Ramus, S.J.; Kjaer, S.K.; Hogdall, E.; et al. Multiple loci with different cancer specificities within the 8q24 gene desert. J. Natl. Cancer Inst. 2008, 100, 962–966. [Google Scholar] [CrossRef] [Green Version]
  25. Wright, J.B.; Brown, S.J.; Cole, M.D. Upregulation of c-MYC in cis through a large chromatin loop linked to a cancer risk-associated single-nucleotide polymorphism in colorectal cancer cells. Mol. Cell Biol. 2010, 30, 1411–1420. [Google Scholar] [CrossRef] [Green Version]
  26. Yang, C.; Li, F.; Zhou, W.; Huang, J. Knockdown of long non-coding RNA CCAT2 suppresses growth and metastasis of esophageal squamous cell carcinoma by inhibiting the β-catenin/WISP1 signaling pathway. J. Int. Med. Res. 2021, 49, 3000605211019938. [Google Scholar] [CrossRef] [PubMed]
  27. He, P.; Xiong, G.; Guo, W.; Jiang, G.; Li, Y.; Li, H. Long non-coding RNA CCAT2 promotes prostate cancer cell proliferation and invasion by regulating the Wnt/β-catenin signaling pathway. Oncol. Lett. 2020, 20, 97. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, B.; Liu, M.; Zhuang, R.; Jiang, J.; Gao, J.; Wang, H.; Chen, H.; Zhang, Z.; Kuang, Y.; Li, P. Long non-coding RNA CCAT2 promotes epithelial-mesenchymal transition involving Wnt/beta-catenin pathway in epithelial ovarian carcinoma cells. Oncol. Lett. 2018, 15, 3369–3375. [Google Scholar]
  29. Huang, J.-L.; Liao, Y.; Qiu, M.-X.; Li, J.; An, Y. Long non-coding RNA CCAT2 promotes cell proliferation and invasion through regulating Wnt/beta-catenin signaling pathway in clear cell renal cell carcinoma. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2017, 39, 1010428317711314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Cai, Y.; He, J.; Zhang, D. Long noncoding RNA CCAT2 promotes breast tumor growth by regulating the Wnt signaling pathway. OncoTargets Ther. 2015, 8, 2657–2664. [Google Scholar]
  31. Redis, R.S.; Vela, L.E.; Lu, W.; Ferreira de Oliveira, J.; Ivan, C.; Rodriguez-Aguayo, C.; Adamoski, D.; Pasculli, B.; Taguchi, A.; Chen, Y.; et al. Allele-Specific Reprogramming of Cancer Metabolism by the Long Non-coding RNA CCAT2. Mol. Cell. 2016, 61, 520–534. [Google Scholar] [CrossRef] [Green Version]
  32. Chen, B.; Dragomir, M.P.; Fabris, L.; Bayraktar, R.; Knutsen, E.; Liu, X.; Tang, C.; Li, Y.; Shimura, T.; Ivkovic, T.C.; et al. The Long Noncoding RNA CCAT2 Induces Chromosomal Instability Through BOP1-AURKB Signaling. Gastroenterology 2020, 159, 2146–2162.e33. [Google Scholar] [CrossRef]
  33. Baysal, B.E.; Sharma, S.; Hashemikhabir, S.; Janga, S.C. RNA Editing in Pathogenesis of Cancer. Cancer Res. 2017, 77, 3733–3739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Christofi, T.; Zaravinos, A. RNA editing in the forefront of epitranscriptomics and human health. J. Transl. Med. 2019, 17, 319. [Google Scholar] [CrossRef] [PubMed]
  35. Shah, M.Y.; Ferracin, M.; Pileczki, V.; Chen, B.; Redis, R.; Fabris, L.; Zhang, X.; Ivan, C.; Shimizu, M.; Rodriguez-Aguayo, C.; et al. Cancer-associated rs6983267 SNP and its accompanying long noncoding RNA CCAT2 induce myeloid malignancies via unique SNP-specific RNA mutations. Genome Res. 2018, 28, 432–447. [Google Scholar] [CrossRef] [Green Version]
  36. Zhou, R.-S.; Zhang, E.-X.; Sun, Q.-F.; Ye, Z.-J.; Liu, J.-W.; Zhou, D.-H.; Tang, Y. Integrated analysis of lncRNA-miRNA-mRNA ceRNA network in squamous cell carcinoma of tongue. BMC Cancer 2019, 19, 779. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, J.; Li, H.; Zheng, B.; Sun, L.; Yuan, Y.; Xing, C. Competitive Endogenous RNA (ceRNA) Regulation Network of lncRNA-miRNA-mRNA in Colorectal Carcinogenesis. Dig. Dis. Sci. 2019, 64, 1868–1877. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, Z.; Qian, W.; Wang, S.; Ji, D.; Wang, Q.; Li, J.; Peng, W.; Gu, J.; Hu, T.; Ji, B.; et al. Analysis of lncRNA-Associated ceRNA Network Reveals Potential lncRNA Biomarkers in Human Colon Adenocarcinoma. Cell Physiol. Biochem. Int. J. Exp. Cell Physiol. Biochem. Pharmacol. 2018, 49, 1778–1791. [Google Scholar] [CrossRef]
  39. Hua, F.; Li, C.-H.; Chen, X.-G.; Liu, X.-P. Long Noncoding RNA CCAT2 Knockdown Suppresses Tumorous Progression by Sponging miR-424 in Epithelial Ovarian Cancer. Oncol. Res. 2018, 26, 241–247. [Google Scholar] [CrossRef]
  40. Thomson, D.W.; Dinger, M.E. Endogenous microRNA sponges: Evidence and controversy. Nat. Rev. Genet. 2016, 17, 272–283. [Google Scholar] [CrossRef] [PubMed]
  41. Xie, P.; Cao, H.; Li, Y.; Wang, J.; Cui, Z. Knockdown of lncRNA CCAT2 inhibits endometrial cancer cells growth and metastasis via sponging miR-216b. Cancer Biomark. Sect Dis. Markers 2017, 21, 123–133. [Google Scholar] [CrossRef]
  42. Hu, G.-D.; Wang, C.-X.; Wang, H.-Y.; Wang, Y.-Q.; Hu, S.; Cao, Z.-W.; Min, B.; Li, L.; Tian, X.-F.; Hu, H.-B. Long noncoding RNA CCAT2 functions as a competitive endogenous RNA to regulate FOXC1 expression by sponging miR-23b-5p in lung adenocarcinoma. J. Cell Biochem. 2018, 120, 7998–8007. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, S.-L.; Shu, Y.-G.; Tao, M.-Y. LncRNA CCAT2 promotes angiogenesis in glioma through activation of VEGFA signalling by sponging miR-424. Mol. Cell Biochem. 2020, 468, 69–82. [Google Scholar] [CrossRef] [PubMed]
  44. Ding, J.; Zhang, L.; Chen, S.; Cao, H.; Xu, C.; Wang, X. lncRNA CCAT2 Enhanced Resistance of Glioma Cells Against Chemodrugs by Disturbing the Normal Function of miR-424. OncoTargets Ther. 2020, 13, 1431–1445. [Google Scholar] [CrossRef] [Green Version]
  45. Niu, C.; Wang, L.; Ye, W.; Guo, S.; Bao, X.; Wang, Y.; Xia, Z.; Chen, R.; Liu, C.; Lin, X.; et al. CCAT2 contributes to hepatocellular carcinoma progression via inhibiting miR-145 maturation to induce MDM2 expression. J. Cell Physiol. 2020, 235, 6307–6320. [Google Scholar] [CrossRef] [PubMed]
  46. Farooqi, A.A.; Qureshi, M.Z.; Attar, R.; Alhewairini, S.S.; Fayyaz, S.; Sabitaliyevich, U.Y.; Duisenbayevich, T.M.; Alaaeddine, N. MicroRNA-143 as a new weapon against cancer: Overview of the mechanistic insights and long non-coding RNA mediated regulation of miRNA-143 in different cancers. Cell Mol. Biol. 2019, 65, 1–5. [Google Scholar] [CrossRef]
  47. Zhang, J.; Jiang, Y.; Zhu, J.; Wu, T.; Ma, J.; Du, C.; Chen, S.; Li, T.; Han, J.; Wang, X. Overexpression of long non-coding RNA colon cancer-associated transcript 2 is associated with advanced tumor progression and poor prognosis in patients with colorectal cancer. Oncol. Lett. 2017, 14, 6907–6914. [Google Scholar] [CrossRef] [PubMed]
  48. Kasagi, Y.; Oki, E.; Ando, K.; Ito, S.; Iguchi, T.; Sugiyama, M.; Nakashima, Y.; Ohgaki, K.; Saeki, H.; Mimori, K.; et al. The Expression of CCAT2, a Novel Long Noncoding RNA Transcript, and rs6983267 Single-Nucleotide Polymorphism Genotypes in Colorectal Cancers. Oncology 2017, 92, 48–54. [Google Scholar] [CrossRef]
  49. Yu, Y.; Nangia-Makker, P.; Farhana, L.; Majumdar, A.P.N. A novel mechanism of lncRNA and miRNA interaction: CCAT2 regulates miR-145 expression by suppressing its maturation process in colon cancer cells. Mol. Cancer 2017, 16, 155. [Google Scholar] [CrossRef] [PubMed]
  50. Ozawa, T.; Matsuyama, T.; Toiyama, Y.; Takahashi, N.; Ishikawa, T.; Uetake, H.; Yamada, Y.; Kusunoki, M.; Calin, G.; Goel, A. CCAT1 and CCAT2 long noncoding RNAs, located within the 8q.24.21 “gene desert”, serve as important prognostic biomarkers in colorectal cancer. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2017, 28, 1882–1888. [Google Scholar] [CrossRef]
  51. Siddique, H.; Al-Ghafari, A.; Choudhry, H.; AlTurki, S.; Alshaibi, H.; Al Doghaither, H.; Alsufiani, H. Long Noncoding RNAs as Prognostic Markers for Colorectal Cancer in Saudi Patients. Genet. Test Mol. Biomark. 2019, 23, 509–514. [Google Scholar] [CrossRef]
  52. Gharib, E.; Nazemalhosseini-Mojarad, E.; Baghdar, K.; Nayeri, Z.; Sadeghi, H.; Rezasoltani, S.; Jamshidi-Fard, A.; Larki, P.; Sadeghi, A.; Hashemi, M.; et al. Identification of a stool long non-coding RNAs panel as a potential biomarker for early detection of colorectal cancer. J. Clin. Lab. Anal. 2021, 35, e23601. [Google Scholar] [CrossRef] [PubMed]
  53. Gao, P.; Sun, D.; Guo, H.; Wu, Z.; Chen, J. LncRNA CCAT2 promotes proliferation and suppresses apoptosis of colorectal cancer cells. J. BUON Off. J. Balk Union Oncol. 2020, 25, 1840–1846. [Google Scholar]
  54. Thean, L.F.; Blöcker, C.; Li, H.H.; Lo, M.; Wong, M.; Tang, C.L.; Tan, E.K.W.; Rozen, S.G.; Cheah, P.Y. Enhancer-derived long non-coding RNAs CCAT1 and CCAT2 at rs6983267 has limited predictability for early stage colorectal carcinoma metastasis. Sci. Rep. 2021, 11, 404. Available online: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7801656/ (accessed on 24 May 2021). [CrossRef]
  55. Zhang, X.; Xu, Y.; He, C.; Guo, X.; Zhang, J.; He, C.; Zhang, L.; Kong, M.; Chen, B.; Zhu, C. Elevated expression of CCAT2 is associated with poor prognosis in esophageal squamous cell carcinoma. J. Surg. Oncol. 2015, 111, 834–839. [Google Scholar] [CrossRef]
  56. Wang, J.; Qiu, M.; Xu, Y.; Li, M.; Dong, G.; Mao, Q.; Yin, R.; Xu, L. Long noncoding RNA CCAT2 correlates with smoking in esophageal squamous cell carcinoma. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2015, 36, 5523–5528. [Google Scholar] [CrossRef] [PubMed]
  57. Wang, X.; Wang, X. Long non-coding RNA colon cancer-associated transcript 2 may promote esophageal cancer growth and metastasis by regulating the Wnt signaling pathway. Oncol. Lett. 2019, 18, 1745–1754. Available online: http://www.spandidos-publications.com/10.3892/ol.2019.10488 (accessed on 27 October 2019). [CrossRef]
  58. Ma, Y.; Hu, X.; Shang, C.; Zhong, M.; Guo, Y. Silencing of long non-coding RNA CCAT2 depressed malignancy of oral squamous cell carcinoma via Wnt/beta-catenin pathway. Tumour Biol. J. Int. Soc. Oncodevel. Biol. Med. 2017, 39, 1010428317717670. [Google Scholar]
  59. Wang, M.; Wang, L.; He, X.; Zhang, J.; Zhu, Z.; Zhang, M.; Li, X. lncRNA CCAT2 promotes radiotherapy resistance for human esophageal carcinoma cells via the miR-145/p70S6K1 and p53 pathway. Int. J. Oncol. 2020, 56, 327–336. [Google Scholar] [CrossRef] [PubMed]
  60. Wu, X.; Fan, Y.; Liu, Y.; Shen, B.; Lu, H.; Ma, H. Long Non-Coding RNA CCAT2 Promotes the Development of Esophageal Squamous Cell Carcinoma by Inhibiting miR-200b to Upregulate the IGF2BP2/TK1 Axis. Front. Oncol. 2021, 11, 680642. [Google Scholar] [CrossRef]
  61. Wang, C.-Y.; Hua, L.; Yao, K.-H.; Chen, J.-T.; Zhang, J.-J.; Hu, J.-H. Long non-coding RNA CCAT2 is up-regulated in gastric cancer and associated with poor prognosis. Int. J. Clin. Exp. Pathol. 2015, 8, 779–785. [Google Scholar]
  62. Wang, Y.-J.; Liu, J.-Z.; Lv, P.; Dang, Y.; Gao, J.-Y.; Wang, Y. Long non-coding RNA CCAT2 promotes gastric cancer proliferation and invasion by regulating the E-cadherin and LATS2. Am. J. Cancer Res. 2016, 6, 2651–2660. [Google Scholar] [PubMed]
  63. Lin, S.; Wang, H.; Yang, W.; Wang, A.; Geng, C. Silencing of Long Non-Coding RNA Colon Cancer-Associated Transcript 2 Inhibits the Growth and Metastasis of Gastric Cancer Through Blocking mTOR Signaling. OncoTargets Ther. 2020, 13, 337–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zhang, K.; Shi, H.; Xi, H.; Wu, X.; Cui, J.; Gao, Y.; Liang, W.; Hu, C.; Liu, Y.; Li, J. Genome-Wide lncRNA Microarray Profiling Identifies Novel Circulating lncRNAs for Detection of Gastric Cancer. Theranostics 2017, 7, 213–227. [Google Scholar] [CrossRef]
  65. Zhou, N.; Si, Z.; Li, T.; Chen, G.; Zhang, Z.; Qi, H. Long non-coding RNA CCAT2 functions as an oncogene in hepatocellular carcinoma, regulating cellular proliferation, migration and apoptosis. Oncol. Lett. 2016, 12, 132–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Chen, F.; Bai, G.; Li, Y.; Feng, Y.; Wang, L. A positive feedback loop of long noncoding RNA CCAT2 and FOXM1 promotes hepatocellular carcinoma growth. Am. J. Cancer Res. 2017, 7, 1423–1434. [Google Scholar]
  67. Xu, Y.; Wang, B.; Zhang, F.; Wang, A.; Du, X.; Hu, P.; Zhu, Y.; Fang, Z. Long non-coding RNA CCAT2 is associated with poor prognosis in hepatocellular carcinoma and promotes tumor metastasis by regulating Snail2-mediated epithelial-mesenchymal transition. OncoTargets Ther. 2017, 10, 1191–1198. [Google Scholar] [CrossRef] [Green Version]
  68. Liu, Y.; Wang, D.; Li, Y.; Yan, S.; Dang, H.; Yue, H.; Ling, J.; Chen, F.; Zhao, Y.; Gou, L.; et al. Long noncoding RNA CCAT2 promotes hepatocellular carcinoma proliferation and metastasis through up-regulation of NDRG1. Exp. Cell Res. 2019, 379, 19–29. [Google Scholar] [CrossRef]
  69. Shi, J.; Guo, C.; Ma, J. CCAT2 enhances autophagy-related invasion and metastasis via regulating miR-4496 and ELAVL1 in hepatocellular carcinoma. J. Cell Mol. Med. 2021, 25, 8985–8996. [Google Scholar] [CrossRef]
  70. Cai, Y.; Li, X.; Shen, P.; Zhang, D. CCAT2 is an oncogenic long non-coding RNA in pancreatic ductal adenocarcinoma. Biol. Res. 2018, 51, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Lang, H.-L.; Hu, G.-W.; Zhang, B.; Kuang, W.; Chen, Y.; Wu, L.; Xu, G.-H. Glioma cells enhance angiogenesis and inhibit endothelial cell apoptosis through the release of exosomes that contain long non-coding RNA CCAT2. Oncol. Rep. 2017, 38, 785–798. [Google Scholar] [CrossRef] [Green Version]
  72. Guo, H.; Hu, G.; Yang, Q.; Zhang, P.; Kuang, W.; Zhu, X.; Wu, L. Knockdown of long non-coding RNA CCAT2 suppressed proliferation and migration of glioma cells. Oncotarget 2016, 7, 81806–81814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Zeng, J.; Du, T.; Song, Y.; Gao, Y.; Li, F.; Wu, R.; Chen, Y.; Li, W.; Zhou, H.; Yang, Y.; et al. Knockdown of Long Noncoding RNA CCAT2 Inhibits Cellular Proliferation, Invasion, and Epithelial-Mesenchymal Transition in Glioma Cells. Oncol. Res. 2017, 25, 913–921. [Google Scholar] [CrossRef] [PubMed]
  74. Filippella, M.; Galland, F.; Kujas, M.; Young, J.; Faggiano, A.; Lombardi, G.; Colao, A.; Meduri, G.; Chanson, P. Pituitary tumour transforming gene (PTTG) expression correlates with the proliferative activity and recurrence status of pituitary adenomas: A clinical and immunohistochemical study. Clin. Endocrinol. 2006, 65, 536–543. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, M.; Zhao, M.; Hou, Y.; Zhu, B. Expression of lncRNA CCAT2 in children with neuroblastoma and its effect on cancer cell growth. Mol. Cell Biochem. 2021, 476, 1871–1879. [Google Scholar] [CrossRef]
  76. Qiu, M.; Xu, Y.; Yang, X.; Wang, J.; Hu, J.; Xu, L.; Yin, R. CCAT2 is a lung adenocarcinoma-specific long non-coding RNA and promotes invasion of non-small cell lung cancer. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2014, 35, 5375–5380. [Google Scholar] [CrossRef]
  77. Chen, S.; Wu, H.; Lv, N.; Wang, H.; Wang, Y.; Tang, Q.; Shao, H.; Sun, C. LncRNA CCAT2 predicts poor prognosis and regulates growth and metastasis in small cell lung cancer. Biomed. Pharmacother. 2016, 82, 583–588. [Google Scholar] [CrossRef]
  78. Zhao, C.; Qiao, C.; Zong, L.; Chen, Y. Long non-coding RNA-CCAT2 promotes the occurrence of non-small cell lung cancer by regulating the Wnt/beta-catenin signaling pathway. Oncol. Lett. 2018, 16, 4600–4606. [Google Scholar]
  79. Esfandi, F.; Fallah, H.; Arsang-Jang, S.; Taheri, M.; Ghafouri-Fard, S. The Expression of CCAT2, UCA1, PANDA and GHET1 Long Non-coding RNAs in Lung Cancer. Rep. Biochem. Mol. Biol. 2019, 8, 36–41. [Google Scholar]
  80. Yu, W.-L.; Yao, J.-J.; Xie, Z.-Z.; Huang, Y.-J.; Xiao, S. LncRNA PRNCR1 rs1456315 and CCAT2 rs6983267 Polymorphisms on 8q24 Associated with Lung Cancer. Int. J. Gen. Med. 2021, 14, 255–266. [Google Scholar] [CrossRef]
  81. Ruan, R.; Zhao, X.-L. LncRNA CCAT2 enhances cell proliferation via GSK3beta/beta-catenin signaling pathway in human osteosarcoma. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 2978–2984. [Google Scholar]
  82. Yan, L.; Wu, X.; Yin, X.; Du, F.; Liu, Y.; Ding, X. LncRNA CCAT2 promoted osteosarcoma cell proliferation and invasion. J. Cell Mol. Med. 2018, 22, 2592–2599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Bi, F.; Chen, C.; Fu, J.; Yu, L.; Geng, J. Inhibiting proliferation and metastasis of osteosarcoma cells by downregulation of long non-coding RNA colon cancer-associated transcript 2 targeting microRNA-143. Oncol. Lett. 2021, 21, 265. [Google Scholar] [CrossRef] [PubMed]
  84. Xin, S.; Ye, X. Knockdown of long non-coding RNA CCAT2 suppresses the progression of thyroid cancer by inhibiting the Wnt/β-catenin pathway. Int. J. Mol. Med. 2020, 46, 2047–2056. [Google Scholar] [CrossRef]
  85. Fu, W.; Wang, X.-D.; Ye, J.-D.; Jin, J.; Chen, L.; Qi, Q.-Y. CCAT2 contributes to progression and treatment resistance of thyroid carcinoma. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 12224–12231. [Google Scholar]
  86. Xu, H.; Yin, Q.; Shen, X.; Ju, S. Long non-coding RNA CCAT2 as a potential serum biomarker for diagnosis and prognosis of multiple myeloma. Ann. Hematol. 2020, 99, 2159–2171. [Google Scholar] [CrossRef] [PubMed]
  87. Li, J.-J.; Zhu, J.-F.; Zhang, F.; Zhang, P.-P.; Zhang, J.-J. Upregulation of lncRNA CCAT2 predicts poor prognosis in patients with acute myeloid leukemia and is correlated with leukemic cell proliferation. Int. J. Clin. Exp. Pathol. 2018, 11, 5658–5666. [Google Scholar]
  88. Redis, R.S.; Sieuwerts, A.M.; Look, M.P.; Tudoran, O.; Ivan, C.; Spizzo, R.; Zhang, X.; de Weerd, V.; Shimizu, M.; Ling, H. CCAT2, a novel long non-coding RNA in breast cancer: Expression study and clinical correlations. Oncotarget 2013, 4, 1748–1762. [Google Scholar] [CrossRef] [Green Version]
  89. Cai, Y.; He, J.; Zhang, D. Suppression of long non-coding RNA CCAT2 improves tamoxifen-resistant breast cancer cells’ response to tamoxifen. Mol. Biol. 2016, 50, 821–827. [Google Scholar] [CrossRef]
  90. Sarrafzadeh, S.; Geranpayeh, L.; Tasharrofi, B.; Soudyab, M.; Nikpayam, E.; Iranpour, M.; Mirfakhraie, R.; Gharesouran, J.; Ghafouri-Fard, S.; Ghafouri-Fard, S. Expression Study and Clinical Correlations of MYC and CCAT2 in Breast Cancer Patients. Iran Biomed. J. 2017, 21, 303–311. [Google Scholar] [CrossRef] [Green Version]
  91. Deng, X.; Zhao, Y.; Wu, X.; Song, G. Upregulation of CCAT2 promotes cell proliferation by repressing the P15 in breast cancer. Biomed. Pharmacother. 2017, 91, 1160–1166. [Google Scholar] [CrossRef]
  92. Wu, Z.-J.; Li, Y.; Wu, Y.-Z.; Wang, Y.; Nian, W.-Q.; Wang, L.-L.; Li, L.-C.; Luo, H.-L.; Wang, D.-L. Long non-coding RNA CCAT2 promotes the breast cancer growth and metastasis by regulating TGF-beta signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 706–714. [Google Scholar] [PubMed]
  93. Chen, X.; Liu, L.; Zhu, W. Up-regulation of long non-coding RNA CCAT2 correlates with tumor metastasis and poor prognosis in cervical squamous cell cancer patients. Int. J. Clin. Exp. Pathol. 2015, 8, 13261–13266. [Google Scholar] [PubMed]
  94. Wu, L.; Jin, L.; Zhang, W.; Zhang, L. Roles of Long Non-Coding RNA CCAT2 in Cervical Cancer Cell Growth and Apoptosis. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2016, 22, 875–879. [Google Scholar] [CrossRef] [Green Version]
  95. Wang, W.-J.; Wang, D.; Zhao, M.; Sun, X.-J.; Li, Y.; Lin, H.; Che, Y.-Q.; Huang, C. Serum lncRNAs (CCAT2, LINC01133, LINC00511) with Squamous Cell Carcinoma Antigen Panel as Novel Non-Invasive Biomarkers for Detection of Cervical Squamous Carcinoma. Cancer Manag. Res. 2020, 12, 9495–9502. [Google Scholar] [CrossRef]
  96. Wang, J.; Liu, Y.; Cai, H.; Jiang, H.; Li, W.; Shi, Y. Long coding RNA CCAT2 enhances the proliferation and epithelial-mesenchymal transition of cervical carcinoma cells via the microRNA-493-5p/CREB1 axis. Bioengineered 2021, 12, 6264–6274. [Google Scholar] [CrossRef] [PubMed]
  97. Huang, S.; Qing, C.; Huang, Z.; Zhu, Y. The long non-coding RNA CCAT2 is up-regulated in ovarian cancer and associated with poor prognosis. Diagn. Pathol. 2016, 11, 49. [Google Scholar] [CrossRef] [Green Version]
  98. Wang, L.; Zhou, S.; Guo, B. Vitamin D Suppresses Ovarian Cancer Growth and Invasion by Targeting Long Non-Coding RNA CCAT2. Int. J. Mol. Sci. 2020, 21, 2334. [Google Scholar] [CrossRef] [Green Version]
  99. Zheng, J.; Zhao, S.; He, X.; Zheng, Z.; Bai, W.; Duan, Y.; Che, Y.-Q.; Huang, C. The up-regulation of long non-coding RNA CCAT2 indicates a poor prognosis for prostate cancer and promotes metastasis by affecting epithelial-mesenchymal transition. Biochem. Biophys. Res. Commun. 2016, 480, 508–514. [Google Scholar] [CrossRef]
  100. Li, J.; Zhuang, C.; Liu, Y.; Chen, M.; Zhou, Q.; Chen, Z.; He, A.; Zhao, G.; Guo, Y.; Wu, H. shRNA targeting long non-coding RNA CCAT2 controlled by tetracycline-inducible system inhibits progression of bladder cancer cells. Oncotarget 2016, 7, 28989–28997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Xu, J.; Li, Y.; Wang, F.; Wang, X.; Cheng, B.; Ye, F.; Xie, X.; Zhou, C.; Lu, W. Suppressed miR-424 expression via upregulation of target gene Chk1 contributes to the progression of cervical cancer. Oncogene 2013, 32, 976–987. [Google Scholar] [CrossRef] [PubMed]
  102. Huang, X.; Liu, F.; Jiang, Z.; Guan, H.; Jia, Q. CREB1 Suppresses Transcription of microRNA-186 to Promote Growth, Invasion and Epithelial-Mesenchymal Transition of Gastric Cancer Cells Through the KRT8/HIF-1α Axis. Cancer Manag. Res. 2020, 12, 9097–9111. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, S.; Wang, X.; Li, J.; Meng, S.; Liang, Z.; Xu, X.; Zhu, Y.; Li, S.; Wu, J.; Xu, M.; et al. c-Met, CREB1 and EGFR are involved in miR-493-5p inhibition of EMT via AKT/GSK-3β/Snail signaling in prostate cancer. Oncotarget 2017, 8, 82303–82313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Janikova, M.; Zizkova, V.; Skarda, J.; Kharaishvili, G.; Radova, L.; Kolar, Z. Prognostic significance of miR-23b in combination with P-gp, MRP and LRP/MVP expression in non-small cell lung cancer. Neoplasma 2016, 63, 576–587. [Google Scholar] [CrossRef]
  105. Wei, L.-X.; Zhou, R.-S.; Xu, H.-F.; Wang, J.-Y.; Yuan, M.-H. High expression of FOXC1 is associated with poor clinical outcome in non-small cell lung cancer patients. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2013, 34, 941–946. [Google Scholar] [CrossRef]
  106. Chand, V.; Pandey, A.; Kopanja, D.; Guzman, G.; Raychaudhuri, P. Opposing Roles of the Fork-head box genes FoxM1 and FoxA2 in Liver Cancer. Mol. Cancer Res. 2019, 1, 383604. [Google Scholar]
  107. Xu, X.; Chen, W.; Miao, R.; Zhou, Y.; Wang, Z.; Zhang, L.; Wan, Y.; Dong, Y.; Qu, K.; Liu, C. miR-34a induces cellular senescence via modulation of telomerase activity in human hepatocellular carcinoma by targeting FoxM1/c-Myc pathway. Oncotarget 2015, 6, 3988–4004. [Google Scholar] [CrossRef] [Green Version]
  108. Abdelmohsen, K.; Gorospe, M. Posttranscriptional regulation of cancer traits by HuR. Wiley Interdiscip Rev. RNA 2010, 1, 214–229. [Google Scholar] [CrossRef]
  109. Takagi, T.; Iio, A.; Nakagawa, Y.; Naoe, T.; Tanigawa, N.; Akao, Y. Decreased expression of microRNA-143 and -145 in human gastric cancers. Oncology 2009, 77, 12–21. [Google Scholar] [CrossRef]
  110. Zhang, X.; Wu, J. Prognostic role of microRNA-145 in prostate cancer: A systems review and meta-analysis. Prostate Int. 2015, 3, 71–74. [Google Scholar] [CrossRef] [Green Version]
  111. Sachdeva, M.; Zhu, S.; Wu, F.; Wu, H.; Walia, V.; Kumar, S.; Elble, R.; Watabe, K.; Mo, Y.-Y. p53 represses c-Myc through induction of the tumor suppressor miR-145. Proc. Natl. Acad. Sci. USA 2009, 106, 3207–3212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Zhang, J.; Guo, H.; Zhang, H.; Wang, H.; Qian, G.; Fan, X.; Hoffman, A.R.; Hu, J.-F.; Ge, S. Putative tumor suppressor miR-145 inhibits colon cancer cell growth by targeting oncogene Friend leukemia virus integration 1 gene. Cancer 2011, 117, 86–95. [Google Scholar] [CrossRef] [Green Version]
  113. Yu, Y.; Nangia-Makker, P.; Farhana, L.; GRajendra, S.; Levi, E.; Majumdar, A.P.N. miR-21 and miR-145 cooperation in regulation of colon cancer stem cells. Mol. Cancer 2015, 14, 98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Sciot, R. MDM2 Amplified Sarcomas: A Literature Review. Diagnostics 2021, 11, 496. [Google Scholar] [CrossRef]
  115. Meng, X.; Franklin, D.A.; Dong, J.; Zhang, Y. MDM2-p53 pathway in hepatocellular carcinoma. Cancer Res. 2014, 74, 7161–7167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Zhang, J.; Sun, Q.; Zhang, Z.; Ge, S.; Han, Z.-G.; Chen, W.-T. Loss of microRNA-143/145 disturbs cellular growth and apoptosis of human epithelial cancers by impairing the MDM2-p53 feedback loop. Oncogene 2013, 32, 61–69. [Google Scholar] [CrossRef] [Green Version]
  117. Zhang, H.-F.; Alshareef, A.; Wu, C.; Jiao, J.-W.; Sorensen, P.H.; Lai, R.; Xu, L.-Y.; Li, E.-M. miR-200b induces cell cycle arrest and represses cell growth in esophageal squamous cell carcinoma. Carcinogenesis 2016, 37, 858–869. [Google Scholar] [CrossRef] [Green Version]
  118. Zhang, H.-F.; Zhang, K.; Liao, L.-D.; Li, L.-Y.; Du, Z.-P.; Wu, B.-L.; Wu, J.-Y.; Xu, X.-E.; Zeng, F.-M.; Chen, B.; et al. miR-200b suppresses invasiveness and modulates the cytoskeletal and adhesive machinery in esophageal squamous cell carcinoma cells via targeting Kindlin-2. Carcinogenesis 2014, 35, 292–301. [Google Scholar] [CrossRef] [Green Version]
  119. Li, T.; Hu, P.-S.; Zuo, Z.; Lin, J.-F.; Li, X.; Wu, Q.-N.; Chen, Z.-H.; Zeng, Z.-L.; Wang, F.; Zheng, J.; et al. METTL3 facilitates tumor progression via an m6A-IGF2BP2-dependent mechanism in colorectal carcinoma. Mol. Cancer 2019, 18, 112. [Google Scholar] [CrossRef] [PubMed]
  120. Sun, X.; Dai, G.; Yu, L.; Hu, Q.; Chen, J.; Guo, W. miR-143-3p inhibits the proliferation, migration and invasion in osteosarcoma by targeting FOSL2. Sci. Rep. 2018, 8, 606. [Google Scholar] [CrossRef] [Green Version]
  121. Wan, X.; Guan, S.; Hou, Y.; Qin, Y.; Zeng, H.; Yang, L.; Qiao, Y.; Liu, S.; Li, Q.; Jin, T.; et al. FOSL2 promotes VEGF-independent angiogenesis by transcriptionnally activating Wnt5a in breast cancer-associated fibroblasts. Theranostics 2021, 11, 4975–4991. [Google Scholar] [CrossRef]
  122. Yin, J.; Hu, W.; Fu, W.; Dai, L.; Jiang, Z.; Zhong, S.; Deng, B.; Zhao, J. HGF/MET Regulated Epithelial-Mesenchymal Transitions And Metastasis By FOSL2 In Non-Small Cell Lung Cancer. OncoTargets Ther. 2019, 12, 9227–9237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Gong, W.-J.; Peng, J.-B.; Yin, J.-Y.; Li, X.-P.; Zheng, W.; Xiao, L.; Tan, L.-M.; Xiao, D.; Chen, Y.-X.; Li, X.; et al. Association between well-characterized lung cancer lncRNA polymorphisms and platinum-based chemotherapy toxicity in Chinese patients with lung cancer. Acta Pharmacol. Sin. 2017, 38, 581–590. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, P.; Wei, Y.; Wang, L.; Debeb, B.G.; Yuan, Y.; Zhang, J.; Yuan, J.; Wang, M.; Chen, D.; Sun, Y.; et al. ATM-mediated stabilization of ZEB1 promotes DNA damage response and radioresistance through CHK1. Nat. Cell Biol. 2014, 16, 864–875. [Google Scholar] [CrossRef] [PubMed]
  125. Gong, W.-J.; Yin, J.-Y.; Li, X.-P.; Fang, C.; Xiao, D.; Zhang, W.; Zhou, H.-H.; Li, X.; Liu, Z.-Q. Association of well-characterized lung cancer lncRNA polymorphisms with lung cancer susceptibility and platinum-based chemotherapy response. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2016, 37, 8349–8358. [Google Scholar] [CrossRef] [PubMed]
  126. David, L.; Fernandez-Vidal, A.; Bertoli, S.; Grgurevic, S.; Lepage, B.; Deshaies, D.; Prade, N.; Cartel, M.; Larrue, C.; Sarry, J.-E.; et al. CHK1 as a therapeutic target to bypass chemoresistance in AML. Sci. Signal 2016, 9, ra90. [Google Scholar] [CrossRef]
  127. Ohba, S.; Yamashiro, K.; Hirose, Y. Inhibition of DNA Repair in Combination with Temozolomide or Dianhydrogalactiol Overcomes Temozolomide-Resistant Glioma Cells. Cancers 2021, 13, 2570. [Google Scholar] [CrossRef] [PubMed]
  128. Yang, P.; Yang, Y.; An, W.; Xu, J.; Zhang, G.; Jie, J.; Zhang, Q. The long noncoding RNA-ROR promotes the resistance of radiotherapy for human colorectal cancer cells by targeting the p53/miR-145 pathway. J. Gastroenterol. Hepatol. 2017, 32, 837–845. [Google Scholar] [CrossRef]
  129. Liu, L.-Z.; Zhou, X.-D.; Qian, G.; Shi, X.; Fang, J.; Jiang, B.-H. AKT1 amplification regulates cisplatin resistance in human lung cancer cells through the mammalian target of rapamycin/p70S6K1 pathway. Cancer Res. 2007, 67, 6325–6332. [Google Scholar] [CrossRef] [Green Version]
  130. Xu, Q.; Liu, L.-Z.; Qian, X.; Chen, Q.; Jiang, Y.; Li, D.; Lai, L.; Jiang, B.-H. MiR-145 directly targets p70S6K1 in cancer cells to inhibit tumor growth and angiogenesis. Nucleic Acids Res. 2012, 40, 761–774. [Google Scholar] [CrossRef]
  131. Fu, D.; Zhang, Y.; Cui, H. Long noncoding RNA CCAT2 is activated by E2F1 and exerts oncogenic properties by interacting with PTTG1 in pituitary adenomas. Am. J. Cancer Res. 2018, 8, 245–255. [Google Scholar]
  132. Zang, Y.; Li, J.; Wan, B.; Tai, Y.; Liu, H.; Li, Q.; Ji, Y. Long non-coding RNA CCAT2 drives the growth of laryngeal squamous cell carcinoma via regulating YAP activity. Hum. Cell 2021, 34, 1878–1887. [Google Scholar] [CrossRef]
  133. Cheng, J.; Xie, H.-Y.; Xu, X.; Wu, J.; Wei, X.; Su, R.; Zhang, W.; Lv, Z.; Zheng, S.; Zhou, L. NDRG1 as a biomarker for metastasis, recurrence and of poor prognosis in hepatocellular carcinoma. Cancer Lett. 2011, 310, 35–45. [Google Scholar] [CrossRef]
  134. Yu, Z.-Y.; Wang, Z.; Lee, K.-Y.; Yuan, P.; Ding, J. Effect of silencing colon cancer-associated transcript 2 on the proliferation, apoptosis and autophagy of gastric cancer BGC-823 cells. Oncol. Lett. 2018, 15, 3127–3132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Hayashi, H.; Arao, T.; Togashi, Y.; Kato, H.; Fujita, Y.; De Velasco, M.A.; Kimura, H.; Matsumoto, K.; Tanaka, K.; Okamoto, I.; et al. The OCT4 pseudogene POU5F1B is amplified and promotes an aggressive phenotype in gastric cancer. Oncogene 2015, 34, 199–208. [Google Scholar] [CrossRef]
  136. Tian, G.-W.; Li, N.; Xin, Y. Prognostic and clinicopathological significance of CCAT2 in Chinese patients with various tumors. Int. J. Biol. Markers 2017, 32, e344–e351. [Google Scholar] [CrossRef]
  137. Wang, D.; Chen, Z.; Xu, H.; He, A.; Liu, Y.; Huang, W. Long noncoding RNA CCAT2 as a novel biomaker of metastasis and prognosis in human cancer: A meta-analysis. Oncotarget 2017, 8, 75664–75674. [Google Scholar] [CrossRef]
  138. Fan, Y.-H.; Fang, H.; Ji, C.-X.; Xie, H.; Xiao, B.; Zhu, X.-G. Long noncoding RNA CCAT2 can predict metastasis and poor prognosis: A meta-analysis. Clin. Chim. Acta 2017, 466, 120–126. [Google Scholar] [CrossRef] [PubMed]
  139. Shaker, O.G.; Senousy, M.A.; Elbaz, E.M. Association of rs6983267 at 8q24, HULC rs7763881 polymorphisms and serum lncRNAs CCAT2 and HULC with colorectal cancer in Egyptian patients. Sci. Rep. 2017, 7, 16246. [Google Scholar] [CrossRef] [PubMed]
  140. Fosselteder, J.; Calin, G.A.; Pichler, M. Long non-coding RNA CCAT2 as a therapeutic target in colorectal cancer. Expert Opin. Ther. Targets 2018, 22, 973–976. [Google Scholar] [CrossRef]
Figure 1. Overall schematic of the main tumor-promoting regulatory mechanisms of CCAT2 described to date.
Figure 1. Overall schematic of the main tumor-promoting regulatory mechanisms of CCAT2 described to date.
Ijms 22 12491 g001
Figure 2. Currently known interactions of CCAT2 and other ncRNAs. The figure presents the effect of CCAT2 sponging on tumor suppressor miRNAs and how it influences essential cancer processes.
Figure 2. Currently known interactions of CCAT2 and other ncRNAs. The figure presents the effect of CCAT2 sponging on tumor suppressor miRNAs and how it influences essential cancer processes.
Ijms 22 12491 g002
Table 1. Currently known regulatory targets and biological effects of different expression patterns of CCAT2 in various cancer types.
Table 1. Currently known regulatory targets and biological effects of different expression patterns of CCAT2 in various cancer types.
Cancer TypeStudy SamplesCCAT2 ExpressionRegulatory TargetsBiological EffectRef.
Colorectal cancerTissue: 191 CRC and ANTT
Cell lines: COLO320D, MHCT116, RKO, HEK293
↑Wnt/β-catenin
↑MYC, ↑TCF7L2
↑ MS status
Invasion, Distant Metastasis[16]
Cell lines: HCT116, KM12C,
KM12SM, COLO320, DLD-1, HT29
In vivo model: CCAT2 transgenic mice, WT mice
↑BOP1
↑AURK B
Chromosomal instability, Chemoresistance to 5 fluorouracil and oxaliplatin[32]
Tissue: 218 CRC and ANTT Differentiation grade
TNM stage, Lymph nodes metastasis, Distant metastasis, Vascular invasion, Poor prognosis
[47]
Tissue: 149 CRC and ANTT Distant metastasis[48]
Cell lines: HCT-116, HT-29↓ pre-miR-145,
↑ miR-21
Proliferation, Invasion[49]
Tissue: 280 CRC and ANTT↑MS status
↑ MYC
Poor prognosis, Lymph nodes metastasis, TNM stage[50]
Blood: 63 CRC and 40 ControlsN/A--[51]
Tissue: 60 CRC, 30 Colon polyps, and 60 non-cancers. ↑ along with CCAT1, CCAT2,
H19, HOTAIR, HULC, MALAT1, PCAT1, MEG3, PTENP1, and TUSC7
Part of a stool lncRNA panel for CRC detection [52]
Tissue: 80 CRC and ANTT
Cell lines: FHC, HT29, Lovo, HCT-116.
-Cellular growth, Proliferation, Antiapoptotic[53]
Tissue: 150 CRC and ANTT↑MYCMetastasis[54]
Esophageal cancerTissue: 229 ESCC and ANTT-Poor prognosis, Lymph node metastasis, TNM stage[55]
Tissue: 57 ESCC and ANTT
Cell lines: TE13, KYSE410, ECA109, TE1/N: HEEC
↑TE1, TE13, KYSE410
↓ ECA109
-CCAT2 expression correlated with smoking status [56]
Cell lines: Eca-109, EC9706, KYSE150, TE-1/N: HEEC ↑BCL-2, ↓BAX, ↓CYCLIN D1, ↑Wnt pathwayProliferation, Migration, Invasion[57]
Tissue: 62 OSCC and ANTT
Cell lines: Tca8113, Cal27/hNOK
↑CCND1, ↑MYC,
↑Wnt/β-catenin
Poor prognosis, Invasion
proliferation, T stage differentiation
[58]
Tissue: 60 ESCC and 21 esophageal mucosa
Cell lines: HEEC, TE-1, TE-3, ECA109, KYSE410, KYSE520
↓miR-145, ↑p70S6K1,
↓p53 pathway
Radiotherapy resistance, cellular proliferation.[59]
Tissue: 33 ESCC and ANTT
Cell lines: KYSE-410, KYSE-150,
TE10, TE11, TE13/HET-1A
β -catenin/
WISP1 signaling pathway
Cell proliferation, Invasion, Poor prognosis. [26]
Tissue: 93 ESCC and ANTT
Cell lines: Eca109, TE-1, EC-1, ESC410/HET-1A
↓miR-200b ↑IGF2BP2/TK1 AxisMigration, Invasion, Tumorigenesis[60]
Gastric cancerTissue: 85 GC and ANTT-Lymph node metastasis
Distant metastasis, Poor prognosis
[61]
Tissue: 108 GC and ANTT
Cell lines: Tu: SGC7901,
MKN45, BGC-823, MKN-28/N: GES-1
ZEB2, ↑VIM, ↑CHD1,
CHD2
EZH2, ↑3K27me3 ↑LSD1, ↑LATS
Poor prognosis Proliferation
Migration, Invasion, EMT
[62]
Tissue: 60 GC and ANTT
Cell lines: GES-1, RGM-1, SGC-7901, SNU-1, HGC-27
↑mTOR signalingProliferation, Metastasis[63]
Serum: 167 GC and 110 controls-Tumor stage, Invasion, Lymph node metastasis[64]
Hepatocellular carcinomaTissue: 50 HCC and ANTT
Cell lines: Tu: HepG2, HEP3B, HCCLM3, HuH7/L02
-Proliferation, Migration
Antiapoptotic
[65]
Tissue: 60 HCC and ANTT
Cell lines: SMMC-7721, PLC/PRF/5, Huh7, SK-Hep-1, Hep3B
↑FOXM1
↓ miR-34a
Poor prognosis, Proliferation, Tumor growth, Antiapoptotic[66]
Tissue: 96 HCC and ANTT
Cell lines: HepG2, SMMC772, MHCC97H /MIHA
↑CDH1,
↑CDH2,
↑VIM, ↑SNAI2
Poor prognosis, TNM stage, Vascular invasion, Alcoholism history, Migration, Invasion, EMT[67]
Cell lines: SMMC7721, SK-hep1, HepG2, Huh7/L02↑NDRG1 promoterProliferation, Migration, Invasion [68]
Cell lines: Hep3B, HepG2, and THLE-3, MHCC97H↓miR-145
MDM2
Proliferation, Metastasis [45]
Tissue: 61 HCC and ANTT
Cell lines: HepG2 HCCLM3
↓miR- 4496
↑ELAVL1
Advanced stage, Venous invasion, Migration, Invasion[69]
Pancreatic cancerTissue: 80 PDAC and ANTT
Cell lines: PANC-1, SW1990, PC-3/HPDE6-C7
↑KRAS,
↑MEK/ERK
Poor prognosis, Proliferation, Invasion
Tumor growth
[70]
GliomaCell lines: A172, U87-MG, U251, T98G/HUVECs ↑VEGF, ↑TGFβ, ↑FGFAngiogenesis, Migration
Proliferation
[71]
Tissue:134 Glioma and ANTT
Cell lines: U87-MG, U251
↑Wnt/β-cateninTNM stage, Proliferation
Cell cycle, Migration
Tumor growth
[72]
Tissue: 134 Glial tumors and ANTT
Cell lines: U87, U251, A172, SHG44/Normal human astrocyte cell line
↑CDH1, ↑CDH2, ↑VIM, ↑TWIST, ↑ SNAI1Poor prognosis, Tumor grade, Tumor size, Proliferation, Migration, Invasion, Apoptosis, EMT[73]
Tissue: 74 PA and ANTT
Cell line: HP75
↑E2F1
↑PTTG1
Poor prognosis, Proliferation, Antiapoptotic, Cell cycle, Migration, Invasion[74]
Tissue: 138 Gliomas and ANTT
Cell lines: U251, U87, A172, SHG44.
↓ miR-424
↑CHK1
Proliferation, Invasion, Migration via miR-424 sponging and CHK1 regulation[44]
Cell lines: A172, U251↓ miR-424
↑ VEGFA
Proliferation, Migration, Angiogenesis[43]
NeuroblastomaTissue: 96 Neuroblastomas and ANTT
Cell lines: SH-SY5Y, SK-N-SH/HUVEC
↓P53
↑BCL-2
Antiapoptotic, Cell growth, Poor prognosis[75]
Lung cancerTissue: 57 NSCLC and ANTT
cell lines: A549, NCI-H1975, NCI-H358, NCI-H1650, NCI-H1299, SK-MES-1, Pc-9/HBE

↑ H1975, Pc9, NCI-H358
↓NCI-H1299 NCI-H1650, A549 SK-MES-1
Proliferation, Invasion[76]
Tissue: 112 SCLC and ANTT
Cell lines: DMS-53, H446/16 HBE
Poor progression, Clinical stage, Tumor size, Distant metastasis
Proliferation, Invasion
[77]
Tissue: 36 NSCLC and ANTT
Cell lines: NCI-H1975
↑Wnt/β-cateninTumor size, Lymph node metastasis[78]
Cell lines:A549, SPC-A- 1, H1395, H441, H1975/BEAS-2B↑FOXC1
↓ miR-23b-5p
Proliferation, Migration[42]
Tissue: 32 NSCLC and ANTTN/A--[79]
Serum: 438 LC and 438 controls--[80]
OsteosarcomaTissue: 50 OS and ANTT
Cell lines: SAOS-2, MG63, U2-OS/Normal osteoblast cell line
↑GSK3β/β-catenin Tumor size, Poor prognosis, Proliferation[81]
Tissue: 40 OS/ANTT
Cell lines: SOSP-9607, MG-63, U2OS, SAOS-2/hFOB
↑LATS2, ↑MYC ↑CDH1, ↑CHD2, ↑SNAI1Poor prognosis, Proliferation.
EMT
[82]
Cell lines: SOSP-9607, MG-63, U2OS, SAOS-2 /hFOB↓miR-143, ↑FOSL2Proliferation, Metastasis[83]
Thyroid cancerTissues: 30 pairs TC and ANTT
(papillary, follicular, and anaplastic)
Cell lines: TPC- 1, TH83, IHH4, FTC- 133, FTC- 238/Nthy-ori3-1
↑Wnt/β-cateninProliferation, Migration, Invasion, Apoptosis [84]
Tissue: 60 anaplastic and papillary TC and ANTT
Cell lines: TC cell lines
-Doxorubicin and cisplatin resistance, Increased tumor size, Poor prognosis,[85]
Multiple myelomaSerum: 106 MM and 106 matched normal controls-ISS stages, Renal dysfunction, Serum creatinine[86]
Acute myeloid leukemiaBone marrow samples: 46 patients and 46 healthy volunteers
Cell lines: KG-1
Cell cycle arrest in S phaseCellular proliferation, Poor prognosis[87]
Breast cancerTissue: 997 BC and ANTT and 56 BC and ANTT
Cell lines: MDA-MB-231, MDA-MB-436
-Poor prognosis, Therapeutic response[88]
Tissue: 67 BC and ANTT
Cell lines: MDA-MB-231, MCF-7/Hs578Bst
↑Wnt/β-catenin
CCND1
↑MYC
Poor prognosis, Proliferation, Invasion,
Tumorigenesis
[30]
Cell lines: MCF-7, T47 D tamoxifen resistant/MCF-7, T47D–tamoxifen responsive↑ tamoxifen-resistant cell lines Suppressing CCAT2 expression improves sensitivity to tamoxifen in resistant cells[89]
Tissue: 48 BC and ANTT-Lymph node metastasis[90]
Tissue: 67 BC and ANTT
Cell lines: MDA-MB-231, MCF-7/MCF10A
↑P15
↑EZH2
Poor prognosis, Proliferation, Invasion,
Cell cycle, Tumor growth
[91]
Tissue: 60 BC and ANTT
Cell lines: LCC9, MDA-MB-231 MCF-7/HCC1937
↑TGF-β, ↑Smad2, ↑α-SMALymph node metastasis
Proliferation, Invasion, Migration, Apoptosis, Cell cycle
[92]
Endometrial cancerTissue: 30 EC and ANTT
Cell lines: HEC-1-A and RL95-2
↓miR-216b
↑PI3K/AKT
Proliferation, Migration,
Invasion, Apoptosis
[41]
Cervical cancerTissue: 123 SCCC and ANTT-FIGO stage, Lymph node metastasis, Cervical invasion, Poor prognosis[93]
Cell lines: CaSki, HeLa, SiHa-Proliferation, Apoptosis[94]
Serum: 115 SCCC, 79 CIN, and 110 healthy controlsCCAT2, LINC01133, LINC00511 upregulated in serum of SCCC and CIN patients. [95]
Tissue: 30 SCCC and ANTT
Cell lines: GH329, CaSki, HeLa, SiHa, C4-1.
Xenografts: 2 groups pSilencer, pSilencer/sh-CCAT2
↓miR-493-5p
↑CREB1
EMT, Proliferation[96]
Ovarian CancerTissue: 31 EOC and ANTT
Cell lines: SKOV3, MC685, A2780, HO8910/IOSE 386
↓miR-424Proliferation, Apoptosis
[39]
Cell lines: SKOV3, A2780, HO8910/HOSE, HUM-CELL-0088↑CDH1, ↑CHD2, ↑SNAI1, ↑SNAI2, ↑TWIST1, ↑Wnt/β-catenin Migration, Invasion,
EMT
[28]
Tissue: 109 EOC and ANTT
Cell lines: SKOV3, IGROV1, A2780, OVCAR3/HOSE 6.3
-FIGO stage, Tumor grade, Distant metastasis, Poor prognosis, Proliferation
Migration, Invasion
[97]
Cell lines: SKOV3 and A2780↑TCF7L2, ↑MYCVitamin D suppresses CCAT2 expression[98]
Prostate cancerTissue: 96 PC and ANTT
Cell lines: DU-145, 22RV1/WPMY-1
-Poor prognosis, Proliferation, Migration
Invasion, EMT
[99]
Renal cell cancerTissue: 61ccRCC and ANTT
Cell lines: 786-O, AHCN ccRCC/HK-2
↑Wnt/β-cateninPoor prognosis, Proliferation, Migration,
Apoptosis, Invasion
[29]
Bladder cancerTissue: 48 BC and ANTT
Cell lines: SV-HUC-1/T24, 5637
-Tumor grade, TNM stage, Proliferation, Migration, Apoptosis[100]
AML, acute myeloid leukemia; ANTT, adjacent non-tumor tissue; BC, bladder cancer; CC, cervical cancer; ccRCC-clear cell renal cell carcinoma; CRC, colorectal cancer; EC, endometrial cancer; HCC- hepatocellular carcinoma; MS status, microsatellite status; EMT, epithelial–mesenchymal transition; GC, gastric cancer; EOC, epithelial ovarian cancer; ESCC, esophageal squamous cellular carcinoma; MM, multiple myeloma; PC, prostate cancer; PDAC, pancreatic ductal adenocarcinoma; TC, thyroid carcinoma.
Table 2. CCAT2 regulatory roles on therapeutic resistance in human cancers.
Table 2. CCAT2 regulatory roles on therapeutic resistance in human cancers.
Cancer TypeRole of CCAT2 in Therapeutic ResistanceRef.
Thyroid cancerUpregulation is associated with chemoresistance to doxorubicin and cisplatin.[85]
Colorectal cancerUpregulation is associated with chromosomal instability and chemotherapy resistance to 5-fluorouracil and oxaliplatin.[32]
Lung cancerPresence of the rs6983267 SNP was associated with reduced hematological toxicity to platinum-based chemotherapy and platinum-based chemotherapy response.[123,125]
Breast cancerUpregulation enhances tamoxifen resistance in breast cancer cell lines.[89]
GlioblastomaUpregulation in glioblastoma cell lines increases resistance to teniposide, temozolomide, vincristine, and cisplatin.[44]
Esophageal squamous cell carcinomaUpregulation promotes radiotherapy resistance in ESCC cell lines by inhibiting miR-145, the expression level of P70 ribosomal protein S6 kinase 1, p53, and p21.[59]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pirlog, R.; Drula, R.; Nutu, A.; Calin, G.A.; Berindan-Neagoe, I. The Roles of the Colon Cancer Associated Transcript 2 (CCAT2) Long Non-Coding RNA in Cancer: A Comprehensive Characterization of the Tumorigenic and Molecular Functions. Int. J. Mol. Sci. 2021, 22, 12491. https://doi.org/10.3390/ijms222212491

AMA Style

Pirlog R, Drula R, Nutu A, Calin GA, Berindan-Neagoe I. The Roles of the Colon Cancer Associated Transcript 2 (CCAT2) Long Non-Coding RNA in Cancer: A Comprehensive Characterization of the Tumorigenic and Molecular Functions. International Journal of Molecular Sciences. 2021; 22(22):12491. https://doi.org/10.3390/ijms222212491

Chicago/Turabian Style

Pirlog, Radu, Rares Drula, Andreea Nutu, George Adrian Calin, and Ioana Berindan-Neagoe. 2021. "The Roles of the Colon Cancer Associated Transcript 2 (CCAT2) Long Non-Coding RNA in Cancer: A Comprehensive Characterization of the Tumorigenic and Molecular Functions" International Journal of Molecular Sciences 22, no. 22: 12491. https://doi.org/10.3390/ijms222212491

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