Next Article in Journal
Melatonin-Mediated Molecular Responses in Plants: Enhancing Stress Tolerance and Mitigating Environmental Challenges in Cereal Crop Production
Previous Article in Journal
Molecular Dynamics Simulations of Different Nanoparticles at Substrates
Previous Article in Special Issue
Synthesis, Characterization, Cytotoxicity, Cellular Imaging, Molecular Docking, and ADMET Studies of Piperazine-Linked 1,8-Naphthalimide-Arylsulfonyl Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring the Key Signaling Pathways and ncRNAs in Colorectal Cancer

by
Yun Ju Lee
1,2,
Woo Ryung Kim
1,2,
Eun Gyung Park
1,2,
Du Hyeong Lee
1,2,
Jung-min Kim
1,2,
Hae Jin Shin
1,2,
Hyeon-su Jeong
1,2,
Hyun-Young Roh
2,3 and
Heui-Soo Kim
2,3,*
1
Department of Integrated Biological Sciences, Pusan National University, Busan 46241, Republic of Korea
2
Institute of Systems Biology, Pusan National University, Busan 46241, Republic of Korea
3
Department of Biological Sciences, College of Natural Sciences, Pusan National University, Busan 46241, Republic of Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(8), 4548; https://doi.org/10.3390/ijms25084548
Submission received: 29 March 2024 / Revised: 19 April 2024 / Accepted: 19 April 2024 / Published: 21 April 2024
(This article belongs to the Special Issue Functional Molecules in Tracing and Cancer Therapeutics)

Abstract

:
Colorectal cancer (CRC) is the third most prevalent cancer to be diagnosed, and it has a substantial mortality rate. Despite numerous studies being conducted on CRC, it remains a significant health concern. The disease-free survival rates notably decrease as CRC progresses, emphasizing the urgency for effective diagnostic and therapeutic approaches. CRC development is caused by environmental factors, which mostly lead to the disruption of signaling pathways. Among these pathways, the Wingless/Integrated (Wnt) signaling pathway, Phosphatidylinositol 3-kinase/protein kinase B/mammalian target of rapamycin (PI3K/AKT/mTOR) signaling pathway, Mitogen-Activated Protein Kinase (MAPK) signaling pathway, Transforming Growth Factor-β (TGF-β) signaling pathway, and p53 signaling pathway are considered to be important. These signaling pathways are also regulated by non-coding RNAs (ncRNAs), including microRNAs (miRNAs), long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs). They have emerged as crucial regulators of gene expression in CRC by changing their expression levels. The altered expression patterns of these ncRNAs have been implicated in CRC progression and development, suggesting their potential as diagnostic and therapeutic targets. This review provides an overview of the five key signaling pathways and regulation of ncRNAs involved in CRC pathogenesis that are studied to identify promising avenues for diagnosis and treatment strategies.

1. Introduction

Colorectal cancer (CRC) is the third most commonly diagnosed cancer in both males and females, with an estimated 81,540 new cases in males and 71,720 new cases in females in 2024, according to the American Cancer Society [1]. In addition, CRC is estimated to cause significant mortality, with approximately 28,700 deaths in males, ranking the top three among males, and 24,310 deaths in females, positioning it as the fourth among females. An examination of the five-year disease-free survival rate revealed a notable trend: as the stage of CRC advances, the survival rates decrease. Specifically, the five-year disease-free survival rates are 91.0% for stage I, 79.8% for stage II, 63.3% for stage III, and 18.9% for stage IV [2]. These statistics highlight the severity of this disease and underscore the need for effective strategies for diagnosing and treating CRC.
CRC develops as a result of abnormal growth cells in the colon or rectum, which over time turn cancerous [3]. It is commonly known that obesity, overweight, a lack of physical activity, smoking, alcohol consumption, and diet are risk factors for CRC [4,5,6]. Most CRCs are thought to be caused by genetic alterations induced by these environmental factors [7]. Consequently, the genetic alterations can induce the aberrant activation of signaling pathways, leading to tumor progression [8]. Many dysregulated signaling pathways in CRC have been identified, most notably the Wingless/Integrated (Wnt) signaling pathway, Phosphatidylinositol 3-kinase/protein kinase B/mammalian target of rapamycin (PI3K/AKT/mTOR) signaling pathway, Mitogen-Activated Protein Kinase (MAPK) signaling pathway, Transforming Growth Factor-β (TGF-β) signaling pathway, and p53 signaling pathway [9]. Aberrant signaling pathways occur due to abnormalities in the expression of genes that function in the pathway, which also affects genes downstream [8]. As these studies continue to expand, there is increasing interest in non-coding RNAs (ncRNAs) as regulators of gene expression.
Unlike coding RNAs, ncRNAs, which are transcribed from DNA but not translated into protein, have been shown to act as a regulator [10]. Well-known ncRNAs such as microRNAs (miRNAs), long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs) have been extensively studied and shown to directly or indirectly regulate the expression of genes involved in CRC, thereby influencing various biological processes [11]. Due to their altered expression patterns and their significant function in CRC, these ncRNAs can also be used not only for the diagnosis but also as potential targets for the treatment. In this review, we aim to summarize the five major signaling pathways and ncRNAs implicated in CRC, to identify potential targets for diagnosis and treatment.

2. Multiple Major Signaling Pathways in CRC

In the human body, various signaling pathways exist, regulating essential cellular processes. However, when these pathways are disrupted, they can lead to various diseases [12]. CRC is no exception, and through numerous studies, differences in signaling pathways between normal and CRC patients have been extensively identified [8]. By comparing these signaling pathways with KEGG pathways (map05210), five signaling pathways play crucial roles in influencing CRC: the Wnt signaling pathway, PI3K/AKT/mTOR signaling pathway, MAPK signaling pathway, TGF-β signaling pathway, and p53 signaling pathway [9,13]. Although these signaling pathways ultimately regulate similar cellular characteristics, there are differences in the genes involved and the factors that trigger the activation of each gene within these pathways (Figure 1). Therefore, in this section, we focus on the changes in gene expression associated with each of these signaling pathways and discuss how these factors affect CRC.

2.1. Wnt Signaling Pathway

The complex and highly conserved Wnt signaling pathway is essential for embryonic development, tissue homeostasis, and various physiological processes [14,15,16]. The presence or absence of Wnt, a ligand, influences the expression levels of genes, particularly the transcriptional co-activator β-catenin, associated with the Wnt signaling pathway. In the absence of Wnt, β-catenin is phosphorylated by the Axin/GSK3/APC/CK1 complex and degraded by the ubiquitin-dependent proteasome, maintaining it at a low level [17]. However, in the presence of Wnt, Wnt binds to the Frizzled receptor, activating the Wnt signaling pathway [14,18,19]. The Frizzled receptor dimerizes with LRP5/6 and recruits the Axin/GSK3/APC/CK1 complex to the cell membrane, preventing the phosphorylation and degradation of β-catenin. Accumulated cytoplasmic β-catenin then translocates to the nucleus, where it interacts with TCF/LEF transcription factors, inducing the expression of Wnt target genes.
The Wnt signaling pathway has been extensively studied to prove its critical role in CRC, and the hyperactivation of this signaling pathway has been observed from early to advanced stages of CRC [20,21,22]. In this signaling pathway, β-catenin and adenomatous polyposis coli (APC) especially have been identified as representative core genes with frequent genetic changes in CRC, which have been found to significantly impact the development and progression of the disease [23,24,25]. According to a study, it was confirmed that β-catenin was activated by Wnt1, which plays an important role in CRC progression, and that a high expression of nuclear β-catenin was associated with poor prognosis [26]. Similarly, it was revealed that the activation of β-catenin in CRC contributed to the aggressiveness of CRC by manipulating cell-cell junctions and inducing epithelial-mesenchymal transition (EMT) [27]. In a different study, it was found that β-catenin was activated by HMGB3, leading to the activation of downstream components of the Wnt signaling pathway, including c-myc and MMP7 [28]. This activation promoted cell growth and migration, influencing the carcinogenesis and development of CRC. Unlike β-catenin, APC is known to act as a negative regulator in the Wnt signaling pathway and is recognized as a tumor suppressor [29]. In one study, it was observed that the overexpression of SMYD2, a negative regulator of APC2, was associated with poor prognosis [30]. The knockdown of SMYD2 led to a decrease in cell proliferation, migration, and invasion, along with an upregulation of APC2 and a downregulation of the expression of E-cadherin, inducing EMT. Additionally, inhibiting SMYD2 reduced lung metastasis, confirming that SMYD2 plays an important role in CRC by regulating the expression of APC2 and influencing the Wnt signaling pathway.

2.2. PI3K/AKT/mTOR Signaling Pathway

The PI3K/AKT/mTOR signaling pathway is an important intracellular signaling cascade that is essential for controlling many cellular functions, including proliferation, survival, DNA repair, apoptosis, and gene transcription [12,31,32]. PI3K is known as a conserved lipid kinase family, divided into three classes based on their structure and specific substrates. While classes II and III remain relatively less-explored areas, class I has been extensively researched. Class I PI3K can be further subdivided into classes IA and IB, depending on the types of catalytic and regulatory subunits. Class IA consists of catalytic subunits, namely p110 (p110α, p110β, and p110 δ), and regulatory subunits p85 (p85 α, p55α, p50α, p85β, and p55γ) [33,34,35]. On the other hand, class IB PI3K is composed of the catalytic subunit p110γ and the regulatory subunit p101. PI3K can be activated downstream of numerous growth factors, including the fibroblast growth factor, vascular endothelial growth factor, and receptors, such as receptor tyrosine kinases [36]. Activated PI3K phosphorylates phosphatidylinositol 4,5-bisphosphate to convert it into phosphatidylinositol 3,4,5-trisphosphate; subsequently, phosphatidylinositol 3,4,5-trisphosphate activates AKT [33,36]. Activated AKT induces the phosphorylation of other substrates in both the cytosol and nucleus, particularly promoting the activation of mTOR, thereby activating the PI3K/AKT/mTOR signaling pathway. In particular, this signaling pathway plays an important role in the early and late stages of CRC, and dysregulation of this signaling pathway leads to carcinogenesis by altering proliferation, angiogenesis, and apoptosis [37,38,39].
One study has been conducted to investigate the correlation between the expression of PI3K and primary versus metastatic CRC [40]. It was found that metastatic lesions in CRC patients exhibited a higher expression of PI3K compared to primary lesions, suggesting a potential association with metastasis. To examine the actual impact of PI3K on CRC, the PI3K inhibitor LY294002 was used and resulted in the induction of apoptosis in CRC cell lines, indicating the functional significance of PI3K. According to another study, the overexpression of AKT in CRC cell lines was shown to increase their proliferative, migratory, and invasive capacities [41]. Moreover, in a xenograft mouse model of CRC, tumors overexpressing AKT exhibited rapid growth, increased vascularization, and the induction of expression of EMT markers. In a different study, it was observed that a higher expression of mTOR in CRC patients correlated with decreased overall survival and disease-free survival probabilities [42]. Additionally, the treatment of CRC cell lines with mTOR inhibitors, rapamycin, and PP242 resulted in a significant decrease in the number and diameter of spheroid cell models derived from CRC cells.
Besides the mentioned studies, there is extensive research on how changes in the expression of genes involved in the PI3K/AKT/mTOR signaling pathway impact CRC. In a study, the artificial overexpression of LACTB, which is downregulated in both CRC patients and cell lines, led to a decrease in PIK3R3 expression and a reduction in the expression of proteins associated with the PI3K/AKT/mTOR signaling pathway [43]. Furthermore, downregulated PIK3R3 increased autophagy and decreased the proliferation, EMT, and tumor size in CRC cell lines or mouse models. Taken together, these results suggest that LACTB influences CRC development by regulating the PI3K/AKT/mTOR signaling pathway. In another study, it was observed that TGM3 was downregulated in CRC patients and cell lines [44]. Increasing the expression of this gene was found to decrease the phosphorylation of PI3K and AKT, indicating the inactivation of the PI3K/AKT signaling pathway suppressed the proliferation and metastasis of CRC cells. Similarly, it has been revealed that MUC3A, which is overexpressed in CRC and associated with poor prognosis, activates the PI3K/AKT/mTOR signaling pathway [45]. The knockout of MUC3A resulted in the inhibition of the PI3K/AKT/mTOR signaling pathway and induced proliferation, cell cycle arrest, and the invasion of CRC cells.

2.3. MAPK Signaling Pathway

The MAPK signaling pathway is an essential intracellular signaling system that controls multiple cellular processes, such as cell proliferation, differentiation, survival, and apoptosis [46,47]. This signaling pathway is initiated when a ligand binds to receptor tyrosine kinases, leading to the phosphorylation and activation of the MAPK kinase kinase (MAP3K) [48,49]. Activated MAP3K phosphorylates MAPK kinase (MAP2K), which, in turn, phosphorylates MAPK, activating various kinases, enzymes, and transcription factors, thereby transmitting signals from cell surface receptors to the nucleus and impacting gene expression.
In the case of MAPK, it consists of three main families: extracellular signal-regulated kinase 1/2 (ERK1/2), c-Jun N-terminal kinase (JNK), and p38. These families are identified to play important roles in the development, progression, and oncogenic impact of CRC, acting individually or concurrently within the specific cellular context and signaling requirements in CRC [5,50]. There are several studies on the independent action of MAPK in CRC. One study revealed the influence of the ERK/MAPK signaling pathway on the progression of CRC by demonstrating a decrease in CRC cell proliferation and migration when the MAP2K inhibitor U0126 was used to block the ERK/MAPK signaling pathway [51]. Similarly, another study revealed that the knockdown of COPB2, which was overexpressed in CRC, led to a reduction in the proliferation and induction of apoptosis in CRC cells through the JNK signaling pathway [52]. In a different study, activated FOXC1, whose degradation was inhibited by phosphorylation through p38, was overexpressed in CRC patients and correlated with poor prognosis [53]. The knockdown of FOXC1 induced a decreased migration and invasion of CRC cells. Additionally, it reduced the number of metastatic tumors in the CRC mouse model.
In addition to cases where a single family independently affects the MAPK signaling pathway, it has been revealed that two MAPK families show combined actions within the signaling pathway. For instance, investigating both JNK and p38 MAPK signaling pathways simultaneously, it was found that S100A16 was overexpressed in CRC patients and correlated with poor prognosis [54]. The overexpression of S100A16 led to an increase in the expression of mesenchymal markers N-cadherin and vimentin, while decreasing the expression of the epithelial marker E-cadherin, implicating its involvement in EMT. Furthermore, treatment with the JNK inhibitor SP600125 and the p38 inhibitor SB203580 resulted in a decrease in cell migration and invasion. This suggests that S100A16 influences CRC through the JNK and p38 MAPK signaling pathways. Similarly, it has been confirmed that both ERK1/2 and p38 MAPK signaling pathways simultaneously influence CRC. The overexpression of CD24 induced the proliferation of CRC cells and increased the tumor size in a CRC mouse model, promoting tumorigenicity [55]. Additionally, increased CD24 promoted the activation of ERK1/2, Raf-1, and p38 MAPK, confirming that CD24-dependent ERK1/2 and p38 MAPK activation were necessary for CRC proliferation.

2.4. TGF-β Signaling Pathway

The TGF-β signaling pathway is a complex and multifunctional pathway that plays a crucial role in the regulation of cell growth, differentiation, immune responses, and homeostasis [56,57]. This pathway is activated by the binding of TGF-β (TGF-β I, TGF-β II, and TGF-β III) to its receptor, TGF-βR (TGF-βR I, TGF-βR II, and TGF-βR III). TGF-β I and TGF-β III bind directly to TGF-βR II, while TGF-β II can interact with TGF- βR II with the assistance of TGF-βR III [56,58,59,60]. Subsequently, TGF-βR II recruits and phosphorylates TGF-βR I for activation. Then, receptor-regulated SMADs (R-SMADs, SMAD 1, SMAD 2, SMAD 3, SMAD 5, and SMAD 8) are phosphorylated and activated to initiate the signaling cascade. In turn, the common-mediator SMAD (Co-SMAD, SMAD 4) interacts with R-SMAD to form a trimeric complex. This complex translocates into the nucleus and binds to DNA binding sites as transcription factors, fulfilling their role in gene expression. The activated TGF-β signaling pathway influences tumor progression and development in CRC [56].
A study confirmed that the expression of SMAD 1 was increased in CRC patient tissues [61]. When the expression of SMAD 1 was inhibited, the migration ability of CRC and the expression of N-cadherin were decreased, while the expression of E-cadherin was increased, indicating a reduction in EMT ability. Therefore, it can be anticipated that SMAD1 induces the migration of CRC cells and increases EMT in CRC patients. Another study confirmed the influence of ITGΒ5 on CRC by regulating the TGF-β signaling pathway [62]. ITGΒ5 was overexpressed in CRC patients, which correlated with poor survival. The knockdown of ITGB5 resulted in the decreased proliferation and invasion of CRC cells. Additionally, in a CRC mouse model, the suppression of ITGΒ5 led to reduced tumor size and metastasis. In addition, the repression of ITGΒ5 resulted in the decreased expression of TGF-βR I and phospho-SMAD 2, while increasing the expression of E-cadherin. This indicates that ITGΒ5, which was highly expressed in CRC, promotes the TGF-β signaling pathway and induces EMT in CRC. Similarly, it was revealed that the transcription activator EHF was overexpressed in CRC and correlated positively with poor prognosis [63]. EHF knockdown suppressed CRC cell proliferation, viability, and migration while inducing apoptosis. Moreover, in the CRC mouse model, repressed EHF resulted in reduced tumor size and metastasis to the liver. Further experiments revealed a positive correlation between the expressions of EHF and TGF-β1. It was demonstrated that EHF binds to the promoter of TGF-β1, thereby regulating its expression and implicating EHF in the tumorigenesis and metastasis of CRC.

2.5. p53 Signaling Pathway

The p53 signaling pathway is a crucial cellular pathway that plays a central role in regulating cellular responses to stress, including cell cycle, DNA repair, and apoptosis [64,65]. The p53 transcription factor is known to function as a tumor suppressor, aiding in the maintenance of genomic stability [6,66,67]. In unstressed cells, polyubiquitination by the E3 ubiquitin ligase mouse double minute 2 homolog (MDM2) leads to low levels of p53. MDM4 (MDMX), known as both a homology of MDM2 and a negative regulator of p53, interacts with MDM2 to bind to p53. As a result of this interaction, p53 is degraded via MDM2-dependent polyubiquitination, leading to the suppression of its transcriptional activity [6,68]. However, various stress signals, including DNA damage and abnormal growth signals, disrupt the interaction between p53 and MDM2. This disruption results in the stabilization of p53, enabling the activation of the transcription of target genes, which regulate various cellular responses, such as senescence and metabolism [69].
In CRC, mutations in the TP53 gene, encoding p53, occur frequently, with approximately 43% of cases exhibiting mutations [6]. Mutations in p53 typically impair the function of wild-type p53, leading to a loss of its tumor suppressor properties. These alterations can affect cancer development and progression, such as cancer stem cell formation, cell proliferation, invasion, and metastasis [6,70]. One study found that CRC patients with TP53 mutations with a complete loss of p53 had significantly different tumor sizes and decreased overall survival compared to CRC patients with wild-type TP53 [71]. In another study, similar findings were observed where CRC patients with a positive p53 expression had higher overall survival compared to those with a negative p53 expression [72].
In addition to the direct impact of mutations in the TP53 on CRC, it has been revealed that the expression level of other genes located upstream or downstream of the p53 affects CRC by influencing the p53 signaling pathway. One study confirmed that the p53 signaling pathway influenced cancer progression through the low expression of MDM2, a negative regulator of p53, in CRC patients with liver metastasis [73]. Similarly, an analysis using the GEO and GEPIA databases confirmed that PRDX2 was overexpressed in CRC and correlated with poor prognosis [74]. The downregulation of PRDX2 led to reduced cell proliferation and tumor growth, accompanied by the upregulation of p53 expression. Furthermore, PRDX2 has been shown to interact with MDM2, leading to the increased ubiquitination and degradation of p53. Thus, PRDX2 is implicated in CRC progression by promoting the ubiquitinated degradation of p53 via the p53 signaling pathway. In another study, it was found that HAUS6 was overexpressed in CRC patients and was associated with shorter overall survival [75]. The knockdown of HAUS6 resulted in the activation of the p53 signaling pathway, leading to an increased expression of downstream p21. Consequently, HAUS6 induces tumor growth and proliferation by regulating the p53 signaling pathway. In addition, there is research indicating that GADD45B, overexpressed in CRC patient tissues, plays a role in apoptosis in CRC [76]. Additionally, while the anti-apoptotic member BCL-2 was upregulated, the pro-apoptotic member Bax was downregulated in these tissues. Furthermore, the knockdown of GADD45B disrupted the balance between anti-apoptotic and pro-apoptotic proteins, indicating that GADD45B influenced apoptosis through the p53 signaling pathway. In a different study, the knockdown of RPS15A, which was highly expressed in CRC patient tissues, has been shown to suppress cell proliferation and induce cell cycle arrest at the G0/G1 phase [77]. Additionally, the downregulation of RPS15A increased the expression of p53 and p21, confirming that RPS15A induces CRC malignancy through the p53 signaling pathway.

3. ncRNA Regulation in CRC

In the mid-1900s, non-protein-coding regions of the human genome were initially dismissed as junk DNA, with no apparent function unlike protein-coding genes [78]. However, with the completion of the Human Genome Project (HGP) in 2003, it became clear that protein-coding genes constitute only about 1.5% (21,000 genes) of the human genome, with non-protein-coding regions making up a significant proportion [79,80]. Subsequent projects, such as the Encyclopedia of DNA Elements (ENCODE) and the Functional Annotation of the Mammalian Genome (FANTOM), demonstrated that junk DNA contains essential regulatory elements for transcription, as well as sequences encoding ncRNAs involved in various biological mechanisms and functions [78,81]. Through these projects, sequencing techniques have significantly advanced and have been used to study ncRNAs in the human genome, revealing substantial insights [82]. ncRNAs, not translated to protein, are generally classified into two categories based on RNA transcript length: those shorter than 200 nucleotides are small ncRNAs, while longer ones are called lncRNAs [83]. miRNAs, representative small ncRNAs, and lncRNAs have been extensively studied in various diseases, including CRC, due to their direct and indirect regulation of protein-coding gene expression [84]. In recent years, circRNAs, ncRNA characterized by their circular structure, have gained attention and have been implicated in the regulation of gene expression, similar to lncRNAs [85]. Therefore, in this section, we aim to discuss the regulatory relationships of miRNAs, lncRNAs, and circRNAs, as well as their impact on CRC (Figure 2).

3.1. miRNA Regulation

miRNAs are a class of small RNA molecules that play a crucial role in the regulation of gene expression in various organisms [86]. They are typically short and single-stranded RNA molecules composed of approximately 18 to 24 nucleotides [87,88]. miRNAs regulate gene expression by binding to the miRNA response element (MRE) of messenger RNA (mRNA) transcribed from genes, thereby modulating post-transcriptional processes [89,90]. By regulating the levels of proteins produced by specific genes, miRNAs play important roles in various biological processes, including development, cell proliferation, differentiation, apoptosis, and immune response [91,92]. The biogenesis of mature miRNAs involves a series of steps (Figure 2). Initially, long primary miRNA transcripts (pri-miRNAs) are transcribed by RNA polymerase II from the miRNA gene [93]. These pri-miRNAs are cleaved in the cell nucleus by an enzyme called Drosha and processed into precursor miRNAs (pre-miRNAs) [94]. Pre-miRNAs are transported to the cytoplasm, where they are further processed by an enzyme called Dicer, resulting in a double-stranded mature, functional miRNA composed of approximately 22 nucleotides. Subsequently, Argonaute binds to the double-stranded mature miRNA [95]. One of the two strands, known as the guide strand, remains in Argonaute, forming the RNA-induced silencing complex, while the other strand, referred to as the passenger strand, is degraded. The guide strand within the RNA-induced silencing complex binds complementarily to the target mRNA, leading to its degradation or the inhibition of translation, thereby directly regulating the expression of the target gene [96].
Research on miRNAs has significantly expanded our understanding of gene regulation and their involvement in various diseases, including cancer, neurodegenerative disorders, and cardiovascular diseases [95,97,98]. Among these, miRNAs exhibit aberrant expression patterns and play a significant role in gene regulation, making them potential targets for diagnosis and treatment (Table 1). In particular, miRNAs influence major signaling pathways in CRC, as discussed in Section 2, resulting in changes in tumor phenotype [99]. For example, miR-590-5p was overexpressed in CRC patient tissues and cell lines and was verified to directly regulate the expression of PDCD4, thereby reducing its expression [100]. It was confirmed that the decreased miR-590-5p induced the expression of PDCD4, resulting in modulating the expression of TGF-β and SMAD2/3, genes involved in the TGF-β signaling pathway. In addition, the downregulation of miR-590-5p induced declined viability, migration, and invasion, suggesting its potential as a therapeutic target for CRC treatment. Similarly, miR-150 was downregulated in both CRC patient tissues and cell lines, exhibiting an inverse expression pattern with β-catenin [101]. Further experiments confirmed CTNNB1, encoding β-catenin, as a target gene of miR-150. The overexpression of miR-150 resulted in decreased cell proliferation, inhibiting CRC progression in mouse models, indicating its involvement in the Wnt signaling pathway and potential as a therapeutic target. In another study, miR-130a-3p was downregulated in CRC patient tissues and various cell lines [102]. The overexpression of miR-130a-3p led to decreased proliferation, viability, and tumor size, implicating its role in the Wnt signaling pathway by reducing the expression of its target gene, Wnt1. These findings suggest the therapeutic potential of targeting miR-130a-3p in CRC treatment.
In addition, numerous studies have shown that miRNAs affect CRC by regulating the expression of genes for which the signaling mechanisms involved have not yet been exactly identified. For instance, RNF187, known for its tumor-suppressive role in other cancers, was overexpressed in CRC patient tissues and cell lines, correlating negatively with poor prognosis [103]. Furthermore, miR-144-5p was decreased in patient tissues and directly regulated the expression of RNF187. Artificial knockdown of RNF187 resulted in suppressed proliferation, migration, and invasion while increasing apoptosis. This suggests that miR-144-5p regulates the expression of RNF187, impacting the progression of CRC and highlighting its potential as a biomarker for CRC diagnosis and therapy. Similarly, research on miR-1-3p suggests its potential as a novel therapeutic strategy for CRC [104]. Downregulated in CRC patient tissues and cell lines, miR-1-3p directly regulated overexpressed YWHAZ. The overexpression of miR-1-3p led to reduced proliferation and invasion, as well as the inhibition of EMT, impacting the development and progression of CRC.
Table 1. The functions of miRNAs with differential expression relating to the main signaling pathways implicated in CRC, as well as experimental information for validation.
Table 1. The functions of miRNAs with differential expression relating to the main signaling pathways implicated in CRC, as well as experimental information for validation.
miRNA
Expression
miRNAsTarget GenesClinical ValueImpact on CRCFunctionSignaling PathwayHuman
Samples
Study ModelReferences
UPmiR-590-5pPDCD4therapeutic targetpathogenesisinduce cell viability, migration, and invasionTGF-β signaling pathway30 tumor and adjacent normal tissuesin vitro[100]
miR-21VMP1therapeutic targetdevelopmentpromote migration and invasion, while repressing autophagy and drug sensitivityPTEN/AKT/TFEB signaling pathway4 tumor and adjacent normal tissuesOncomine dataset, in vitro[105]
miR-21-5pTGF-β1therapeutic targetNAinduce pyroptosis while inhibiting cell viabilityNA5 tumor and normal tissuesin vitro[106]
miR-452-5pPKN2therapeutic targetprogressionenhance cell proliferation, cell cycle transition, and chemoresistance, while suppressing apoptosis MAPK/ERK signaling pathway87 tumor and matched para-carcinoma mucosal tissuesTCGA, in vitro, in vivo[107]
DUSP6
miR-106-5pFAT4NAcarcinogenicpromote proliferation, migration, invasion, and angiogenesisNAtumor and adjacent normal tissuesTCGA, in vitro[108]
miR-496RASSF6therapeutic targetprogressioninduce cell motility, migration, invasion, and EMTWnt signaling pathway28 tumor and adjacent normal tissuesTCGA, in vitro[109]
miR-125bCFTRdiagnosis biomarker and therapeutic targetmetastasisenhance migration, invasion, metastasis, and EMTRhoA/ROCK signaling pathway58 tumor and adjacent normal tissueshuman tissue microarray chips, in vitro, in vivo[110]
CGN
miR-298PTENtherapeutic targetdevelopmentfacilitate cell metabolic activity, cell cycle progression, migration, and invasion, while inhibiting apoptosisAKT/ERK and AKT/mTOR/P70 S6K signaling pathways100 tumors and matched adjacent normal tissues,
100 tumor and normal plasma
in vitro[111]
miR-429LATS2diagnosis biomarker and therapeutic targettumorigenesisstimulate proliferation and tumor growthYAP/TAZ signaling pathway21 tumor and normal tissuesin vitro, in vivo[112]
miR-645EFNA5therapeutic targetmetastasisenhance migration, invasion, metastasis, and tumor growthNA28 tumor and adjacent normal tissuesin vitro, in vivo[113]
DOWNmiR-150CTNNB1therapeutic targetprogressioninhibit cell proliferationWnt signaling pathway30 tumor and adjacent normal tissuesin vitro, in vivo[101]
miR-130a-3pWNT1biomarker, therapeutic targetdevelopmentrepress cell proliferation and tumor growthWnt signaling pathway30 tumor and adjacent normal tissuesTCGA, in vitro, in vivo[102]
miR-144-5pRNF187diagnosis biomarker and therapeutic targetprogressionsuppress migration and invasionNA83 tumor and adjacent normal tissuesin vitro[103]
miR-1-3pYWHAZtherapeutic targetprogressioninhibit cell proliferation and EMTNA20 tumor and adjacent normal tissuesGEO database, in vitro[104]
miR-144-3pBCL6prognosis biomarker and therapeutic targetprogressionrepress cell proliferation and cell cycle progressionWnt signaling pathway20 tumor and adjacent normal tissuesin vitro[114]
miR-450a-5pSOX2progressionsuppress stemness, vasculature, and tumor growthNA90 glass-slide tumor tissues for arrayin vitro, in vivo[115]
miR-215-5pCTNNBIP1therapeutic targetprogression and metastasisinhibit clonogenic potential, cell cycle, migration, invasion, metastasis, and tumor growth, while inducing apoptosisWnt signaling pathwayprimary tumor, paired liver metastatic, and adjacent normal tissuesin vitro, in vivo[116]
miR-148bp55PIKtherapeutic targetprogressionsuppress proliferation, tumor growthp53 signaling pathway10 tumor and normal tissuesin vitro, in vivo[117]
miR-16Survivintherapeutic targetNArepress proliferation and induce apoptosisp53 signaling pathway52 tumor and adjacent normal tissuesin vitro[118]
miR-139-3pKRT80therapeutic targetNAinhibit proliferation, migration, and invasionNA27 tumor and normal tissuesin vitro[119]
miR-125a-5pFNDC3Bprognosis biomarker and therapeutic targetprogressionreduce proliferationPI3K/mTOR signaling pathway36 tumor and adjacent normal tissuesin vitro[120]
miR-217
miR-654-3pSRCdiagnosis and prognosis biomarker, and therapeutic targetdevelopmentrepress proliferation, migration, and invasion, while promoting apoptosisNA103 tumor and adjacent normal tissuesin vitro, in vivo[121]
miR-539TIPEtherapeutic targetprogressionsuppress proliferation and tumor growth, while enhancing ferroptosisSAPK/JNK signaling pathway26 tumor and adjacent tissuesin vitro, in vivo[122]
miR-3622a-3pSALL4therapeutic targetprogression and metastasisreduce proliferation and EMTWnt signaling pathway80 tumor and adjacent normal tissuesTCGA, in vitro, in vivo[123]

3.2. Competing Endogenous RNA (ceRNA) Regulation in CRC

CeRNAs refer to various RNA molecules containing MREs that compete for miRNA binding, of which lncRNAs and circRNAs are well known [124,125]. This dynamic interaction establishes the ceRNA network, where miRNAs compete for interactions with lncRNAs or circRNAs as well as mRNAs. Consequently, it is known that ceRNAs function as a miRNA sponge because they are involved in various physiological processes, such as development, by binding to miRNAs and regulating the availability and activity of miRNAs [126,127,128]. Recent studies have extensively investigated ceRNAs in numerous diseases, particularly CRC [129]. CeRNAs play a pivotal role in CRC initiation and progression by interfering with the binding between miRNAs and mRNAs, thereby serving as potential therapeutic targets or diagnostic biomarkers [130].

3.2.1. LncRNA/miRNA/mRNA

LncRNAs are a class of RNA molecules longer than 200 nucleotides that do not encode proteins but play diverse roles in many biological processes, such as transcriptional control, post-transcriptional processing, and cellular organization, exerting a crucial influence on gene expression regulation [131,132,133,134,135,136]. LncRNAs are transcribed from lncRNA genes by RNA polymerase II (Pol II), synthesizing the nascent lncRNA transcript (Figure 2) [137,138,139]. Similar to mRNA, the nascent lncRNA transcript undergoes processes, including 5′ capping, splicing, and 3′ poly-A tailing, to be stabilized. Some of the produced lncRNAs function in the nucleus, while others transport to the cytoplasm for regulatory roles, where they interact with other RNA molecules, such as miRNAs [140,141]. This interaction influences cellular processes associated with cancer, particularly revealing significant impacts on CRC, suggesting their potential as therapeutic targets for CRC (Table 2) [141,142,143,144].
Accumulated studies have indicated that abnormal expression levels of lncRNAs play an important role in the development and progression of CRC acting as miRNA sponges. In particular, genes implicated in CRC-related signaling pathways, as highlighted in Section 2, are regulated through the interaction of lncRNAs and miRNAs. For example, the overexpression of CTBP1-AS2 in CRC has been identified as a miRNA sponge, influencing CRC development [145]. Increased CTBP1-AS2 activated TGF-β, thereby promoting the proliferation and metastasis of CRC cells while concurrently suppressing apoptosis. Experimental evidence further supported the notion that upregulated miR-93-5p led to a reduction in the expression of both CTBP1-AS2 and TGF-β1, establishing miR-93-5p as a common target for these two factors. Similarly, the increase in SNHG16 contributed to the proliferation of CRC cells, promoting tumor development in CRC [146]. This study revealed that the overexpressed SNHG16 interacted with miR-302-3p, influencing the expression of AKT, a target gene of miR-302-3p. Additionally, the upregulation of miR-302-3p resulted in the increased proliferation of CRC cells, providing evidence that SHNG16 acts as a sponge for miR-302-3p.
In addition to the previously mentioned genes related to important CRC pathways, lncRNAs also play a significant role in CRC by regulating the expression of genes involved in various signaling pathways. In a study, both HIF1A, a target gene of miR-20b-5p, and COL4A2-AS1 expression were overexpressed in CRC patient tissues and cell lines, while miR-20b-5p was decreased [147]. COL4A2-AS1 was observed to compete with HIF1A and acted as a miRNA sponge for miR-20b-3p. Furthermore, reducing COL4A2-AS1 expression in cells and mouse models resulted in decreased cell viability, proliferation, and aerobic glycolysis, thus inhibiting tumorigenesis. These findings suggest the potential of COL4A2-AS1 as a novel target for CRC therapy. In another study, MIR503HG was upregulated in CRC patients, with a more pronounced decrease observed in stages III-IV compared to stages I-II [148]. Its expression was downregulated in patients with lymph node metastasis compared to those without. Increased MIR503HG led to the inhibition of cell proliferation, migration, and invasion, and the induction of apoptosis. Further investigations revealed that MIR503HG inversely regulated miR-107, and PAR4 was negatively controlled by miR-107, establishing a positive relationship between PAR4 and MIR503HG. It was demonstrated that MIR503HG acts as a tumor suppressor by inhibiting CRC development and metastasis, suggesting its potential as a target for CRC treatment.
Table 2. The functions of lncRNAs with differential expression relating to the main signaling pathways implicated in CRC, as well as experimental information for validation.
Table 2. The functions of lncRNAs with differential expression relating to the main signaling pathways implicated in CRC, as well as experimental information for validation.
LncRNA
Expression
LncRNAsmiRNAsGenesClinical ValueImpact on CRCFunctionSignaling PathwayHuman
Samples
Study ModelReferences
UPEGFR-AS1miR-133bSTAT3early diagnosis biomarkerdevelopment and progressioninduce cell proliferation, migration, and invasionNA130 tumor and 30 normal tissuesin vitro[142]
CTBP1-AS2miR-93-5pTGF-β1NAprogression and metastasispromote proliferation, invasion, and metastasis, while inhibiting apoptosisTGF-β signaling pathway50 tumor and normal tissuesGEPIA, in vitro, in vivo[145]
SNHG16miR-302a-3pAKTtherapeutic targetdevelopmentenhance proliferationAKT signaling pathwayNAin vitro[146]
COL4A2-AS1miR-20b-5pHIF1Abiomarker and therapeutic targetprogressionfacilitate proliferation and aerobic glycolysisNA55 tumor and adjacent normal tissuesin vitro, in vivo[147]
MIR4435-2HGmiR-206YAP1prognosis biomarker and therapeutic targetmetastasisstimulate invasion, migration, EMT, metastasis, and growthHippo signaling pathway90 tumor and normal tissuesin vitro, in vivo[149]
NEAT1miR-205-5pVEGFAdiagnosis biomarker and therapeutic targetdevelopmentinduce proliferation, migration, and invasionNA30 tumor and adjacent normal tissuesin vitro[150]
miR-34aSIRT1prognosis biomarker and therapeutic targetmetastasisenhance proliferation, invasion, and tumor growthWnt signaling pathway100 tumor and normal tissuesGEO database, in vitro, in vivo[151]
HOTAIRmiR-206CCL2therapeutic targetprogressionstimulate proliferation and invasion, while inhibiting apoptosisNA32 tumor and normal tissuesTCGA, GEPIA, UALCAN, SurvExpress, in vitro[152]
HCG18miR-1271MTDHtherapeutic targetdevelopmentpromote proliferation and invasionWnt signaling pathway20 tumor and adjacent normal tissuesStarBase, in vitro[153]
RoRmiR-6833-3pSMC4NAtumorigenesisfacilitate proliferation and viability, while inhibiting apoptosisNA24 tumor and normal tissuesin vitro[154]
SNHG8miR-588ATG7therapeutic targetdevelopmentinduce proliferation and autophagyNANATCGA, in vitro[155]
MAFG-AS1miR-149-3pHOXB8prognosis biomarker and therapeutic targetprogressionstimulate proliferation, migration, and invasionNA30 tumor and adjacent normal tissuesin vitro, in vivo[156]
SNHG6miR-181b-5pJAK2biomarker and therapeutic targetprogressionenhance proliferation, while repressing apoptosisNA40 tumor and adjacent normal tissuesin vitro[157]
miR-181c-5p
RHPN1-AS1miR-7-5pOGTtherapeutic targetprogressionpromote proliferation, migration, and invasion, while suppressing apoptosisNANAin vitro, in vivo[158]
CASC21miR-7-5pYAP1therapeutic targetprogressionfacilitate migration, invasion, and EMT, while inhibiting apoptosisNANAGEPIA, in vitro[159]
HOXD-AS1miR-526b-3pCCND1NAprogressionstimulate proliferation, migration, and invasionNA54 tumor and adjacent normal tissuesin vitro[160]
RP11-757G1.5miR-139-5pYAP1prognosis biomarker and therapeutic targetprogression and metastasisenhance proliferation, cell cycle progression, migration, invasion, tumor growth, and metastasisNA43 tumor and 6 adjacent normal tissuesin vitro, in vivo[161]
MCF2L-AS1miR-874-3pCCNE1diagnosis and prognosis biomarkers and therapeutic targetinitiation and progressioninduce proliferation, migration, invasion, and EMT, while inhibiting apoptosisNA130 tumor and normal tissuesin vitro[162]
DANCRmiR-185-5pHMGA2NAprogressionpromote proliferation, migration, and invasionNA50 tumor and adjacent normal tissuesin vitro[163]
RNCR3miR-1301-3pAKT1therapeutic targetprogressionfacilitate proliferation and invasion, while suppressing apoptosisNA76 tumor and adjacent normal tissuesin vitro, in vivo[164]
DOWNMIR503HGmiR-107PAR4therapeutic target suppress migration and invasionNA80 tumor and adjacent normal tissuesin vitro, in vivo[148]
LINC00485miR-581EDEM1therapeutic targetprogression and metastasisrepress proliferation, migration, invasion, tumor growth, and metastasisNA52 tumor and adjacent normal tissuesTCGA, GEO database, in vitro, in vivo[165]
DPP10-AS1miR-127-3pADCY1therapeutic targetprogressioninhibit stemness, sphere formation, proliferation, migration, invasion, and tumor growth, while enhancing apoptosisNA54 rumor and adjacent normal tissuesin vitro, in vivo[166]
MBNL1-AS1miR-412-3pMYL9therapeutic targetprogression, occurrencerepress proliferation, invasion, migration, and tumor formation, while increasing apoptosisNANATCGA, GEO database, in vitro, in vivo[167]
FENDRRmiR-18a-5pING4NAprogression and metastasissuppress proliferation, migration, invasion, tumor growth, and metastasisNA42 tumor and adjacent normal tissuesGEPIA, in vitro, in vivo[168]
MCM3AP-AS1miR-19a-3pFOXF2biomarker and therapeutic targetprogressioninhibit proliferation and migrationNA53 tumor and adjacent normal tissuesGEPIA, in vitro[169]

3.2.2. CircRNA/miRNA/mRNA

CircRNAs are endogenous ncRNAs found in various species, including mammals [170,171]. These molecules exhibit a continuous circular structure, connecting the 3′ and 5′ ends through a phosphodiester bond, and notably lack both a 5′ cap or a 3′ poly-A tail [172,173]. This unique characteristic renders them considerably more stable compared to linear RNAs, providing resistance to exonuclease activity. While circRNAs are not fully understood, they are known to be produced through various synthetic mechanisms, including back-splicing, intron-pairing-driven circularization, and lariat-driven circularization (Figure 2). Back-splicing, a non-canonical splicing event, deviates from the typical splicing pathway that leads to the formation of linear mRNA from precursor mRNA (pre-mRNA) [174,175]. Here, a canonical splice site (5′-GU and 3′-AG at intron) is used, where the 5′ splice donor site downstream of the pre-mRNA connects with the 3′ splice acceptor site upstream, forming a 3′, 5′-phosphodiester bond [173,176,177]. This process results in circRNAs, as the end of the downstream exon is ligated to the beginning of the upstream exon. In intron-pairing-driven circularization, circRNAs are formed by the base pairing of reverse complementary sequences, such as Alu elements, within the flanking long introns of pre-mRNA [178,179]. The third biogenesis type, lariat-driven circularization, occurs through canonical splicing events [178,180,181]. In canonical splicing events, the intronic region of pre-mRNA is debranched, leading to the excision of the intron, and exons are ligated to generate mature mRNA. However, if the debranching process escapes, the lariat structure persists due to the 2′, 5′-phosphodiester bond between the splice donor and branch point. This failure to dismantle the lariat structure maintains the circular form, making it a precursor for circRNAs. Similar to lncRNAs, circRNAs function in the nucleus and play crucial roles in the cytoplasm. Residing in the cytoplasm and containing MRE, circRNAs function as sponges for miRNA, allowing them to act as transcriptional regulators [173,182]. Ultimately, the ability of circRNAs to indirectly regulate mRNA expression has revealed diverse biological functions. Specifically, it has been demonstrated that changes in circRNA expression in CRC impact the initiation, growth, and metastasis of CRC (Table 3) [183,184,185]. Consequently, the altered expression of circRNAs is recognized as a potential biomarker for the diagnosis or treatment of CRC [186,187].
In particular, there are studies on circRNAs that indirectly regulate the expression of genes involved in the key signaling pathways mentioned in Section 2, which play a significant role in CRC. In a study, circ_0008285 expression was downregulated in CRC patient tissues and cell lines [188]. The expression level of circ_0008285 was inversely correlated with tumor size, lymph node metastasis, and tumor–node–metastasis stage. Furthermore, the knockdown of circ_0008285 resulted in the enhanced proliferation and migration of cell lines, revealing its role as a sponge for miR-382-5p. Since miR-382-5p acts as a regulator of PTEN involved in the PI3K/AKT signaling pathway, circ_0008285 affected CRC by indirectly regulating PTEN and is suggested as a potential target for CRC treatment. In a different investigation, the downregulated circ_0009361 in both colorectal cancer patient tissues and cell lines was associated with the promotion of proliferation, EMT, migration, and invasion [189]. Conversely, the overexpression of hsa_circ_0009361 resulted in reduced tumor size and decreased metastasis. Further experiments revealed that hsa_circ_0009361 acted as a miRNA sponge by binding to miR-582, thereby regulating the expression of APC2. Consequently, hsa_circ_0009361 influences colorectal cancer by interacting with miR-582 to regulate APC2 involved in the Wnt signaling pathway.
In addition, it has been revealed that circRNAs can compete with genes unrelated to the main signaling pathways known to have a significant impact on CRC, thereby regulating miRNA. In a study, the expression of circCTNNA1 was upregulated in CRC tissues and associated with poor prognosis, especially in advanced tumor stages showing node metastasis [190]. It has been revealed that circCTNNA1 competed with FOXM1, a target gene of miR-149-5p, thus binding to miR-149-5p. The knockdown of circCTNNA1 resulted in decreased proliferation and tumor size. Therefore, circCTNNA1 plays an oncogenic role, highlighting its potential as a target for CRC therapy and diagnosis. Similarly, circ_GLG1 exhibited high expression in CRC patients, and the downregulation of circ_GLG1 led to the suppression of tumor cell viability, proliferation, migration, and invasion [191]. Circ_GLG1 was identified to act as a sponge for miR-622, influencing the expression of KRAS, a target gene of miR-622. This suggests that circ_GLG1 holds potential as a novel biomarker in CRC diagnosis. In another study, it was observed that circ0065378 expression was decreased in CRC tissues and cell lines [192]. Artificially increasing circ0065378 expression led to reduced proliferation, migration, and invasion of CRC cells. Circ0065378 was found to interact with miR-4701-5p, indirectly regulating the target gene TUSC1 of miR-4701-5p, thereby influencing malignant behavior. These findings suggest that circ-PLXNB1 can serve as a novel target for the diagnosis and treatment of CRC.
Table 3. The functions of circRNAs with differential expression relating to the main signaling pathways implicated in CRC, as well as experimental information for validation.
Table 3. The functions of circRNAs with differential expression relating to the main signaling pathways implicated in CRC, as well as experimental information for validation.
CircRNA
Expression
CircRNAsmiRNAsGenesClinical ValueImpact on CRCFunctionSignaling PathwayHuman SamplesStudy ModelReferences
UPCircVAPAmiR-125aCREB5therapeutic targetprogressioninduce cycle progression, glycolysis, migration, and invasionNA42 tumor and normal tissuesin vitro[187]
CircCTNNA1miR-149-5pFOXM1diagnosis biomarker and therapeutic targetprogressionfacilitate proliferation, DNA synthesis, migration, invasion, and tumorigenesisNA60 tumor and adjacent normal tissuesTCGA, in vitro, in vivo[190]
CircGLG1miR-622KRASdiagnosis biomarker and therapeutic targetprogressionpromote proliferation, invasion, and migrationNA40 tumor and normal tissuesin vitro[191]
CircSPARCmiR-485-3pJAK2diagnosis and prognosis biomarkers and therapeutic targetprogression and metastasisenhance proliferation, migration, invasion, tumor growth, and metastasisJAK/STAT signaling pathway84 tumor and adjacent normal tissues, 40 plasmas of patientsGEO database, in vitro, in vivo[193]
Circ0007142miR-122-5pCDC25Adiagnosis biomarker and therapeutic targetprogressionstimulate proliferation, migration, invasion, and tumor growthNA31 tumor and adjacent normal tissuesin vitro, in vivo[194]
miR-455-5pSGK1therapeutic targetprogressioninduce proliferation, migration, invasion, and tumor growth while repressing apoptosisNA45 tumor and para-carcinoma tissuesin vitro, in vivo[195]
Circ0000467miR-382-5pEN2therapeutic targetdevelopmentenhance proliferation, migration, invasion, and EMTNA69 tumor and adjacent normal tissuesGEO database, in vitro[196]
Circ0001178miR-382ZEB1therapeutic targetmetastasisstimulate migration, invasion, EMT, and metastasisNA102 tumor tissuesin vitro, in vivo[197]
miR-587
miR-616
Circ0053277miR-2467-3pMMP14therapeutic targetdevelopment and progressionpromote proliferation, migration, and EMTNA3 tumor and normal tissuesin vitro[198]
Circ0060745miR-4736CSE1Ltherapeutic targetNAaccelerate proliferation, migration, and invasionNA28 tumor and para-tumor tissuesin vitro[199]
Circ102209miR-761RIN1therapeutic targetprogressionenhance proliferation, cell cycle progression, migration, invasion, EMT, and tumor growth, while suppressing apoptosisNA56 tumor and para-carcinoma tissueshuman circRNA array v2, in vitro, in vivo[200]
Circ100146miR-149HMGA2early diagnosis biomarker and therapeutic targetprogression and metastasisfacilitate proliferation, migration, invasion, tumor growth, and metastasis, while inhibiting apoptosisNA58 tumor and normal tissuesGEO database, in vitro, in vivo[201]
Circ0004277miR-512-5pPTMANAprogressioninduce proliferation and tumor growth, while inhibiting apoptosisNA50 tumor and para-carcinoma tissuesGEO database, in vitro, in vivo[202]
CircFARSAmiR-330-5pLASP1therapeutic targetprogressionpromote proliferation, migration, invasion, and tumor growthNA40 tumor and adjacent normal tissuesin vitro, in vivo[203]
Circ000166miR-326diagnosis biomarker and therapeutic targetprogressionstimulate proliferation, while repressing apoptosisNA40 tumor and adjacent normal tissuesGEO database, in vitro[204]
CircPRMT5miR-377E2F3therapeutic targetprogressionfacilitate proliferation and tumor growthNA30 tumor and adjacent normal tissuesin vitro, in vivo[205]
CircERBINmiR-125a-5p4EBP-1therapeutic targetprogressionenhance proliferation, migration, invasion, tumor growth, angiogenesis, and metastasisNA59 tumor and adjacent normal tissuesin vitro, in vivo[206]
miR-138-5p
CircMAT2BmiR-610E2F1therapeutic targetprogressioninduce proliferationNA70 tumor and adjacent normal tissuesin vitro[207]
Circ0084615miR-599DNMT3Abiomarker and therapeutic targetprogression and metastasispromote proliferation, migration, invasion, and metastasisNA50 tumor and adjacent normal tissuesGEO database, in vitro, in vivo[208]
CircDENND4CmiR-760GLUT1diagnosis biomarker and therapeutic targetprogressionenhance proliferation, migration, glycolysis, and tumor growthNAtumor and normal tissuessequencing, in vitro, in vivo[209]
CircUBAP2miR-582-5pFOXO1biomarker and therapeutic targetprogression and metastasisstimulate migration, invasion, proliferation, autophagy, tumor growth, and metastasisNA3 tumor and normal tissuesin vitro, in vivo[210]
Circ0089153miR-198SENP1therapeutic targetprogressionfacilitate proliferation, sphere formation, tube formation, and tumor growth, while suppressing apoptosisNA50 tumor and adjacent noncancerous tissuesin vitro, in vivo[211]
CircNOX4miR-485-5pCKS1BNAprogressioninduce proliferation, migration, invasion, glycolysis, and tumor growthNA46 tumor and adjacent normal tissuesin vitro, in vivo[212]
CircHERC4miR-556-5pCTBP2prognosis biomarker and therapeutic targetprogression and metastasispromote proliferation, migration, invasion, tumor growth, and metastasisNotch signaling pathway120 tumor and adjacent normal tissuessequencing, in vitro, in vivo[213]
Circ0000372miR-495IL6prognosis biomarker and therapeutic targetprogressionenhance proliferation, migration, invasion, and tumor growthJAK/STAT signaling pathway60 tumor and adjacent normal tissuesin vitro, in vivo[214]
CircLDLRmiR-30a-3pSOAT1biomarker and therapeutic targetprogression and metastasisstimulate proliferation, DNA synthesis, cholesterol, migration, invasion, tumor growth, and metastasisNA80 tumor and 15 normal tissuesin vitro, in vivo[215]
DOWNCirc0008285miR-382-5pPTENtherapeutic targetprogressionrepress proliferation and migrationPI3K/AKT signaling pathway56 tumor and noncancerous colorectal mucosa tissuesin vitro[188]
Circ_0009361miR-582APC2therapeutic targetprogression and metastasisinhibit proliferation, EMT, migration, invasion, tumor growth, and metastasisWnt signaling pathway30 tumor and paracancerous tissuesmicroarray, in vitro, in vivo[189]
Circ0065378miR-4701-5pTUSC1diagnosis biomarker and therapeutic targetprogressionsuppress proliferation, migration, invasion, EMT, and tumor growthNA25 tumor and adjacent normal tissuessequencing, in vitro, in vivo[192]
CircCUL2miR-208a-3pPPP6Cdiagnosis biomarker and therapeutic targetprogressionrepress proliferation, apoptosis, and tumor growth, while increasing apoptosisNA30 tumor and adjacent normal tissuesmicroarray, in vitro, in vivo[216]
CircSMARCA5miR-93-5pARID4Bbiomarkerprogressioninhibit proliferation, migration, invasion, and tumor growthNAtumor and adjacent tissuesin vitro, in vivo[217]
Circ0003266miR-503-5pPDCD4therapeutic targetprogressionsuppress proliferation, migration, and invasion, while accelerating apoptosisNA46 tumor and paracancerous tissuesGEO database, in vitro[218]

4. Conclusions

According to numerous studies and KEGG pathway analyses, the key signaling pathways in CRC are the Wnt, PI3K/AKT/mTOR, MAPK, TGF-β, and p53 signaling pathways. Genes involved in each signaling pathway have been shown to influence behaviors of CRC cells, such as proliferation, migration, and invasion, ultimately impacting the progression and metastasis of CRC. Additionally, extensive research has been conducted on ncRNAs in CRC, as they are known regulators of genes. Overexpressed miRNAs have been found to suppress the expression of tumor suppressor genes, while downregulated miRNAs increased the expression of oncogenes, exerting negative effects on CRC. Furthermore, the discovery of ceRNAs has drawn attention to their indirect regulation of gene expression through the interaction between ceRNAs and miRNAs, thus influencing CRC (Figure 3). However, there are several limitations in the current research. Firstly, some studies solely rely on either in vitro or in vivo experiments, lacking validation in patient samples. Secondly, certain studies are exclusively based on in vitro experiments, lacking confirmation of reproducibility in vivo. Thirdly, the functional roles of genes regulated by ncRNAs remain ambiguous. As discussed in Section 2, although significant signaling pathways in CRC have been identified, Section 3 shows a deficiency in research on ncRNAs regulating genes involved in major signaling pathways. Instead, much research has been performed on gene regulation whose mechanisms have not yet been revealed. Therefore, for the studied ncRNAs to have clinical value, in-depth studies are needed on the regulation of genes involved in key signaling pathways or the mechanisms of genes whose exact functions remain unidentified. As these studies accumulate, it is expected that it will be possible to discover meaningful biomarkers or suitable treatment targets in CRC.

Author Contributions

Conceptualization, Y.J.L. and H.-S.K.; investigation, Y.J.L.; writing—original draft preparation, Y.J.L.; writing—review and editing, Y.J.L., W.R.K., E.G.P., D.H.L., J.-m.K., H.J.S., H.-s.J., H.-Y.R. and H.-S.K.; supervision, H.-S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Siegel, R.L.; Giaquinto, A.N.; Jemal, A. Cancer statistics, 2024. CA Cancer J. Clin. 2024, 74, 12–49. [Google Scholar] [CrossRef]
  2. Hong, Y.; Kim, J.; Choi, Y.J.; Kang, J.G. Clinical study of colorectal cancer operation: Survival analysis. Korean J. Clin. Oncol. 2020, 16, 3. [Google Scholar] [CrossRef]
  3. Vafapour, Z.; Troy, W.; Rashidi, A. Colon cancer detection by designing and analytical evaluation of a water-based THz metamaterial perfect absorber. IEEE Sens. J. 2021, 21, 19307–19313. [Google Scholar] [CrossRef]
  4. Sawicki, T.; Ruszkowska, M.; Danielewicz, A.; Niedźwiedzka, E.; Arłukowicz, T.; Przybyłowicz, K.E. A review of colorectal cancer in terms of epidemiology, risk factors, development, symptoms and diagnosis. Cancers 2021, 13, 2025. [Google Scholar] [CrossRef]
  5. Stefani, C.; Miricescu, D.; Stanescu-Spinu, I.-I.; Nica, R.I.; Greabu, M.; Totan, A.R.; Jinga, M. Growth factors, PI3K/AKT/mTOR and MAPK signaling pathways in colorectal cancer pathogenesis: Where are we now? Int. J. Mol. Sci. 2021, 22, 10260. [Google Scholar] [CrossRef]
  6. Liebl, M.C.; Hofmann, T.G. The role of p53 signaling in colorectal cancer. Cancers 2021, 13, 2125. [Google Scholar] [CrossRef]
  7. Tomita, N.; Ishida, H.; Tanakaya, K.; Yamaguchi, T.; Kumamoto, K.; Tanaka, T.; Hinoi, T.; Miyakura, Y.; Hasegawa, H.; Takayama, T. Japanese society for cancer of the colon and rectum (JSCCR) guidelines 2020 for the clinical practice of hereditary colorectal cancer. Int. J. Clin. Oncol. 2021, 26, 1353–1419. [Google Scholar] [CrossRef]
  8. Ahmad, R.; Singh, J.K.; Wunnava, A.; Al-Obeed, O.; Abdulla, M.; Srivastava, S.K. Emerging trends in colorectal cancer: Dysregulated signaling pathways. Int. J. Mol. Med. 2021, 47, 14. [Google Scholar] [CrossRef]
  9. Wan, M.-l.; Wang, Y.; Zeng, Z.; Deng, B.; Zhu, B.-s.; Cao, T.; Li, Y.-k.; Xiao, J.; Han, Q.; Wu, Q. Colorectal cancer (CRC) as a multifactorial disease and its causal correlations with multiple signaling pathways. Biosci. Rep. 2020, 40, BSR20200265. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, N.; Wang, Z.-Z.; Zhao, M.; Zhang, Y.; Chen, N.-H. Role of non-coding RNA in the pathogenesis of depression. Gene 2020, 735, 144276. [Google Scholar] [CrossRef] [PubMed]
  11. Ma, B.; Wang, S.; Wu, W.; Shan, P.; Chen, Y.; Meng, J.; Xing, L.; Yun, J.; Hao, L.; Wang, X. Mechanisms of circRNA/lncRNA-miRNA interactions and applications in disease and drug research. Biomed. Pharmacother. 2023, 162, 114672. [Google Scholar] [CrossRef] [PubMed]
  12. Nepstad, I.; Hatfield, K.J.; Grønningsæter, I.S.; Reikvam, H. The PI3K-Akt-mTOR signaling pathway in human acute myeloid leukemia (AML) cells. Int. J. Mol. Sci. 2020, 21, 2907. [Google Scholar] [CrossRef] [PubMed]
  13. Malki, A.; ElRuz, R.A.; Gupta, I.; Allouch, A.; Vranic, S.; Al Moustafa, A.-E. Molecular mechanisms of colon cancer progression and metastasis: Recent insights and advancements. Int. J. Mol. Sci. 2020, 22, 130. [Google Scholar] [CrossRef]
  14. Zhao, H.; Ming, T.; Tang, S.; Ren, S.; Yang, H.; Liu, M.; Tao, Q.; Xu, H. Wnt signaling in colorectal cancer: Pathogenic role and therapeutic target. Mol. Cancer 2022, 21, 144. [Google Scholar] [CrossRef] [PubMed]
  15. Steinhart, Z.; Angers, S. Wnt signaling in development and tissue homeostasis. Development 2018, 145, dev146589. [Google Scholar] [CrossRef] [PubMed]
  16. Hayat, R.; Manzoor, M.; Hussain, A. Wnt signaling pathway: A comprehensive review. Cell Biol. Int. 2022, 46, 863–877. [Google Scholar] [CrossRef] [PubMed]
  17. Li, X.; Ortiz, M.A.; Kotula, L. The physiological role of Wnt pathway in normal development and cancer. Medicus. Exp. Biol. Med. 2020, 245, 411–426. [Google Scholar] [CrossRef] [PubMed]
  18. Najafi, S.M.A. Canonical Wnt Signaling (Wnt/β-Catenin Pathway): A Potential Target for Cancer Prevention and Therapy. Iran. Biomed. J. 2020, 24, 269. [Google Scholar] [CrossRef]
  19. Bian, J.; Dannappel, M.; Wan, C.; Firestein, R. Transcriptional regulation of Wnt/β-catenin pathway in colorectal cancer. Cells 2020, 9, 2125. [Google Scholar] [CrossRef]
  20. Schatoff, E.M.; Leach, B.I.; Dow, L.E. Wnt signaling and colorectal cancer. Curr. Color. Cancer Rep. 2017, 13, 101–110. [Google Scholar] [CrossRef]
  21. Fleming-de-Moraes, C.D.; Rocha, M.R.; Tessmann, J.W.; de Araujo, W.M.; Morgado-Diaz, J.A. Crosstalk between PI3K/Akt and Wnt/β-catenin pathways promote colorectal cancer progression regardless of mutational status. Cancer Biol. Ther. 2022, 23, 1–13. [Google Scholar] [CrossRef]
  22. Basu, S.; Haase, G.; Ben-Ze’ev, A. Wnt signaling in cancer stem cells and colon cancer metastasis. F1000Research 2016, 5, 699. [Google Scholar] [CrossRef] [PubMed]
  23. Aghabozorgi, A.S.; Ebrahimi, R.; Bahiraee, A.; Tehrani, S.S.; Nabizadeh, F.; Setayesh, L.; Jafarzadeh-Esfehani, R.; Ferns, G.A.; Avan, A.; Rashidi, Z. The genetic factors associated with Wnt signaling pathway in colorectal cancer. Life Sci. 2020, 256, 118006. [Google Scholar] [CrossRef]
  24. Network, C.G.A. Comprehensive molecular characterization of human colon and rectal cancer. Nature 2012, 487, 330. [Google Scholar] [CrossRef] [PubMed]
  25. La Vecchia, S.; Sebastián, C. Metabolic pathways regulating colorectal cancer initiation and progression. Semin. Cell Dev. Biol. 2020, 98, 63–70. [Google Scholar] [CrossRef]
  26. Yoshida, N.; Kinugasa, T.; Ohshima, K.; Yuge, K.; Ohchi, T.; Fujino, S.; Shiraiwa, S.; Katagiri, M.; Akagi, Y. Analysis of Wnt and β-catenin expression in advanced colorectal cancer. Anticancer Res. 2015, 35, 4403–4410. [Google Scholar] [PubMed]
  27. Kim, W.K.; Kwon, Y.; Jang, M.; Park, M.; Kim, J.; Cho, S.; Jang, D.G.; Lee, W.-B.; Jung, S.H.; Choi, H.J. β-catenin activation down-regulates cell-cell junction-related genes and induces epithelial-to-mesenchymal transition in colorectal cancers. Sci. Rep. 2019, 9, 18440. [Google Scholar] [CrossRef]
  28. Zhang, Z.; Chang, Y.; Zhang, J.; Lu, Y.; Zheng, L.; Hu, Y.; Zhang, F.; Li, X.; Zhang, W.; Li, X. HMGB3 promotes growth and migration in colorectal cancer by regulating WNT/β-catenin pathway. PLoS ONE 2017, 12, e0179741. [Google Scholar] [CrossRef]
  29. Wachsmannova, L.; Mego, M.; Stevurkova, V.; Zajac, V.; Ciernikova, S. Novel strategies for comprehensive mutation screening of the APC gene. Neoplasma 2017, 64, 338–343. [Google Scholar] [CrossRef]
  30. Meng, F.; Liu, X.; Lin, C.; Xu, L.; Liu, J.; Zhang, P.; Zhang, X.; Song, J.; Yan, Y.; Ren, Z. SMYD2 suppresses APC2 expression to activate the Wnt/β-catenin pathway and promotes epithelial-mesenchymal transition in colorectal cancer. Am. J. Cancer Res. 2020, 10, 997. [Google Scholar]
  31. Miricescu, D.; Totan, A.; Stanescu-Spinu, I.-I.; Badoiu, S.C.; Stefani, C.; Greabu, M. PI3K/AKT/mTOR signaling pathway in breast cancer: From molecular landscape to clinical aspects. Int. J. Mol. Sci. 2020, 22, 173. [Google Scholar] [CrossRef] [PubMed]
  32. Ruchi Sharma, V.; Kumar Gupta, G.; Sharma, A.K.; Batra, N.; Sharma, D.K.; Joshi, A.; Sharma, A.K. PI3K/Akt/mTOR intracellular pathway and breast cancer: Factors, mechanism and regulation. Curr. Pharm. Des. 2017, 23, 1633–1638. [Google Scholar] [CrossRef] [PubMed]
  33. Owusu-Brackett, N.; Shariati, M.; Meric-Bernstam, F. Role of PI3K/AKT/mTOR in Cancer Signaling. In Predictive Biomarkers in Oncology: Applications in Precision Medicine; Springer: Berlin/Heidelberg, Germany, 2019; pp. 263–270. [Google Scholar]
  34. Wang, Q.; Zhang, P.; Zhang, W.; Zhang, X.; Chen, J.; Ding, P.; Li, L.; Lv, X.; Li, L.; Hu, W. PI3K activation is enhanced by FOXM1D binding to p110 and p85 subunits. Signal Transduct. Target. Ther. 2020, 5, 105. [Google Scholar] [CrossRef] [PubMed]
  35. Mazloumi Gavgani, F.; Smith Arnesen, V.; Jacobsen, R.G.; Krakstad, C.; Hoivik, E.A.; Lewis, A.E. Class I phosphoinositide 3-Kinase PIK3CA/p110α and PIK3CB/p110β isoforms in endometrial cancer. Int. J. Mol. Sci. 2018, 19, 3931. [Google Scholar] [CrossRef] [PubMed]
  36. Xue, C.; Li, G.; Lu, J.; Li, L. Crosstalk between circRNAs and the PI3K/AKT signaling pathway in cancer progression. Signal Transduct. Target. Ther. 2021, 6, 400. [Google Scholar] [CrossRef] [PubMed]
  37. Papadatos-Pastos, D.; Rabbie, R.; Ross, P.; Sarker, D. The role of the PI3K pathway in colorectal cancer. Crit. Rev. Oncol. Hematol. 2015, 94, 18–30. [Google Scholar] [CrossRef] [PubMed]
  38. Moafian, Z.; Maghrouni, A.; Soltani, A.; Hashemy, S.I. Cross-talk between non-coding RNAs and PI3K/AKT/mTOR pathway in colorectal cancer. Mol. Biol. Rep. 2021, 48, 4797–4811. [Google Scholar] [CrossRef] [PubMed]
  39. Johnson, S.M.; Gulhati, P.; Rampy, B.A.; Han, Y.; Rychahou, P.G.; Doan, H.Q.; Weiss, H.L.; Evers, B.M. Novel expression patterns of PI3K/Akt/mTOR signaling pathway components in colorectal cancer. J. Am. Coll. Surg. 2010, 210, 767–776. [Google Scholar] [CrossRef] [PubMed]
  40. Zhu, Y.-F.; Yu, B.-H.; Li, D.-L.; Ke, H.-L.; Guo, X.-Z.; Xiao, X.-Y. PI3K expression and PIK3CA mutations are related to colorectal cancer metastases. World J. Gastroenterol. 2012, 18, 3745. [Google Scholar] [CrossRef]
  41. Suman, S.; Kurisetty, V.; Das, T.P.; Vadodkar, A.; Ramos, G.; Lakshmanaswamy, R.; Damodaran, C. Activation of AKT signaling promotes epithelial–mesenchymal transition and tumor growth in colorectal cancer cells. Mol. Carcinog. 2014, 53, E151–E160. [Google Scholar] [CrossRef]
  42. Cai, Z.; Ke, J.; He, X.; Yuan, R.; Chen, Y.; Wu, X.; Wang, L.; Wang, J.; Lan, P.; Wu, X. Significance of mTOR signaling and its inhibitor against cancer stem-like cells in colorectal cancer. Ann. Surg. Oncol. 2014, 21, 179–188. [Google Scholar] [CrossRef]
  43. Xu, W.; Yu, M.; Qin, J.; Luo, Y.; Zhong, M. LACTB regulates PIK3R3 to promote autophagy and inhibit EMT and proliferation through the PI3K/AKT/mTOR signaling pathway in colorectal cancer. Cancer Manag. Res. 2020, 12, 5181–5200. [Google Scholar] [CrossRef]
  44. Feng, Y.; Ji, D.; Huang, Y.; Ji, B.; Zhang, Y.; Li, J.; Peng, W.; Zhang, C.; Zhang, D.; Sun, Y. TGM3 functions as a tumor suppressor by repressing epithelial-to-mesenchymal transition and the PI3K/AKT signaling pathway in colorectal cancer. Oncol. Rep. 2020, 43, 864–876. [Google Scholar] [CrossRef]
  45. Su, W.; Feng, B.; Hu, L.; Guo, X.; Yu, M. MUC3A promotes the progression of colorectal cancer through the PI3K/Akt/mTOR pathway. BMC Cancer 2022, 22, 602. [Google Scholar] [CrossRef]
  46. Guo, Y.J.; Pan, W.W.; Liu, S.B.; Shen, Z.F.; Xu, Y.; Hu, L.L. ERK/MAPK signalling pathway and tumorigenesis. Exp. Ther. Med. 2020, 19, 1997–2007. [Google Scholar] [CrossRef] [PubMed]
  47. Sheng, Y.-N.; Luo, Y.-H.; Liu, S.-B.; Xu, W.-T.; Zhang, Y.; Zhang, T.; Xue, H.; Zuo, W.-B.; Li, Y.-N.; Wang, C.-Y. Zeaxanthin induces apoptosis via ROS-regulated MAPK and AKT signaling pathway in human gastric cancer cells. Onco Targets Ther. 2020, 13, 10995–11006. [Google Scholar] [CrossRef]
  48. Asl, E.R.; Amini, M.; Najafi, S.; Mansoori, B.; Mokhtarzadeh, A.; Mohammadi, A.; Lotfinejad, P.; Bagheri, M.; Shirjang, S.; Lotfi, Z. Interplay between MAPK/ERK signaling pathway and MicroRNAs: A crucial mechanism regulating cancer cell metabolism and tumor progression. Life Sci. 2021, 278, 119499. [Google Scholar] [CrossRef]
  49. Kciuk, M.; Gielecińska, A.; Budzinska, A.; Mojzych, M.; Kontek, R. Metastasis and MAPK pathways. Int. J. Mol. Sci. 2022, 23, 3847. [Google Scholar] [CrossRef] [PubMed]
  50. Fang, J.Y.; Richardson, B.C. The MAPK signalling pathways and colorectal cancer. Lancet Oncol. 2005, 6, 322–327. [Google Scholar] [CrossRef] [PubMed]
  51. Zhou, G.; Yang, J.; Song, P. Correlation of ERK/MAPK signaling pathway with proliferation and apoptosis of colon cancer cells. Oncol. Lett. 2019, 17, 2266–2270. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.; Xie, G.; Li, M.; Du, J.; Wang, M. COPB2 gene silencing inhibits colorectal cancer cell proliferation and induces apoptosis via the JNK/c-Jun signaling pathway. PLoS ONE 2020, 15, e0240106. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, Y.; Liao, Y.; Chen, C.; Sun, W.; Sun, X.; Liu, Y.; Xu, E.; Lai, M.; Zhang, H. p38-regulated FOXC1 stability is required for colorectal cancer metastasis. J. Pathol. Inform. 2020, 250, 217–230. [Google Scholar] [CrossRef]
  54. Ou, S.; Liao, Y.; Shi, J.; Tang, J.; Ye, Y.; Wu, F.; Wang, W.; Fei, J.; Xie, F.; Bai, L. S100A16 suppresses the proliferation, migration and invasion of colorectal cancer cells in part via the JNK/p38 MAPK pathway. Mol. Med. Rep. 2021, 23, 164. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, W.; Wang, X.; Peng, L.; Deng, Q.; Liang, Y.; Qing, H.; Jiang, B. CD24-dependent MAPK pathway activation is required for colorectal cancer cell proliferation. Cancer Sci. 2010, 101, 112–119. [Google Scholar] [CrossRef]
  56. Itatani, Y.; Kawada, K.; Sakai, Y. Transforming growth factor-β signaling pathway in colorectal cancer and its tumor microenvironment. Int. J. Mol. Sci. 2019, 20, 5822. [Google Scholar] [CrossRef] [PubMed]
  57. Zhao, H.; Wei, J.; Sun, J. Roles of TGF-β signaling pathway in tumor microenvirionment and cancer therapy. Int. Immunopharmacol. 2020, 89, 107101. [Google Scholar] [CrossRef]
  58. Pawlak, J.B.; Blobe, G.C. TGF-β superfamily co-receptors in cancer. Dev. Dynam. 2022, 251, 117–143. [Google Scholar] [CrossRef]
  59. Tzavlaki, K.; Moustakas, A. TGF-β Signaling. Biomolecules 2020, 10, 487. [Google Scholar] [CrossRef]
  60. Yu, X.-Y.; Sun, Q.; Zhang, Y.-M.; Zou, L.; Zhao, Y.-Y. TGF-β/Smad signaling pathway in tubulointerstitial fibrosis. Front. Pharmacol. 2022, 13, 860588. [Google Scholar] [CrossRef]
  61. Yang, D.; Hou, T.; Li, L.; Chu, Y.; Zhou, F.; Xu, Y.; Hou, X.; Song, H.; Zhu, K.; Hou, Z. Smad1 promotes colorectal cancer cell migration through Ajuba transactivation. Oncotarget 2017, 8, 110415. [Google Scholar] [CrossRef]
  62. Shi, W.; He, J.; Huang, Y.; Zeng, Z.; Feng, Z.; Xu, H.; Nie, Y. Integrin β5 enhances the malignancy of human colorectal cancer by increasing the TGF-β signaling. Anti-Cancer Drugs 2021, 32, 717–726. [Google Scholar] [CrossRef]
  63. Wang, L.; Ai, M.; Nie, M.; Zhao, L.; Deng, G.; Hu, S.; Han, Y.; Zeng, W.; Wang, Y.; Yang, M. EHF promotes colorectal carcinoma progression by activating TGF-β1 transcription and canonical TGF-β signaling. Cancer Sci. 2020, 111, 2310–2324. [Google Scholar] [CrossRef]
  64. Williams, A.B.; Schumacher, B. p53 in the DNA-damage-repair process. Cold Spring Harb. Perspect. Med. 2016, 6, a026070. [Google Scholar] [CrossRef]
  65. Borrero, L.J.H.; El-Deiry, W.S. Tumor suppressor p53: Biology, signaling pathways, and therapeutic targeting. Biochim. Biophys. Acta 2021, 1876, 188556. [Google Scholar]
  66. Fu, X.; Wu, S.; Li, B.; Xu, Y.; Liu, J. Functions of p53 in pluripotent stem cells. Protein Cell 2020, 11, 71–78. [Google Scholar] [CrossRef] [PubMed]
  67. Han, X.; Wei, W.; Sun, Y. PROTAC degraders with ligands recruiting MDM2 E3 ubiquitin ligase: An updated perspective. Acta Mater. Med. 2022, 1, 244. [Google Scholar] [CrossRef] [PubMed]
  68. Eskandari, M.; Shi, Y.; Liu, J.; Albanese, J.; Goel, S.; Verma, A.; Wang, Y. The expression of MDM2, MDM4, p53 and p21 in myeloid neoplasms and the effect of MDM2/MDM4 dual inhibitor. Leuk Lymphoma 2021, 62, 167–175. [Google Scholar] [CrossRef]
  69. Wang, P.; Wang, H.-Y.; Gao, X.-J.; Zhu, H.-X.; Zhang, X.-P.; Liu, F.; Wang, W. Encoding and decoding of p53 dynamics in cellular response to stresses. Cells 2023, 12, 490. [Google Scholar] [CrossRef]
  70. Solomon, H.; Dinowitz, N.; Pateras, I.S.; Cooks, T.; Shetzer, Y.; Molchadsky, A.; Charni, M.; Rabani, S.; Koifman, G.; Tarcic, O. Mutant p53 gain of function underlies high expression levels of colorectal cancer stem cells markers. Oncogene 2018, 37, 1669–1684. [Google Scholar] [CrossRef]
  71. Nagao, K.; Koshino, A.; Sugimura-Nagata, A.; Nagano, A.; Komura, M.; Ueki, A.; Ebi, M.; Ogasawara, N.; Tsuzuki, T.; Kasai, K. The complete loss of p53 expression uniquely predicts worse prognosis in colorectal cancer. Int. J. Mol. Sci. 2022, 23, 3252. [Google Scholar] [CrossRef]
  72. Kruschewski, M.; Mueller, K.; Lipka, S.; Budczies, J.; Noske, A.; Buhr, H.J.; Elezkurtaj, S. The prognostic impact of p53 expression on sporadic colorectal cancer is dependent on p21 status. Cancers 2011, 3, 1274–1284. [Google Scholar] [CrossRef] [PubMed]
  73. Kondo, I.; Iida, S.; Takagi, Y.; Sugihara, K. MDM2 mRNA expression in the p53 pathway may predict the potential of invasion and liver metastasis in colorectal cancer. Dis. Colon Rectum 2008, 51, 1395–1402. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, W.; Wei, J.; Zhang, H.; Zheng, X.; Zhou, H.; Luo, Y.; Yang, J.; Deng, Q.; Huang, S.; Fu, Z. PRDX2 promotes the proliferation of colorectal cancer cells by increasing the ubiquitinated degradation of p53. Cell Death Dis. 2021, 12, 605. [Google Scholar] [CrossRef] [PubMed]
  75. Shen, A.; Liu, L.; Huang, Y.; Shen, Z.; Wu, M.; Chen, X.; Wu, X.; Lin, X.; Chen, Y.; Li, L. Down-regulating HAUS6 suppresses cell proliferation by activating the p53/p21 pathway in colorectal cancer. Front. Cell Dev. Biol. 2022, 9, 772077. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, L.; Xiao, X.; Li, D.; Chi, Y.; Wei, P.; Wang, Y.; Ni, S.; Tan, C.; Zhou, X.; Du, X. Abnormal expression of GADD45B in human colorectal carcinoma. J. Transl. Med. 2012, 10, 215. [Google Scholar] [CrossRef] [PubMed]
  77. Chen, J.; Wei, Y.; Feng, Q.; Ren, L.; He, G.; Chang, W.; Zhu, D.; Yi, T.; Lin, Q.; Tang, W. Ribosomal protein S15A promotes malignant transformation and predicts poor outcome in colorectal cancer through misregulation of p53 signaling pathway. Int. J. Oncol. 2016, 48, 1628–1638. [Google Scholar] [CrossRef] [PubMed]
  78. Science in School. Available online: https://www.scienceinschool.org/article/2023/the-importance-of-non-coding-rnas/ (accessed on 1 February 2024).
  79. Moraes, F.; Góes, A. A decade of human genome project conclusion: Scientific diffusion about our genome knowledge. Biochem. Mol. Biol. Educ. 2016, 44, 215–223. [Google Scholar] [CrossRef] [PubMed]
  80. Oyelade, J.; Isewon, I.; Ogunlana, O.; Aworunse, O.; Oyesola, O.; Aromolaran, O.; Dokumu, T.; Ademuwagun, I.; Iheagwam, F.; Babatunde, E. Overview of the human genome. In Genome Plasticity in Health and Disease; Academic Press: Cambridge, MA, USA, 2020; pp. 9–26. [Google Scholar]
  81. Santosh, B.; Varshney, A.; Yadava, P.K. Non-coding RNAs: Biological functions and applications. Cell Biochem. Funct. 2015, 33, 14–22. [Google Scholar] [CrossRef] [PubMed]
  82. Micheel, J.; Safrastyan, A.; Wollny, D. Advances in non-coding RNA sequencing. Non-Coding RNA 2021, 7, 70. [Google Scholar] [CrossRef]
  83. Bhatti, G.K.; Khullar, N.; Sidhu, I.S.; Navik, U.S.; Reddy, A.P.; Reddy, P.H.; Bhatti, J.S. Emerging role of non-coding RNA in health and disease. Metab. Brain Dis. 2021, 36, 1119–1134. [Google Scholar] [CrossRef]
  84. Luo, R.; Song, J.; Zhang, W.; Ran, L. Identification of MFI2-AS1, a novel pivotal lncRNA for prognosis of stage III/IV colorectal cancer. Dig. Dis. Sci. 2020, 65, 3538–3550. [Google Scholar] [CrossRef] [PubMed]
  85. Shao, T.; Pan, Y.-h.; Xiong, X.-d. Circular RNA: An important player with multiple facets to regulate its parental gene expression. Mol. Ther. Nucleic Acids 2021, 23, 369–376. [Google Scholar] [CrossRef]
  86. Gebremedhn, S.; Ali, A.; Hossain, M.; Hoelker, M.; Salilew-Wondim, D.; Anthony, R.V.; Tesfaye, D. MicroRNA-mediated gene regulatory mechanisms in mammalian female reproductive health. Int. J. Mol. Sci. 2021, 22, 938. [Google Scholar] [CrossRef]
  87. Wang, L.-L.; Chen, W.-Q.; Wang, Y.-R.; Zeng, L.-P.; Chen, T.-T.; Chen, G.-Y.; Chen, J.-H. Numerous long single-stranded DNAs produced by dual amplification reactions for electrochemical detection of exosomal microRNAs. Biosens. Bioelectron. 2020, 169, 112555. [Google Scholar] [CrossRef] [PubMed]
  88. Wallace, D.R.; Taalab, Y.M.; Heinze, S.; Tariba Lovaković, B.; Pizent, A.; Renieri, E.; Tsatsakis, A.; Farooqi, A.A.; Javorac, D.; Andjelkovic, M. Toxic-metal-induced alteration in miRNA expression profile as a proposed mechanism for disease development. Cells 2020, 9, 901. [Google Scholar] [CrossRef] [PubMed]
  89. Soni, D.K.; Biswas, R. Role of non-coding RNAs in post-transcriptional regulation of lung diseases. Front. Genet. 2021, 12, 767348. [Google Scholar] [CrossRef]
  90. Riolo, G.; Cantara, S.; Marzocchi, C.; Ricci, C. miRNA targets: From prediction tools to experimental validation. Methods Protoc. 2020, 4, 1. [Google Scholar] [CrossRef] [PubMed]
  91. Indrieri, A.; Carrella, S.; Carotenuto, P.; Banfi, S.; Franco, B. The pervasive role of the miR-181 family in development, neurodegeneration, and cancer. Int. J. Mol. Sci. 2020, 21, 2092. [Google Scholar] [CrossRef]
  92. Kalayinia, S.; Arjmand, F.; Maleki, M.; Malakootian, M.; Singh, C.P. MicroRNAs: Roles in cardiovascular development and disease. Cardiovasc. Pathol. 2021, 50, 107296. [Google Scholar] [CrossRef]
  93. Prasad, A.; Sharma, N.; Prasad, M. Noncoding but coding: Pri-miRNA into the action. Trends Plant Sci. 2021, 26, 204–206. [Google Scholar] [CrossRef]
  94. Jin, W.; Wang, J.; Liu, C.-P.; Wang, H.-W.; Xu, R.-M. Structural basis for pri-miRNA recognition by Drosha. Mol. Cell 2020, 78, 423–433. [Google Scholar] [CrossRef] [PubMed]
  95. Hill, M.; Tran, N. miRNA interplay: Mechanisms and consequences in cancer. Dis. Model. Mech. 2021, 14, dmm047662. [Google Scholar] [CrossRef] [PubMed]
  96. Klicka, K.; Grzywa, T.M.; Mielniczuk, A.; Klinke, A.; Włodarski, P.K. The role of miR-200 family in the regulation of hallmarks of cancer. Front. Oncol. 2022, 12, 965231. [Google Scholar] [CrossRef] [PubMed]
  97. Paul, S.; Bravo Vázquez, L.A.; Pérez Uribe, S.; Roxana Reyes-Pérez, P.; Sharma, A. Current status of microRNA-based therapeutic approaches in neurodegenerative disorders. Cells 2020, 9, 1698. [Google Scholar] [CrossRef] [PubMed]
  98. Çakmak, H.A.; Demir, M. MicroRNA and cardiovascular diseases. Balkan Med. J. 2020, 37, 60. [Google Scholar] [CrossRef] [PubMed]
  99. Pan, S.; Wu, W.; Ren, F.; Li, L.; Li, Y.; Li, W.; Wang, A.; Liu, D.; Dong, Y. MiR-346-5p promotes colorectal cancer cell proliferation in vitro and in vivo by targeting FBXL2 and activating the β-catenin signaling pathway. Life Sci. 2020, 244, 117300. [Google Scholar] [CrossRef] [PubMed]
  100. Guo, T.; Wang, J.; Cheng, G.; Huang, H. miR-590-5p may regulate colorectal cancer cell viability and migration by targeting PDCD4. Exp. Ther. Med. 2020, 20, 55. [Google Scholar] [CrossRef]
  101. He, Z.; Dang, J.; Song, A.; Cui, X.; Ma, Z.; Zhang, Y. The involvement of miR-150/β-catenin axis in colorectal cancer progression. Biomed. Pharmacother. 2020, 121, 109495. [Google Scholar] [CrossRef] [PubMed]
  102. Song, G.-L.; Xiao, M.; Wan, X.-Y.; Deng, J.; Ling, J.-D.; Tian, Y.-G.; Li, M.; Yin, J.; Zheng, R.-Y.; Tang, Y. MiR-130a-3p suppresses colorectal cancer growth by targeting Wnt Family Member 1 (WNT1). Bioengineered 2021, 12, 8407–8418. [Google Scholar] [CrossRef]
  103. Gao, Z.; Jiang, J.; Hou, L.; Zhang, B. Dysregulation of MiR-144-5p/RNF187 axis contributes to the progression of colorectal cancer. J. Transl. Med. 2022, 10, 65–75. [Google Scholar] [CrossRef]
  104. Du, G.; Yu, X.; Chen, Y.; Cai, W. MiR-1-3p Suppresses Colorectal Cancer Cell Proliferation and Metastasis by Inhibiting YWHAZ-Mediated Epithelial–Mesenchymal Transition. Front. Oncol. 2021, 11, 634596. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, C.; Peng, R.; Zeng, M.; Zhang, Z.; Liu, S.; Jiang, D.; Lu, Y.; Zou, F. An autoregulatory feedback loop of miR-21/VMP1 is responsible for the abnormal expression of miR-21 in colorectal cancer cells. Cell Death Dis. 2020, 11, 1067. [Google Scholar] [CrossRef] [PubMed]
  106. Jiang, R.; Chen, X.; Ge, S.; Wang, Q.; Liu, Y.; Chen, H.; Xu, J.; Wu, J. MiR-21-5p induces pyroptosis in colorectal cancer via TGFBI. Front. Oncol. 2021, 10, 610545. [Google Scholar] [CrossRef] [PubMed]
  107. Lin, X.; Han, L.; Gu, C.; Lai, Y.; Lai, Q.; Li, Q.; He, C.; Meng, Y.; Pan, L.; Liu, S. MiR-452-5p promotes colorectal cancer progression by regulating an ERK/MAPK positive feedback loop. Aging 2021, 13, 7608. [Google Scholar] [CrossRef] [PubMed]
  108. Pan, M.; Chen, Q.; Lu, Y.; Wei, F.; Chen, C.; Tang, G.; Huang, H. MiR-106b-5p regulates the migration and invasion of colorectal cancer cells by targeting FAT4. Biosci. Rep. 2020, 40, BSR20200098. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, H.; Yan, B.; Zhang, P.; Liu, S.; Li, Q.; Yang, J.; Yang, F.; Chen, E. MiR-496 promotes migration and epithelial-mesenchymal transition by targeting RASSF6 in colorectal cancer. J. Cell. Physiol. 2020, 235, 1469–1479. [Google Scholar] [CrossRef] [PubMed]
  110. Zhang, X.; Li, T.; Han, Y.-N.; Ge, M.; Wang, P.; Sun, L.; Liu, H.; Cao, T.; Nie, Y.; Fan, D. miR-125b promotes colorectal cancer migration and invasion by dual-targeting CFTR and CGN. Cancers 2021, 13, 5710. [Google Scholar] [CrossRef]
  111. Arabsorkhi, Z.; Gharib, E.; Yaghmoorian Khojini, J.; Farhadieh, M.E.; Nazemalhosseini-Mojarad, E.; Zali, M.R. miR-298 plays a pivotal role in colon cancer invasiveness by targeting PTEN. J. Cell. Physiol. 2020, 235, 4335–4350. [Google Scholar] [CrossRef]
  112. Chen, X.; Wang, A.-l.; Liu, Y.-y.; Zhao, C.-x.; Zhou, X.; Liu, H.-l.; Lin, M.-b. MiR-429 involves in the pathogenesis of colorectal cancer via directly targeting LATS2. Oxid. Med. Cell. Longev. 2020, 2020, 5316276. [Google Scholar] [CrossRef]
  113. Li, S.; Hou, X.; Wu, C.; Han, L.; Li, Q.; Wang, J.; Luo, S. MiR-645 promotes invasiveness, metastasis and tumor growth in colorectal cancer by targeting EFNA5. Biomed. Pharmacother. 2020, 125, 109889. [Google Scholar] [CrossRef]
  114. Sun, N.; Zhang, L.; Zhang, C.; Yuan, Y. miR-144-3p inhibits cell proliferation of colorectal cancer cells by targeting BCL6 via inhibition of Wnt/β-catenin signaling. Cell. Mol. Biol. Lett. 2020, 25, 19. [Google Scholar] [CrossRef] [PubMed]
  115. Chen, J.; Chen, S.; Zhuo, L.; Zhu, Y.; Zheng, H. Regulation of cancer stem cell properties, angiogenesis, and vasculogenic mimicry by miR-450a-5p/SOX2 axis in colorectal cancer. Cell Death Dis. 2020, 11, 173. [Google Scholar] [CrossRef] [PubMed]
  116. Machackova, T.; Vychytilova-Faltejskova, P.; Souckova, K.; Trachtova, K.; Brchnelova, D.; Svoboda, M.; Kiss, I.; Prochazka, V.; Kala, Z.; Slaby, O. MiR-215-5p reduces liver metastasis in an experimental model of colorectal cancer through regulation of ECM-receptor interactions and focal adhesion. Cancers 2020, 12, 3518. [Google Scholar] [CrossRef] [PubMed]
  117. Wang, G.; Cao, X.; Lai, S.; Luo, X.; Feng, Y.; Wu, J.; Ning, Q.; Xia, X.; Wang, J.; Gong, J. Altered p53 regulation of miR-148b and p55PIK contributes to tumor progression in colorectal cancer. Oncogene 2015, 34, 912–921. [Google Scholar] [CrossRef] [PubMed]
  118. Ma, Q.; Wang, X.; Li, Z.; Li, B.; Ma, F.; Peng, L.; Zhang, Y.; Xu, A.; Jiang, B. microRNA-16 represses colorectal cancer cell growth in vitro by regulating the p53/survivin signaling pathway. Oncol. Rep. 2013, 29, 1652–1658. [Google Scholar] [CrossRef] [PubMed]
  119. Yasudome, R.; Seki, N.; Asai, S.; Goto, Y.; Kita, Y.; Hozaka, Y.; Wada, M.; Tanabe, K.; Idichi, T.; Mori, S. Molecular pathogenesis of colorectal cancer: Impact of oncogenic targets regulated by tumor suppressive miR-139-3p. Int. J. Mol. Sci. 2022, 23, 11616. [Google Scholar] [CrossRef] [PubMed]
  120. Li, Y.; Yang, J.; Wang, H.; Qiao, W.; Guo, Y.; Zhang, S.; Guo, Y. FNDC3B, targeted by miR-125a-5p and miR-217, promotes the proliferation and invasion of colorectal cancer cells via PI3K/mTOR signaling. Onco Targets Ther. 2020, 13, 3501–3510. [Google Scholar] [CrossRef]
  121. Zhang, H.; Shen, Z.; Zhou, Y.; Zhang, Z.; Wang, Q.; Zhang, M.; Wang, S.; Wang, B. Downregulation of miR-654-3p in colorectal cancer indicates poor prognosis and promotes cell proliferation and invasion by targeting SRC. Front. Genet. 2020, 11, 577948. [Google Scholar] [CrossRef]
  122. Yang, Y.; Lin, Z.; Han, Z.; Wu, Z.; Hua, J.; Zhong, R.; Zhao, R.; Ran, H.; Qu, K.; Huang, H. miR-539 activates the SAPK/JNK signaling pathway to promote ferropotosis in colorectal cancer by directly targeting TIPE. Cell Death Discov. 2021, 7, 272. [Google Scholar] [CrossRef]
  123. Chang, S.; Sun, G.; Zhang, D.; Li, Q.; Qian, H. MiR-3622a-3p acts as a tumor suppressor in colorectal cancer by reducing stemness features and EMT through targeting spalt-like transcription factor 4. Cell Death Dis. 2020, 11, 592. [Google Scholar] [CrossRef]
  124. Liu, H.-T.; Ma, R.-R.; Lv, B.-B.; Zhang, H.; Shi, D.-B.; Guo, X.-Y.; Zhang, G.-H.; Gao, P. LncRNA-HNF1A-AS1 functions as a competing endogenous RNA to activate PI3K/AKT signalling pathway by sponging miR-30b-3p in gastric cancer. Br. J. Cancer 2020, 122, 1825–1836. [Google Scholar] [CrossRef]
  125. Ala, U. Competing endogenous RNAs, non-coding RNAs and diseases: An intertwined story. Cells 2020, 9, 1574. [Google Scholar] [CrossRef] [PubMed]
  126. Xu, G.; Xu, W.-Y.; Xiao, Y.; Jin, B.; Du, S.-D.; Mao, Y.-l.; Zhang, Z.-T. The emerging roles of non-coding competing endogenous RNA in hepatocellular carcinoma. Cancer Cell Int. 2020, 20, 496. [Google Scholar] [CrossRef] [PubMed]
  127. Mitra, A.; Pfeifer, K.; Park, K.-S. Circular RNAs and competing endogenous RNA (ceRNA) networks. Transl. Cancer Res. 2018, 7 (Suppl. 5), S624. [Google Scholar] [CrossRef] [PubMed]
  128. Li, D.; Zhang, J.; Li, J. Role of miRNA sponges in hepatocellular carcinoma. Clin. Chim. Acta. 2020, 500, 10–19. [Google Scholar] [CrossRef] [PubMed]
  129. Li, K.; Yao, T.; Zhang, Y.; Li, W.; Wang, Z. NEAT1 as a competing endogenous RNA in tumorigenesis of various cancers: Role, mechanism and therapeutic potential. Int. J. Biol. Sci. 2021, 17, 3428. [Google Scholar] [CrossRef] [PubMed]
  130. Qi, X.; Lin, Y.; Liu, X.; Chen, J.; Shen, B. Biomarker discovery for the carcinogenic heterogeneity between colon and rectal cancers based on lncRNA-associated ceRNA network analysis. Front. Oncol. 2020, 10, 535985. [Google Scholar] [CrossRef] [PubMed]
  131. Fang, C.; Wang, L.; Gong, C.; Wu, W.; Yao, C.; Zhu, S. Long non-coding RNAs: How to regulate the metastasis of non–small-cell lung cancer. J. Cell. Mol. Med. 2020, 24, 3282–3291. [Google Scholar] [CrossRef]
  132. Yang, X.; Xie, Z.; Lei, X.; Gan, R. Long non-coding RNA GAS5 in human cancer. Oncol. Lett. 2020, 20, 2587–2594. [Google Scholar] [CrossRef]
  133. Pisignano, G.; Ladomery, M. Post-transcriptional regulation through long non-coding rnas (lncrnas). Non-Coding RNA 2021, 7, 29. [Google Scholar] [CrossRef]
  134. Mattick, J.S.; Amaral, P.P.; Carninci, P.; Carpenter, S.; Chang, H.Y.; Chen, L.-L.; Chen, R.; Dean, C.; Dinger, M.E.; Fitzgerald, K.A. Long non-coding RNAs: Definitions, functions, challenges and recommendations. Nat. Rev. Mol. Cell Biol. 2023, 24, 430–447. [Google Scholar] [CrossRef] [PubMed]
  135. Neve, B.; Jonckheere, N.; Vincent, A.; Van Seuningen, I. Long non-coding RNAs: The tentacles of chromatin remodeler complexes. Cell. Mol. Life Sci. 2021, 78, 1139–1161. [Google Scholar] [CrossRef] [PubMed]
  136. Gil, N.; Ulitsky, I. Regulation of gene expression by cis-acting long non-coding RNAs. Nat. Rev. Genet 2020, 21, 102–117. [Google Scholar] [CrossRef] [PubMed]
  137. Liu, Y.; Ding, W.; Yu, W.; Zhang, Y.; Ao, X.; Wang, J. Long non-coding RNAs: Biogenesis, functions, and clinical significance in gastric cancer. Mol. Ther.-Oncolytics 2021, 23, 458–476. [Google Scholar] [CrossRef] [PubMed]
  138. Shi, Y.; Liu, J.-B.; Deng, J.; Zou, D.-Z.; Wu, J.-J.; Cao, Y.-H.; Yin, J.; Ma, Y.-S.; Da, F.; Li, W. The role of ceRNA-mediated diagnosis and therapy in hepatocellular carcinoma. Hereditas 2021, 158, 44. [Google Scholar] [CrossRef] [PubMed]
  139. Quinn, J.J.; Chang, H.Y. Unique features of long non-coding RNA biogenesis and function. Nat. Rev. Genet 2016, 17, 47–62. [Google Scholar] [CrossRef] [PubMed]
  140. Tsagakis, I.; Douka, K.; Birds, I.; Aspden, J.L. Long non-coding RNAs in development and disease: Conservation to mechanisms. J. Pathol. Inform. 2020, 250, 480–495. [Google Scholar] [CrossRef] [PubMed]
  141. Sun, B.; Liu, C.; Li, H.; Zhang, L.; Luo, G.; Liang, S.; Lü, M. Research progress on the interactions between long non-coding RNAs and microRNAs in human cancer. Oncol. Lett. 2020, 19, 595–605. [Google Scholar] [CrossRef] [PubMed]
  142. Atef, M.; Amer, A.; Hafez, Y.; Elsebaey, M.; Saber, S.; Abd El-Khalik, S. Long non-coding RNA EGFR-AS1 in colorectal cancer: Potential role in tumorigenesis and survival via miRNA-133b sponge and EGFR/STAT3 axis regulation. J. Biomed. Sci. 2021, 78, 122–129. [Google Scholar] [CrossRef]
  143. Shang, A.; Wang, W.; Gu, C.; Chen, W.; Lu, W.; Sun, Z.; Li, D. Long non-coding RNA CCAT1 promotes colorectal cancer progression by regulating miR-181a-5p expression. Aging 2020, 12, 8301. [Google Scholar] [CrossRef]
  144. Bi, C.; Cui, H.; Fan, H.; Li, L. LncRNA LINC01116 promotes the development of colorectal cancer by targeting miR-9-5p/STMN1. Onco Targets Ther. 2020, 13, 10547–10558. [Google Scholar] [CrossRef] [PubMed]
  145. Li, Q.; Yue, W.; Li, M.; Jiang, Z.; Hou, Z.; Liu, W.; Ma, N.; Gan, W.; Li, Y.; Zhou, T. Downregulating long Non-coding RNAs CTBP1-AS2 inhibits colorectal cancer development by modulating the miR-93-5p/TGF-β/SMAD2/3 pathway. Front. Oncol. 2021, 11, 626620. [Google Scholar] [CrossRef] [PubMed]
  146. Ke, D.; Wang, Q.; Ke, S.; Zou, L.; Wang, Q. Long-non coding RNA SNHG16 supports colon cancer cell growth by modulating miR-302a-3p/AKT axis. Pathol. Oncol. Res. 2020, 26, 1605–1613. [Google Scholar] [CrossRef] [PubMed]
  147. Yu, Z.; Wang, Y.; Deng, J.; Liu, D.; Zhang, L.; Shao, H.; Wang, Z.; Zhu, W.; Zhao, C.; Ke, Q. Long non-coding RNA COL4A2-AS1 facilitates cell proliferation and glycolysis of colorectal cancer cells via miR-20b-5p/hypoxia inducible factor 1 alpha subunit axis. Bioengineered 2021, 12, 6251–6263. [Google Scholar] [CrossRef] [PubMed]
  148. Han, H.; Li, H.; Zhou, J. Long non-coding RNA MIR503HG inhibits the proliferation, migration and invasion of colon cancer cells via miR-107/Par4 axis. Exp. Cell Res. 2020, 395, 112205. [Google Scholar] [CrossRef] [PubMed]
  149. Dong, X.; Yang, Z.; Yang, H.; Li, D.; Qiu, X. Long non-coding RNA MIR4435-2HG promotes colorectal cancer proliferation and metastasis through miR-206/YAP1 axis. Front. Oncol. 2020, 10, 160. [Google Scholar] [CrossRef] [PubMed]
  150. Liu, H.; Li, A.; Sun, Z.; Zhang, J.; Xu, H. Long non-coding RNA NEAT1 promotes colorectal cancer progression by regulating miR-205-5p/VEGFA axis. Hum. Cell 2020, 33, 386–396. [Google Scholar] [CrossRef] [PubMed]
  151. Luo, Y.; Chen, J.-J.; Lv, Q.; Qin, J.; Huang, Y.-Z.; Yu, M.-H.; Zhong, M. Long non-coding RNA NEAT1 promotes colorectal cancer progression by competitively binding miR-34a with SIRT1 and enhancing the Wnt/β-catenin signaling pathway. Cancer Lett. 2019, 440, 11–22. [Google Scholar] [CrossRef] [PubMed]
  152. Shengnan, J.; Dafei, X.; Hua, J.; Sunfu, F.; Xiaowei, W.; Liang, X. Long non-coding RNA HOTAIR as a competitive endogenous RNA to sponge miR-206 to promote colorectal cancer progression by activating CCL2. J. Cancer 2020, 11, 4431. [Google Scholar] [CrossRef]
  153. Li, S.; Wu, T.; Zhang, D.; Sun, X.; Zhang, X. The long non-coding RNA HCG18 promotes the growth and invasion of colorectal cancer cells through sponging miR-1271 and upregulating MTDH/Wnt/β-catenin. Clin. Exp. Pharmacol. Physiol. 2020, 47, 703–712. [Google Scholar] [CrossRef]
  154. Li, X.; Chen, W.; Jia, J.; You, Z.; Hu, C.; Zhuang, Y.; Lin, Z.; Liu, Y.; Yang, C.; Xu, R. The long non-coding RNA-RoR promotes the tumorigenesis of human colorectal cancer by targeting miR-6833-3p through SMC4. Onco Targets Ther. 2020, 13, 2573–2581. [Google Scholar] [CrossRef] [PubMed]
  155. He, C.; Fu, Y.; Chen, Y.; Li, X. Long non-coding RNA SNHG8 promotes autophagy as a ceRNA to upregulate ATG7 by sponging microRNA-588 in colorectal cancer. Oncol. Lett. 2021, 22, 577. [Google Scholar] [CrossRef] [PubMed]
  156. Ruan, Z.; Deng, H.; Liang, M.; Xu, Z.; Lai, M.; Ren, H.; Deng, X.; Su, X. Downregulation of long non-coding RNA MAFG-AS1 represses tumorigenesis of colorectal cancer cells through the microRNA-149-3p-dependent inhibition of HOXB8. Cancer Cell Int. 2020, 20, 511. [Google Scholar] [CrossRef] [PubMed]
  157. Lai, F.; Deng, W.; Fu, C.; Wu, P.; Cao, M.; Tan, S. Long non-coding RNA SNHG6 increases JAK2 expression by targeting the miR-181 family to promote colorectal cancer cell proliferation. J. Gene Med. 2020, 22, e3262. [Google Scholar] [CrossRef] [PubMed]
  158. Zheng, W.; Li, H.; Zhang, H.; Zhang, C.; Zhu, Z.; Liang, H.; Zhou, Y. Long noncoding RNA RHPN1-AS1 promotes colorectal cancer progression via targeting miR-7-5p/OGT axis. Cancer Cell Int. 2020, 20, 54. [Google Scholar] [CrossRef] [PubMed]
  159. Zheng, Y.; Nie, P.; Xu, S. Long noncoding RNA CASC21 exerts an oncogenic role in colorectal cancer through regulating miR-7-5p/YAP1 axis. Biomed. Pharmacother. 2020, 121, 109628. [Google Scholar] [CrossRef]
  160. Yan, F.; Ma, Y.; Liu, L.; Li, L.; Deng, J.; Sun, J. Long noncoding RNA HOXD-AS1 promotes the proliferation, migration, and invasion of colorectal cancer via the miR-526b-3p/CCND1 axis. J. Surg. Res. 2020, 255, 525–535. [Google Scholar] [CrossRef]
  161. Zhu, X.; Bu, F.; Tan, T.; Luo, Q.; Zhu, J.; Lin, K.; Huang, J.; Luo, C.; Zhu, Z. Long noncoding RNA RP11-757G1. 5 sponges miR-139-5p and upregulates YAP1 thereby promoting the proliferation and liver, spleen metastasis of colorectal cancer. J. Exp. Clin. Cancer Res. 2020, 39, 207. [Google Scholar] [CrossRef]
  162. Huang, F.K.; Zheng, C.Y.; Huang, L.K.; Lin, C.Q.; Zhou, J.F.; Wang, J.X. Long non-coding RNA MCF2L-AS1 promotes the aggressiveness of colorectal cancer by sponging miR-874-3p and thereby up-regulating CCNE1. J. Gene Med. 2021, 23, e3285. [Google Scholar] [CrossRef]
  163. Lu, W.; Huang, Z.; Wang, J.; Liu, H. Long non-coding RNA DANCR accelerates colorectal cancer progression via regulating the miR-185-5p/HMGA2 axis. J. Biochem. 2022, 171, 389–398. [Google Scholar] [CrossRef]
  164. Xu, G.; Wang, H.; Yuan, D.; Yao, J.; Meng, L.; Li, K.; Zhang, Y.; Dang, C.; Zhu, K. RUNX1-activated upregulation of lncRNA RNCR3 promotes cell proliferation, invasion, and suppresses apoptosis in colorectal cancer via miR-1301-3p/AKT1 axis in vitro and in vivo. Clin. Transl. Oncol. 2020, 22, 1762–1777. [Google Scholar] [CrossRef] [PubMed]
  165. Li, C.; Pan, B.; Liu, X.; Qin, J.; Wang, X.; He, B.; Pan, Y.; Sun, H.; Xu, T.; Xu, X. Long intergenic non-coding RNA LINC00485 exerts tumor-suppressive activity by regulating miR-581/EDEM1 axis in colorectal cancer. Aging 2021, 13, 3866. [Google Scholar] [CrossRef] [PubMed]
  166. Liu, G.; Zhao, H.; Song, Q.; Li, G.; Lin, S.; Xiong, S. Long non-coding RNA DPP10-AS1 exerts anti-tumor effects on colon cancer via the upregulation of ADCY1 by regulating microRNA-127-3p. Aging 2021, 13, 9748. [Google Scholar] [CrossRef] [PubMed]
  167. Zhu, K.; Wang, Y.; Liu, L.; Li, S.; Yu, W. Long non-coding RNA MBNL1-AS1 regulates proliferation, migration, and invasion of cancer stem cells in colon cancer by interacting with MYL9 via sponging microRNA-412-3p. Clin. Res. Hepatol. Gastroenterol. 2020, 44, 101–114. [Google Scholar] [CrossRef] [PubMed]
  168. Yin, S.L.; Xiao, F.; Liu, Y.F.; Chen, H.; Guo, G.C. Long non-coding RNA FENDRR restrains the aggressiveness of CRC via regulating miR-18a-5p/ING4 axis. J. Cell. Biochem. 2020, 121, 3973–3985. [Google Scholar] [CrossRef] [PubMed]
  169. Dai, W.; Zeng, W.; Lee, D. lncRNA MCM3AP-AS1 inhibits the progression of colorectal cancer via the miR-19a-3p/FOXF2 axis. J. Gene Med. 2021, 23, e3306. [Google Scholar] [CrossRef] [PubMed]
  170. Chu, Q.; Zheng, W.; Su, H.; Zhang, L.; Chang, R.; Gao, W.; Xu, T. A highly conserved circular RNA, circRasGEF1B, enhances antiviral immunity by regulating the miR-21-3p/MITA pathway in lower vertebrates. Virol. J. 2021, 95, e02145-20. [Google Scholar] [CrossRef] [PubMed]
  171. Guo, W.; Zhang, B.; Sun, C.; Duan, H.Q.; Liu, W.X.; Mu, K.; Zhao, L.; Li, H.R.; Dong, Z.Y.; Cui, Q. Circular RNA derived from TIMP2 functions as a competitive endogenous RNA and regulates intervertebral disc degeneration by targeting miR-185-5p and matrix metalloproteinase 2. Int. J. Mol. Med. 2020, 46, 621–632. [Google Scholar] [CrossRef] [PubMed]
  172. Tang, X.; Ren, H.; Guo, M.; Qian, J.; Yang, Y.; Gu, C. Review on circular RNAs and new insights into their roles in cancer. Comput. Struct. Biotechnol. J. 2021, 19, 910–928. [Google Scholar] [CrossRef]
  173. Zhou, W.-Y.; Cai, Z.-R.; Liu, J.; Wang, D.-S.; Ju, H.-Q.; Xu, R.-H. Circular RNA: Metabolism, functions and interactions with proteins. Mol. Cancer 2020, 19, 172. [Google Scholar] [CrossRef]
  174. Huang, Y.; Zhu, Q. Mechanisms regulating abnormal circular RNA biogenesis in cancer. Cancers 2021, 13, 4185. [Google Scholar] [CrossRef] [PubMed]
  175. Pitolli, C.; Marini, A.; Sette, C.; Pagliarini, V. Non-canonical splicing and its implications in brain physiology and cancer. Int. J. Mol. Sci. 2022, 23, 2811. [Google Scholar] [CrossRef] [PubMed]
  176. Xu, B.; Meng, Y.; Jin, Y. RNA structures in alternative splicing and back-splicing. Wiley Interdiscip. Rev. RNA 2021, 12, e1626. [Google Scholar] [CrossRef] [PubMed]
  177. Van Der Steen, N.; Lyu, Y.; Hitzler, A.K.; Becker, A.C.; Seiler, J.; Diederichs, S. The circular RNA landscape of non-small cell lung cancer cells. Cancers 2020, 12, 1091. [Google Scholar] [CrossRef] [PubMed]
  178. Zhao, X.; Zhong, Y.; Wang, X.; Shen, J.; An, W. Advances in circular RNA and its applications. Int. J. Med. Sci. 2022, 19, 975. [Google Scholar] [CrossRef]
  179. Ma, Y.; Zhang, X.; Wang, Y.-Z.; Tian, H.; Xu, S. Research progress of circular RNAs in lung cancer. Cancer Biol. Ther. 2019, 20, 123–129. [Google Scholar] [CrossRef] [PubMed]
  180. Chen, L.-L. The expanding regulatory mechanisms and cellular functions of circular RNAs. Nat. Rev. Mol. Cell Biol. 2020, 21, 475–490. [Google Scholar] [CrossRef] [PubMed]
  181. Robic, A.; Kühn, C. Beyond back splicing, a still poorly explored world: Non-canonical circular RNAs. Genes 2020, 11, 1111. [Google Scholar] [CrossRef]
  182. Huang, A.; Zheng, H.; Wu, Z.; Chen, M.; Huang, Y. Circular RNA-protein interactions: Functions, mechanisms, and identification. Theranostics 2020, 10, 3503. [Google Scholar] [CrossRef]
  183. Yang, Y.; Yujiao, W.; Fang, W.; Linhui, Y.; Ziqi, G.; Zhichen, W.; Zirui, W.; Shengwang, W. The roles of miRNA, lncRNA and circRNA in the development of osteoporosis. Biol. Res. 2020, 53, 40. [Google Scholar] [CrossRef]
  184. Li, T.; Wang, W.C.; McAlister, V.; Zhou, Q.; Zheng, X. Circular RNA in colorectal cancer. J. Cell. Mol. Med. 2021, 25, 3667–3679. [Google Scholar] [CrossRef] [PubMed]
  185. Bhuyan, R.; Bagchi, A. Prediction of the differentially expressed circRNAs to decipher their roles in the onset of human colorectal cancers. Gene 2020, 762, 145035. [Google Scholar] [CrossRef] [PubMed]
  186. Tang, Q.; Hann, S.S. Biological roles and mechanisms of circular RNA in human cancers. Onco Targets Ther. 2020, 13, 2067–2092. [Google Scholar] [CrossRef] [PubMed]
  187. Zhang, X.; Xu, Y.; Yamaguchi, K.; Hu, J.; Zhang, L.; Wang, J.; Tian, J.; Chen, W. Circular RNA circVAPA knockdown suppresses colorectal cancer cell growth process by regulating miR-125a/CREB5 axis. Cancer Cell Int. 2020, 20, 103. [Google Scholar] [CrossRef] [PubMed]
  188. Wang, J.; Luo, J.; Liu, G.; Li, X. Circular RNA hsa_circ_0008285 inhibits colorectal cancer cell proliferation and migration via the miR-382-5p/PTEN axis. Biochem. Biophys. Res. Commun. 2020, 527, 503–510. [Google Scholar] [CrossRef] [PubMed]
  189. Geng, Y.; Zheng, X.; Hu, W.; Wang, Q.; Xu, Y.; He, W.; Wu, C.; Zhu, D.; Wu, C.; Jiang, J. Has_circ_0009361 acts as the sponge of miR-582 to suppress colorectal cancer progression by regulating APC2 expression. Clin. Sci. 2019, 133, 1197–1213. [Google Scholar] [CrossRef] [PubMed]
  190. Chen, P.; Yao, Y.; Yang, N.; Gong, L.; Kong, Y.; Wu, A. Circular RNA circCTNNA1 promotes colorectal cancer progression by sponging miR-149-5p and regulating FOXM1 expression. Cell Death Dis. 2020, 11, 557. [Google Scholar] [CrossRef] [PubMed]
  191. Hao, S.; Qu, R.; Hu, C.; Wang, M.; Li, Y. A circular RNA derived from golgi glycoprotein 1 mRNA regulates KRAS expression and promotes colorectal cancer progression by targeting microRNA-622. Onco Targets Ther. 2020, 13, 12637–12648. [Google Scholar] [CrossRef] [PubMed]
  192. Yan, D.; Liu, W.; Liu, Y.; Zhu, X. Circular RNA circ_0065378 upregulates tumor suppressor candidate 1 by competitively binding with miR-4701-5p to alleviate colorectal cancer progression. J. Gastroenterol. Hepatol. 2022, 37, 1107–1118. [Google Scholar] [CrossRef]
  193. Wang, J.; Zhang, Y.; Song, H.; Yin, H.; Jiang, T.; Xu, Y.; Liu, L.; Wang, H.; Gao, H.; Wang, R. The circular RNA circSPARC enhances the migration and proliferation of colorectal cancer by regulating the JAK/STAT pathway. Mol. Cancer 2021, 20, 81. [Google Scholar] [CrossRef]
  194. Yin, W.; Xu, J.; Li, C.; Dai, X.; Wu, T.; Wen, J. Circular RNA circ_0007142 facilitates colorectal cancer progression by modulating CDC25A expression via miR-122-5p. Onco Targets Ther. 2020, 13, 3689–3701. [Google Scholar] [CrossRef] [PubMed]
  195. Wen, T.; Wu, H.; Zhang, L.; Li, K.; Xiao, X.; Zhang, L.; Zhang, Y. Circular RNA circ_0007142 regulates cell proliferation, apoptosis, migration and invasion via miR-455-5p/SGK1 axis in colorectal cancer. Anti-Cancer Drugs 2021, 32, 22–33. [Google Scholar] [CrossRef] [PubMed]
  196. Xie, L.; Pan, Z. Circular RNA circ_0000467 regulates colorectal cancer development via miR-382-5p/EN2 axis. Bioengineered 2021, 12, 886–897. [Google Scholar] [CrossRef] [PubMed]
  197. Ren, C.; Zhang, Z.; Wang, S.; Zhu, W.; Zheng, P.; Wang, W. Circular RNA hsa_circ_0001178 facilitates the invasion and metastasis of colorectal cancer through upregulating ZEB1 via sponging multiple miRNAs. Biol. Chem. 2020, 401, 487–496. [Google Scholar] [CrossRef] [PubMed]
  198. Xiao, H.; Liu, M. Circular RNA hsa_circ_0053277 promotes the development of colorectal cancer by upregulating matrix metallopeptidase 14 via miR-2467-3p sequestration. J. Cell. Physiol. 2020, 235, 2881–2890. [Google Scholar] [CrossRef] [PubMed]
  199. Wang, X.; Ren, Y.; Ma, S.; Wang, S. Circular RNA 0060745, a novel circRNA, promotes colorectal cancer cell proliferation and metastasis through miR-4736 sponging. Onco Targets Ther. 2020, 13, 1941–1951. [Google Scholar] [CrossRef] [PubMed]
  200. Li, C.; Zhou, H. Circular RNA hsa_circRNA_102209 promotes the growth and metastasis of colorectal cancer through miR-761-mediated Ras and Rab interactor 1 signaling. Cancer Med. 2020, 9, 6710–6725. [Google Scholar] [CrossRef] [PubMed]
  201. Liu, K.; Mou, Y.; Shi, X.; Liu, T.; Chen, Z.; Zuo, X. Circular RNA 100146 promotes colorectal cancer progression by the MicroRNA 149/HMGA2 Axis. Mol. Cell. Biol. 2021, 41, e00445-20. [Google Scholar] [CrossRef] [PubMed]
  202. Yang, L.; Sun, H.; Liu, X.; Chen, J.; Tian, Z.; Xu, J.; Xiang, B.; Qin, B. Circular RNA hsa_circ_0004277 contributes to malignant phenotype of colorectal cancer by sponging miR-512-5p to upregulate the expression of PTMA. J. Cell. Physiol. 2020; ahead of print. [Google Scholar]
  203. Lu, C.; Fu, L.; Qian, X.; Dou, L.; Cang, S. Knockdown of circular RNA circ-FARSA restricts colorectal cancer cell growth through regulation of miR-330-5p/LASP1 axis. Arch. Biochem. Biophys. 2020, 689, 108434. [Google Scholar] [CrossRef]
  204. Hao, Q.; Zhang, Z. hsa_circRNA_000166 facilitated cell growth and limited apoptosis through targeting miR-326/LASP1 axis in colorectal cancer. Gastroent. Res. Pract. 2020, 2020, 8834359. [Google Scholar] [CrossRef]
  205. Yang, B.; Du, K.; Yang, C.; Xiang, L.; Xu, Y.; Cao, C.; Zhang, J.; Liu, W. CircPRMT5 circular RNA promotes proliferation of colorectal cancer through sponging miR-377 to induce E2F3 expression. J. Cell. Mol. Med. 2020, 24, 3431–3437. [Google Scholar] [CrossRef] [PubMed]
  206. Chen, L.-Y.; Wang, L.; Ren, Y.-X.; Pang, Z.; Liu, Y.; Sun, X.-D.; Tu, J.; Zhi, Z.; Qin, Y.; Sun, L.-N. The circular RNA circ-ERBIN promotes growth and metastasis of colorectal cancer by miR-125a-5p and miR-138-5p/4EBP-1 mediated cap-independent HIF-1α translation. Mol. Cancer 2020, 19, 164. [Google Scholar] [CrossRef]
  207. Zhao, J.P.; Chen, L.L. Circular RNA MAT2B induces colorectal cancer proliferation via sponging miR-610, resulting in an increased E2F1 expression. Cancer Manag. Res. 2020, 12, 7107–7116. [Google Scholar] [CrossRef] [PubMed]
  208. Zhang, B.; Yang, S.; Wang, J. Circ_0084615 is an oncogenic circular RNA in colorectal cancer and promotes DNMT3A expression via repressing miR-599. Ppathol. Res. Pract. 2021, 224, 153494. [Google Scholar] [CrossRef]
  209. Zhang, Z.-J.; Zhang, Y.-H.; Qin, X.-J.; Wang, Y.-X.; Fu, J. Circular RNA circDENND4C facilitates proliferation, migration and glycolysis of colorectal cancer cells through miR-760/GLUT1 axis. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 2387–2400. [Google Scholar]
  210. Chen, F.; Guo, L.; Di, J.; Li, M.; Dong, D.; Pei, D. Circular RNA ubiquitin-associated protein 2 enhances autophagy and promotes colorectal cancer progression and metastasis via miR-582-5p/FOXO1 signaling. J. Genet. Genom. 2021, 48, 1091–1103. [Google Scholar] [CrossRef]
  211. Fang, G.; Chen, T.; Mao, R.; Huang, X.; Ji, L. Circular RNA circ_0089153 acts as a competing endogenous RNA to regulate colorectal cancer development by the miR-198/SUMO-specific peptidase 1 (SENP1) axis. Bioengineered 2021, 12, 5664–5678. [Google Scholar] [CrossRef] [PubMed]
  212. Wang, X.; Tao, G.; Huang, D.; Liang, S.; Zheng, D. Circular RNA NOX4 promotes the development of colorectal cancer via the microRNA-485-5p/CKS1B axis. Oncol. Rep. 2020, 44, 2009–2020. [Google Scholar] [CrossRef] [PubMed]
  213. He, J.; Chu, Z.; Lai, W.; Lan, Q.; Zeng, Y.; Lu, D.; Jin, S.; Xu, H.; Su, P.; Yin, D. Circular RNA circHERC4 as a novel oncogenic driver to promote tumor metastasis via the miR-556-5p/CTBP2/E-cadherin axis in colorectal cancer. J. Hematol. Oncol. 2021, 14, 194. [Google Scholar] [CrossRef]
  214. Liu, X.; Qin, Y.; Tang, X.; Wang, Y.; Bian, C.; Zhong, J. Circular RNA circ_0000372 contributes to the proliferation, migration and invasion of colorectal cancer by elevating IL6 expression via sponging miR-495. Anti-Cancer Drugs 2021, 32, 296–305. [Google Scholar] [CrossRef]
  215. Wang, R.; Wang, J.; Chen, Y.; Chen, Y.; Xi, Q.; Sun, L.; Zhang, X.; Zhang, G.; Ding, X.; Shi, T. Circular RNA circLDLR facilitates cancer progression by altering the miR-30a-3p/SOAT1 axis in colorectal cancer. Cell Death Discov. 2022, 8, 314. [Google Scholar] [CrossRef] [PubMed]
  216. Yang, B.-L.; Liu, G.-Q.; Li, P.; Li, X.-H. Circular RNA CUL2 regulates the development of colorectal cancer by modulating apoptosis and autophagy via miR-208a-3p/PPP6C. Aging 2022, 14, 497. [Google Scholar] [CrossRef] [PubMed]
  217. Miao, X.; Xi, Z.; Zhang, Y.; Li, Z.; Huang, L.; Xin, T.; Shen, R.; Wang, T. Circ-SMARCA5 suppresses colorectal cancer progression via downregulating miR-39-3p and upregulating ARID4B. Dig. Liver Dis. 2020, 52, 1494–1502. [Google Scholar] [CrossRef] [PubMed]
  218. Wen, C.; Feng, X.; Yuan, H.; Gong, Y.; Wang, G. Circ_0003266 sponges miR-503-5p to suppress colorectal cancer progression via regulating PDCD4 expression. BMC Cancer 2021, 21, 284. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of activated signaling pathways: (a) the Wnt signaling pathway is activated in the presence of Wnt. When Wnt binds to the Frizzled receptor, it dimerizes with LRP5/6, leading to the recruitment of the Axin/GSK3/APC/CK1 complex to the cell membrane. This results in the translocation of β-catenin from the cytoplasm to the nucleus. In the nucleus, β-catenin interacts with TCF/LEF transcription factors, thereby inducing the expression of downstream target genes. (b) The PI3K/AKT/mTOR signaling pathway is activated when a ligand binds to receptors such as RTK. Activated PI3K phosphorylates PIP2 to convert it to PIP3. Then, PIP3 activates downstream genes, such as AKT and mTOR. (c) The MAPK signaling pathway is initiated when a ligand binds to RTK, leading to the phosphorylation and activation of MAP3K. Activated MAP3K phosphorylates the downstream MAP2K, which in turn phosphorylates MAPK. This phosphorylation cascade transmits signals from the cell surface receptor to the nucleus, influencing gene expression by activating various kinases, enzymes, and transcription factors downstream. (d) The TGF-β signaling pathway is activated when TGF-β directly or indirectly binds to TGF-βR II. TGF-βR II recruits and phosphorylates TGF-βR I, which, in turn, phosphorylates R-SAMD. The phosphorylated R-SMAD then forms a trimeric complex with Co-SMAD. This complex translocates to the nucleus, where it acts as a transcription factor, influencing gene expression. (e) The p53 signaling pathway is initiated when p53 remains stabilized due to its failure to bind with MDM2, an E3 ubiquitin ligase. This stabilization of p53 leads to the induction of downstream gene expression, resulting in the activation of the pathway. Wnt signaling pathway: Wingless/Integrated signaling pathway; PI3K/AKT/mTOR signaling pathway: Phosphatidylinositol 3-kinase/protein kinase B/mammalian target of rapamycin signaling pathway; RTK: receptor tyrosine kinase; PIP2: phosphatidylinositol 4,5-bisphosphate; PIP3: phosphatidylinositol 3,4,5-trisphosphate; MAPK: Mitogen-Activated Protein Kinase; MAP3K: MAP kinase kinase kinase; MAP2K: MAP kinase kinase; TGF-β: Transforming Growth Factor-β; TGF-βR: Transforming Growth Factor-β receptor; R-SAMD: receptor-regulated SMAD; Co-SMAD: common-mediator SMAD; MDM2: mouse double minute 2 homolog.
Figure 1. Schematic diagram of activated signaling pathways: (a) the Wnt signaling pathway is activated in the presence of Wnt. When Wnt binds to the Frizzled receptor, it dimerizes with LRP5/6, leading to the recruitment of the Axin/GSK3/APC/CK1 complex to the cell membrane. This results in the translocation of β-catenin from the cytoplasm to the nucleus. In the nucleus, β-catenin interacts with TCF/LEF transcription factors, thereby inducing the expression of downstream target genes. (b) The PI3K/AKT/mTOR signaling pathway is activated when a ligand binds to receptors such as RTK. Activated PI3K phosphorylates PIP2 to convert it to PIP3. Then, PIP3 activates downstream genes, such as AKT and mTOR. (c) The MAPK signaling pathway is initiated when a ligand binds to RTK, leading to the phosphorylation and activation of MAP3K. Activated MAP3K phosphorylates the downstream MAP2K, which in turn phosphorylates MAPK. This phosphorylation cascade transmits signals from the cell surface receptor to the nucleus, influencing gene expression by activating various kinases, enzymes, and transcription factors downstream. (d) The TGF-β signaling pathway is activated when TGF-β directly or indirectly binds to TGF-βR II. TGF-βR II recruits and phosphorylates TGF-βR I, which, in turn, phosphorylates R-SAMD. The phosphorylated R-SMAD then forms a trimeric complex with Co-SMAD. This complex translocates to the nucleus, where it acts as a transcription factor, influencing gene expression. (e) The p53 signaling pathway is initiated when p53 remains stabilized due to its failure to bind with MDM2, an E3 ubiquitin ligase. This stabilization of p53 leads to the induction of downstream gene expression, resulting in the activation of the pathway. Wnt signaling pathway: Wingless/Integrated signaling pathway; PI3K/AKT/mTOR signaling pathway: Phosphatidylinositol 3-kinase/protein kinase B/mammalian target of rapamycin signaling pathway; RTK: receptor tyrosine kinase; PIP2: phosphatidylinositol 4,5-bisphosphate; PIP3: phosphatidylinositol 3,4,5-trisphosphate; MAPK: Mitogen-Activated Protein Kinase; MAP3K: MAP kinase kinase kinase; MAP2K: MAP kinase kinase; TGF-β: Transforming Growth Factor-β; TGF-βR: Transforming Growth Factor-β receptor; R-SAMD: receptor-regulated SMAD; Co-SMAD: common-mediator SMAD; MDM2: mouse double minute 2 homolog.
Ijms 25 04548 g001
Figure 2. Biogenesis pathways and mutual regulatory relationships of mRNA, miRNA, lncRNA, and circRNA. (a) mRNA generated from a gene undergoes capping, tailing, and splicing to produce mature mRNA. This mature mRNA is regulated directly by miRNA or indirectly by lncRNA or circRNA in the cytoplasm, thereby controlling the translation of the mRNA. (b) CircRNA is generated through back-splicing, intron-pairing-driven circularization, and lariat-driven circularization. Back-splicing involves connecting the 5′ splice donor site downstream to the 3ʹ splice acceptor site upstream, forming a circular structure. In intron-pairing-driven circularization, circRNA is formed by the binding of inverted complementary sequences on the sides of long intron base pairs. Lariat-driven circularization occurs when the excision process of regular splicing fails, resulting in the generation of a circRNA. CircRNA binds to miRNA, indirectly influencing gene expression by preventing miRNA from binding to target genes. (c) Pri-miRNA is transcribed from the miRNA gene and is cleaved by Drosha to produce pre-miRNA. The pre-miRNA is then transported to the cytoplasm and processed to form a miRNA duplex by Dicer. One of the strands binds with AGO to form the RISC, which subsequently interacts with the target mRNA, leading to an inhibition of translation. (d) LncRNA is generated from DNA similarly to mRNA, undergoing capping, splicing, and tailing. Once generated, lncRNA in the cytoplasm binds to miRNA, indirectly affecting gene expression by inhibiting miRNA from binding to target genes. mRNA: messenger RNA; miRNA: microRNA; lncRNA: long non-coding RNA; circRNA: circular RNA; pri-miRNA: primary miRNA; pre-miRNA: precursor miRNA; AGO: Argonaute; RISC: RNA-induced silencing complex.
Figure 2. Biogenesis pathways and mutual regulatory relationships of mRNA, miRNA, lncRNA, and circRNA. (a) mRNA generated from a gene undergoes capping, tailing, and splicing to produce mature mRNA. This mature mRNA is regulated directly by miRNA or indirectly by lncRNA or circRNA in the cytoplasm, thereby controlling the translation of the mRNA. (b) CircRNA is generated through back-splicing, intron-pairing-driven circularization, and lariat-driven circularization. Back-splicing involves connecting the 5′ splice donor site downstream to the 3ʹ splice acceptor site upstream, forming a circular structure. In intron-pairing-driven circularization, circRNA is formed by the binding of inverted complementary sequences on the sides of long intron base pairs. Lariat-driven circularization occurs when the excision process of regular splicing fails, resulting in the generation of a circRNA. CircRNA binds to miRNA, indirectly influencing gene expression by preventing miRNA from binding to target genes. (c) Pri-miRNA is transcribed from the miRNA gene and is cleaved by Drosha to produce pre-miRNA. The pre-miRNA is then transported to the cytoplasm and processed to form a miRNA duplex by Dicer. One of the strands binds with AGO to form the RISC, which subsequently interacts with the target mRNA, leading to an inhibition of translation. (d) LncRNA is generated from DNA similarly to mRNA, undergoing capping, splicing, and tailing. Once generated, lncRNA in the cytoplasm binds to miRNA, indirectly affecting gene expression by inhibiting miRNA from binding to target genes. mRNA: messenger RNA; miRNA: microRNA; lncRNA: long non-coding RNA; circRNA: circular RNA; pri-miRNA: primary miRNA; pre-miRNA: precursor miRNA; AGO: Argonaute; RISC: RNA-induced silencing complex.
Ijms 25 04548 g002
Figure 3. The regulatory mechanisms of ceRNA/miRNA/mRNA to enhance CRC progression or metastasis: (a) Upregulated ceRNA binds to miRNA, interfering with the binding between miRNA and target mRNA. This interference results in increased mRNA expression, promoting CRC progression or metastasis. (b) Downregulated ceRNA induces an increase in miRNA expression, leading to the binding to mRNA, which acts as an inhibitor of signaling pathways. Consequently, this interaction induces a decrease in mRNA expression, promoting the progression or metastasis of CRC. ceRNA: competing endogenous RNA; miRNA: microRNA; mRNA: messenger RNA; CRC: colorectal cancer.
Figure 3. The regulatory mechanisms of ceRNA/miRNA/mRNA to enhance CRC progression or metastasis: (a) Upregulated ceRNA binds to miRNA, interfering with the binding between miRNA and target mRNA. This interference results in increased mRNA expression, promoting CRC progression or metastasis. (b) Downregulated ceRNA induces an increase in miRNA expression, leading to the binding to mRNA, which acts as an inhibitor of signaling pathways. Consequently, this interaction induces a decrease in mRNA expression, promoting the progression or metastasis of CRC. ceRNA: competing endogenous RNA; miRNA: microRNA; mRNA: messenger RNA; CRC: colorectal cancer.
Ijms 25 04548 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lee, Y.J.; Kim, W.R.; Park, E.G.; Lee, D.H.; Kim, J.-m.; Shin, H.J.; Jeong, H.-s.; Roh, H.-Y.; Kim, H.-S. Exploring the Key Signaling Pathways and ncRNAs in Colorectal Cancer. Int. J. Mol. Sci. 2024, 25, 4548. https://doi.org/10.3390/ijms25084548

AMA Style

Lee YJ, Kim WR, Park EG, Lee DH, Kim J-m, Shin HJ, Jeong H-s, Roh H-Y, Kim H-S. Exploring the Key Signaling Pathways and ncRNAs in Colorectal Cancer. International Journal of Molecular Sciences. 2024; 25(8):4548. https://doi.org/10.3390/ijms25084548

Chicago/Turabian Style

Lee, Yun Ju, Woo Ryung Kim, Eun Gyung Park, Du Hyeong Lee, Jung-min Kim, Hae Jin Shin, Hyeon-su Jeong, Hyun-Young Roh, and Heui-Soo Kim. 2024. "Exploring the Key Signaling Pathways and ncRNAs in Colorectal Cancer" International Journal of Molecular Sciences 25, no. 8: 4548. https://doi.org/10.3390/ijms25084548

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