Next Article in Journal
Molecular Aspects of the Development and Function of Auditory Neurons
Next Article in Special Issue
Current and Prospective Methods for Assessing Anti-Tumor Immunity in Colorectal Cancer
Previous Article in Journal
N-Acetyl-d-Glucosamine Kinase Interacts with NudC and Lis1 in Dynein Motor Complex and Promotes Cell Migration
Previous Article in Special Issue
Dehydroxyhispolon Methyl Ether, A Hispolon Derivative, Inhibits WNT/β-Catenin Signaling to Elicit Human Colorectal Carcinoma Cell Apoptosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms of Colon Cancer Progression and Metastasis: Recent Insights and Advancements

1
Department of Biomedical Science, College of Health Sciences, QU Health, Qatar University, Doha 2713, Qatar
2
College of Medicine, QU Health, Qatar University, Doha 2713, Qatar
3
Biomedical and Pharmaceutical Research Unit, QU Health, Qatar University, Doha 2713, Qatar
4
Biomedical Research Center, QU Health, Qatar University, Doha 2713, Qatar
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(1), 130; https://doi.org/10.3390/ijms22010130
Submission received: 21 November 2020 / Revised: 9 December 2020 / Accepted: 10 December 2020 / Published: 24 December 2020
(This article belongs to the Special Issue Molecular and Translational Research on Colorectal Cancer 2.0)

Abstract

:
Colorectal cancer (CRC), the third most common type of cancer, is the second leading cause of cancer-related mortality rates worldwide. Although modern research was able to shed light on the pathogenesis of CRC and provide enhanced screening strategies, the prevalence of CRC is still on the rise. Studies showed several cellular signaling pathways dysregulated in CRC, leading to the onset of malignant phenotypes. Therefore, analyzing signaling pathways involved in CRC metastasis is necessary to elucidate the underlying mechanism of CRC progression and pharmacotherapy. This review focused on target genes as well as various cellular signaling pathways including Wnt/β-catenin, p53, TGF-β/SMAD, NF-κB, Notch, VEGF, and JAKs/STAT3, which are associated with CRC progression and metastasis. Additionally, alternations in methylation patterns in relation with signaling pathways involved in regulating various cellular mechanisms such as cell cycle, transcription, apoptosis, and angiogenesis as well as invasion and metastasis were also reviewed. To date, understanding the genomic and epigenomic instability has identified candidate biomarkers that are validated for routine clinical use in CRC management. Nevertheless, better understanding of the onset and progression of CRC can aid in the development of early detection molecular markers and risk stratification methods to improve the clinical care of CRC patients.

1. Introduction

Colorectal cancers (CRCs) encompass two types of highly aggressive and common types of cancers, namely, colon and rectal. Globally, while colon cancer is the fourth most common malignancy, rectum cancer is the eight most common one [1]. Collectively, CRCs present the third most commonly diagnosed form of cancer worldwide, accounting for 11% of all diagnosed cancer cases [1]. Additionally, CRC is the second most lethal cancer worldwide [2]. In this context, the prevalence of CRC cases varies from one geographical location to another, as Hungary has the highest incidence of CRC among males and Norway has the highest incidence amongst females [2]. Nevertheless, CRC is frequently diagnosed in men from Japan, South Korea, the Middle East region, and Slovakia, with mortality being highest in Saudi Arabia, Oman, and UAE [2].
Two molecular pathological classifications for CRC are described [3,4]. The Cancer Genome Atlas (TCGA) project first classified CRC into two groups using integrated molecular analysis (array-based and sequencing technologies) [3]. The first group is comprised of approximately hypermutated tumors (~16%) with either microsatellite instability (MSI) due to defective mismatch repair (MMR) (∼13%) or ultra-mutated tumors with DNA Polymerase Epsilon or Delta 1 (POLE or POLD1) exonuclease domain (proofreading) mutations (EDM) (∼3%) [3]. The second group consisted of non-hypermutated tumors (~84%), microsatellite stable (MSS) cancers with a high frequency of DNA somatic copy number alterations (SCNAs) and dysregulated Wnt pathway with frequent mutations in genes including Adenomatous polyposis coli (APC), Kirsten ras (KRAS), Phosphatidylinositol-4,5-Bisphosphate 3-Kinase Catalytic Subunit Alpha (PIK3CA), Small mothers against decapentaplegic 4 (SMAD4), and Tumor protein 53 (TP53) [3].
On the other hand, Guinney et al. (2016) [4] aggregated gene expression datasets using 18 different CRC and described the four consensus molecular subtypes (CMS) of CRC: CMS1 (MSI-immune), CMS2 (canonical), CMS3 (metabolic), and CMS4 (mesenchymal) (Table 1).
To determine the extent of CRC spread, according to the American Joint Committee on Cancer, CRC staging is done according to the tumor-nodes-metastasis (TNM) system [5].
CRC pathogenesis is highly complex and diverse and is induced by several risk factors including sporadic, familial, and inherited [6]. Sporadic cases comprise 70% of CRC cases and are comprised of environmental and dietary factors (cigarette smoking, excessive alcohol consumption, sedentary lifestyle, obesity, and diets high in fat and low in fiber) [6,7]. Familial CRC cases comprise 25% of the cases and affect individuals with family history of CRC [6]. Lastly, genetic or inherited cases account for only 5–10% of the cases and are categorized based on the presence or absence of colonic polyps [6]. While diseases with polyposis include familial adenomatous polyposis (FAP) (OMIM#: 608456), MUTYH-associated polyposis (MAP) (OMIM#: 175100), hamartomatous polyposis syndromes (Peutz-Jeghers, juvenile polyposis (OMIM#: 175200) [8] and Cowden syndrome (OMIM#: 158350), diseases without polyposis are known as hereditary nonpolyposis CRC (HNPCC; Lynch syndrome (OMIM#: 120435)) [6]. Other risk factors involve the presence of existing diseases such long-standing inflammatory bowel diseases (IBD), either Crohn or ulcerative colitis [9].
On the other hand, genetic and epigenetic alterations are involved in events that lead to the initiation of neoplastic transformation of healthy epithelium that progresses toward the malignant stages; in 1990, Fearon and Vogelstein defined the genetic paradigm of CRC formation [10]. In their multistep genetic model, formation of CRC occurs due to accumulation of several genetic and epigenetic alterations in key genes involved in silencing of TSGs and activation of oncogenes [10,11,12]. In this regard, two major pathways were defined for CRC formation. One pathway involves inhibiting tumor suppressor genes’ (TSGs) expression and adenomatous polyposis coli (APC). This comprises 85% of all CRC and is mutated in the germline of patients with FAP. The other pathway is via mutational inactivation of proteins involved in MMR (MSH2, MLH1, and PMS2) and accounts for around 15% of all sporadic cases and HNPCC syndrome [10,11,12]. The different pathways are characterized by distinctive models of genetic instability, subsequent clinical manifestations, and pathological characteristics [11,12]. However, the majority of CRC cases follow the chromosomal instability (CIN) pathway, characterized by an extensive loss of heterozygosis (LOH) and gross chromosomal abnormalities [13,14].
Today, CRC is one of the leading causes of morbidity and represents a formidable health burden. Therefore, understanding the molecular and genetic features of its onset and progression is essential. This review focused on several novel cellular signaling pathways associated with CRC metastasis, including Wnt/β-catenin, p53, cyclooxygenase (COX), TGF-β/SMAD, nuclear factor kappa beta (NF-κB), Notch, vascular epidermal growth factor (VEGF), and Janus kinase/signal transducer and activator of transcription 3 (JAKs/STAT3) pathways. Additionally, targeted therapeutic agents developed based on these pathways are also briefly discussed.

2. Molecular Basis of Colorectal Cancer

As mentioned above, both genetic and epigenetic alterations are involved in regulating tumorigenesis of CRC [10]. The adenocarcinoma sequence was described much earlier, in 1980, when the transformation of normal colorectal epithelium to an adenoma and, finally, to an invasive and metastatic tumor was elucidated [10]. There are three major pathways involved in the genetic instability of CRC and its pathogenesis: chromosomal instability (CIN), microsatellite instability (MSI), and CpG island methylator phenotype (CIMP) pathways (Figure 1) [15].

2.1. Chromosomal Instability (CIN) Pathway

Chromosomal instability refers to a significant increase in the gain or loss of either the entire or the large portions of chromosomes and is the most commonly occurring genetic instability in CRC. CIN is found in around 85% of adenocarcinoma transitions [16,17] and is characterized by the activation of oncogenes (KRAS and BRAF), inactivation of TSGs (APC and TP53), and a loss of heterozygosity for the long arm of chromosome 18 (18q LOH), thus, promoting CRC tumorigenesis [15,16].
According to the multistep genetic model defined by Fearon and Vogelstein [10], the first step includes the silencing of APC, followed by oncogenic KRAS mutations in the adenomatous stage, and, finally, deletion of chromosome 18q and inactivation of TP53 occurring during the transition to malignancy (Figure 1) [10,18,19]. In addition, recently, genetic aberrations in TGF-βR and PI3KCA were found to be involved in the adenocarcinoma sequence model [20,21].
Furthermore, array-based comparative genomic hybridization and single nucleotide polymorphism techniques have identified a few frequently recurring allele losses of all chromosomal arms in CRC that include losses at chromosomal arms 1p, 5q, 8p, 17p, 18p, 18q, 20p, and 22q [3,22,23,24]. On the other hand, chromosomal gains were identified at chromosome 7 and chromosomal arms 1q, 8q, 12q, 13q, and 20q [3,22,23,24]. The allelic loss or gain of material denotes presence of candidate TSGs or oncogenes, respectively, thus allowing growth of mutated cells, leading to transformation of normal cells into cancerous [25].

2.1.1. Adenomatous Polyposis Coli (APC) Gene and Wnt Signaling Pathway

The APC gene, located on chromosome 5q21-q22, consists of 8535 nucleotides spanning 21 exons and encodes a 310 kDa protein [26]. Around 75% of the coding sequence is present on exon 15, the most frequent region for both germline and somatic mutations of APC [26]. Furthermore, APC is a multi-domain protein. From the N to C-terminus, it consists of oligomerization, armadillo repeat, 15- or 20-residue repeat, SAMP repeats, a basic domain, and C-terminal domains [27]. Since APC interacts with various binding proteins through its different domains, APC plays a role in regulating cellular processes including chromosome segregation, cell migration, apoptosis, adhesion, proliferation, and differentiation [28,29,30,31,32].
APC mutations are present at the preliminary stages of neoplasia and are majorly linked with the classic tubular adenoma pathway and CIN cancers [33,34,35,36]. APC germline mutations give rise to familial adenomatous polyposis (FAP) syndrome (OMIM#: 608456), attenuated FAP, Gardner syndrome (OMIM#: 175100), Turcot syndrome (OMIM#: 276300), and other major hereditary predisposition events leading to CRC development [37]. FAP, an autosomal-dominant genetic disorder, is characterized by the development of around 1000 precancerous colon polyps in the large bowel in individuals aged 10–25 years. Germline mutation is found in 60–80% of families with FAP [38,39]. CIN is also known to stimulate tumorigenesis in FAP [40,41,42,43]. The attenuated form of FAP (AFAP) is marked by less than 100 adenomas or polyps in the more proximal part of the colon. Germline mutations occur in the 5′ or 3′ region of the APC gene [44]. More importantly, 16–40% of patients with less than 100 polyps present the biallelic inactivation of the MUTYH-based excision repair gene, a condition called MUTYH-associated polyposis (MAP) [45]. AFAP and MAP are phenotypically very alike [45]. While Gardner syndrome (OMIM#: 175100) is the subtype of FAP and is linked with osteomas and soft tissue tumors, Turcot syndrome (OMIM#: 276300) is characterized by colonic polyps along with tumors of the central nervous system [44]. On the other hand, somatic APC mutations (frameshift or nonsense mutations) are present in more than 70–80% of sporadic CRCs and 5q LOH is reported in approximately 40% of CRC cases [46,47,48].
In both familial and sporadic CRCs, the APC/β-catenin/Wnt-Tcf pathway plays a major role in the onset and progression of CRC carcinogenesis [49]. The APC gene inhibits transition from the G0/G1 to the S phase of the cell cycle. The Wnt signaling pathway maintains the undifferentiated stem cells in the base of the colonic crypts, allowing survival of both normal and cancer stem cells [32]. In this regard, it is well known that β-catenin is the main controller of the Wnt signaling pathway [50]. The wild-type APC protein negatively controls the Wnt signaling by regulating the ubiquitin-mediated proteasomal breakdown of the transcription factor β-catenin [51,52]. Disruption of the APC protein results in enhanced Wnt signaling by intracellular β-catenin stabilization, which stimulates transcription of Wnt targeted genes and enhances TCF targets with increase in cell growth, differentiation, spread, and adhesion of colorectal cells [53]. Mutations in genes involved in APC/β-catenin/Tcf pathway in CRC cells without APC mutations are also present in sporadic CIN tumors. Activating mutations in the gene for β-catenin (CTNNB1) block APC-regulated breakdown and are present in colorectal neoplasia. Although CTNNB1 mutations are more common in adenomas (12.5%) than invasive cancer (1.4%), they are found in the preliminary stages of CRC pathogenesis and plausibly substitute APC mutations in cancer onset and progression [54,55,56]. In addition, distinct units of the APC/β-catenin/Wnt pathway can be either directly or indirectly changed, by constitutively triggering β-catenin or Tcf. Among the different regulatory genes that act with APC, the mitotic checkpoint protein BubR1 plays a vital role [28]. BubR1 is a part of the mitotic checkpoint machinery along with Bub1, Bub3, Mad1, Mad2, Mad3, Mps-1, CENP-E, and cell division cycle 20 (CDC20) [57]. Binding of BubR1 to Cdc20 blocks APC activity by triggering a “wait anaphase” signal [58,59], thus contributing to the development of polyploid cells, extended cell survival, and uncontrolled cell proliferation, suggesting a plausible pathogenic mechanism in the initiation of CIN in CRC sporadic forms [60].
Recently, activation of leucine-rich repeat-containing G-protein-coupled receptors (LGR-4 and LGR-5) triggered signaling by binding with proteins in the R-responding family [61,62,63,64]. Moreover, cyclin D1 (CCND1) was implicated in APC signaling: Mutated APC cells activate downstream targets, such as cyclin D1 and Myc [65,66]. CCND1, along with other cyclin-dependent kinases (CDKs) that block cyclins, such as p27 (CDKN1B) and p21 (CDKN1A), are vital for cell growth and apoptosis as well as cell cycle control, majorly during the transition from G1 to S phase [67]. Prolonged activation of CCND1 by APC mutation leads to the onset of colonic neoplasia by allowing the cell to divert from apoptosis [68]. Arber and colleagues (1997) studied the expression of CCND1 in normal colonic mucosa, adenoma, and adenocarcinoma and confirmed its high expression in colonic tumors, thus indicating CCND1 expression as an early event during multistage process of CRC tumorigenesis that may deregulate cell-cycle control in benign adenomas and stimulate tumor progression [69].
On the other hand, β-catenin activity can be indirectly triggered by aberrations in oncogenes controlling its activity at different levels. β-catenin mutually interacts with various members of the Notch pathway that are vital regulators of cell differentiation and play a role in CRC carcinogenesis [70,71]. Kwon and colleagues showed that Notch1 stimulates the assembly of active β-catenin protein in the absence of ligand-receptor activation [72]. Further genetic changes that modulate β-catenin activity include CDK8 (cyclin-dependent kinase-8) gene amplification and occurs in over 60% of CRC cases. Increased CDK8 enhances both β-catenin [73] and Notch1, thus stimulating transcription and cell differentiation [73,74].

2.1.2. TP53 Pathway

The TP53 gene, located on the short arm of chromosome 17, is known as the “guardian of the genome” and encodes proteins regulating cell cycle, DNA repair, senescence, and apoptosis [75]. TP53 mutations or loss of function are reported in 50–75% of CRC cases [76]; loss of p53-mediated pathways of apoptosis is an important determinant of progression from adenoma to malignant tumor [77]. The p53 loss of function enhances high cell proliferation activities and uncontrolled cell cycle, a key step in colorectal carcinogenesis [77]. Research showed missense mutations (48%) that substitute AT for GC as the most common TP53 mutations in CRC [78,79], followed by point mutations (37.5%) with transitions at CpG sites [79].
Commonly, p53 is considered a controller for BubR1 transcription and expression; loss of p53 expression downregulates BubR1, thus comprising checkpoint function to mitotic aberrations leading to progression of CRC [80]. Additionally, a wild type of p53 is identified as a direct activator for WAD-1, a gene highly induced to suppress tumor cell growth in the p53 pathway [81].
Moreover, mutations in CDK are likely to occur during CRC development and progression [82]. Adenosine monophosphate-activated protein kinase (AMPK) pathway induces phosphorylation of p53 and induces AMPK-dependent cell-cycle arrest that upregulates the cyclin-dependent kinases inhibitor 1A (CDKN1A or p21). This eventually controls cell cycle regulation, cellular senescence, and stem cell aging [83]. Furthermore, in CRC, the stimulation of p21 occurs in a p53-dependent pathway; p21 also inhibits the activity of cyclin D1 [84]. Loss of p21 is significantly associated with poor prognosis in CRC [84]. Another type of CDK associated with TP53 loss of function is CDK inhibitor 1B (CDKN1B or p27) [85]. Specifically, p27 is an enzyme inhibitor that encodes CDK inhibitor proteins responsible for regulating cell cycle progression into S phase and its degradation [85]. It has been established that p27 expression is inversely related with the MSI-H and CIMP-H types of CRC and TP53-negative cancers [86].
The p53 also interacts with cyclooxygenase-2 (COX-2) and is involved in enhancing inflammation and CRC cell proliferation [87]. Interestingly, COX-2-positive tumors are significantly linked with cancer-specific mortality regardless of p53 status, thus suggesting COX-2 as an independent CRC prognostic factor [88,89].

2.1.3. The 18q Loss of Heterozygosity (LOH)

Loss of heterozygosity (LOH) refers to the absence of one of the two copies or alleles of a gene, with the remaining allele frequently being affected by mutation [44]. LOH in the 18q region is most commonly observed in advanced CRC, accounting for approximately 70% of the cases [90], and is associated with poor prognosis in CRC [91,92]. LOH at 18q indicates presence of several TSGs including Deleted in Colorectal Carcinoma (DCC), SMAD2, and SMAD4; loss of expression of 18q LOH plays a significant role in CRC pathogenesis [19].
DCC, located on chromosome 18q21.2, encodes netrin-1 and is indicated as a plausible TSG [93]; LOH in the DCC gene region is present in approximately 70% of CRCs. Moreover, a few somatic mutations in DCC are found in CRC [94]. Netrin-1 is built within the cryptos of colorectal mucosa; epithelial cell differentiation results in loss of netrin-1 expression [95]. Furthermore, mutation in DCC gene inhibits binding of netrin-1 to DCC transmembrane protein, leading to abnormal cell survival [95].
On the other hand, SMAD2 and SMAD4 are present on 18q21.1, the prevalent region lost during CRC progression, and correlate with adenoma development and adenocarcinoma progression in mice models, thus suggesting a plausible tumor suppressor role for SMAD genes [19,96]. Furthermore, immunohistochemical analysis reported a loss of SMAD4 expression in >50% of CRCs, which is associated with lymph node metastases [97]. Since the frequency of somatic mutations in SMAD2 and SMAD4 is comparatively low in CRC [98], other TSGs might be responsible for chromosome 18q loss. Research has indicated that SMAD genes encode for TGF-β [99]. Dysregulation of TGF-β signaling occurs in the majority of CRCs [36]. Additionally, inactivating mutations in receptor genes including TGF-βR1, TGF-βR2, and TGF-β superfamily members Activin Receptor type 2 (ACVR2) are reported in CRC [100]. Functionally, marked mutations in TGF-βR2 are present in ~30% of all CRC cases and correlate with malignant transformation of late adenomas. TGF-βR2 mutations are most frequent in MSI tumors; however, they are also present in around 15% of MSS tumors [12,20,101].

2.2. The Microsatellite Instability Pathway

Another type of genomic instability in CRC is microsatellite instability (MSI), a distinctive characteristic of cancerous cells [102]. MSI is the hallmark of HNPCC or Lynch syndrome and occurs in >95% of HNPCC cases [103]. However, in the majority of sporadic CRCs, the underlying mechanism for CIN remains nascent and MSI comprises merely 15–20% of all CRC cases [103].
A subset of tumors with unstable loci in ≥30% markers are defined as “Microsatellite high” (MSI-H), a subset of tumors with 10–29% unstable loci are classified “Microsatellite low” (MSI-L), and “microsatellite stable” (MSS) is marked with no unstable markers [104]. In cancers with MSI-H, small insertions/deletions result in frameshift mutations within repetitive tracts in the coding regions of TSGs or oncogenes, further contributing to tumorigenesis [105]. Mori et al. [106] performed large-scale genomic screening of the coding region of microsatellites and found mutations in nine loci (TGF-βR2, Bax, MSH3, ActRIIB, SEC63, AIM2, NADH-ubiquinone oxidoreductase, COBLL1, and EBP1) in >20% of tumors. TGF-βR2 was the most commonly mutated loci and instability in the poly-adenine tract of TGF-βR2 is present in approximately 85% of MSI-H CRCs [107]. Moreover, Bax, the other frequently mutated gene, was found to have frameshift mutations within the poly-guanine sequence in almost 50% of the MSI-H CRCs, resulting in the inactivation of Bax and inhibition of apoptosis [108].
MSI is rarely found in polyps, except in Lynch syndrome [109,110]. Furthermore, individuals with Lynch syndrome frequently develop MSI CRCs due to germline mutations in one of the MMR genes (MLH1, MSH2, MSH6, and PMS2); mutations in MLH1 or MSH2 gene lead to an increased risk (70–80%) of developing cancer, while mutations in the MSH6 or PMS2 gene have a comparatively lower risk (25–60%) of cancer development [111] On the contrary, sporadic MSI CRCs frequently display loss of MMR activity due to MLH1 silencing by aberrant DNA methylation [112,113].
Furthermore, according to modeling studies and absence of colon cancer in individuals with biallelic germline mutations in MMR genes suggest that lack of MMR activity is insufficient to trigger polyp formation [114,115]. Relevant to its clinical impact, there is considerable indirect data that polyps arising as a result of MMR activity loss has a lesser transition interval from a polyp to colorectal cancer; polyps can develop into MSI CRCs in as less than 2–3 years [116]. Evidence suggesting lack of MMR progress and onset of MSI induces tumor development and progression is based on the fact that MSI is generally observed in polyps adjacent to cancers and is infrequent in non-advanced polyps [116]. It is known that sporadic MSI CRCs correlate with the serrated neoplasia pathway and commonly carry BRAFV600E mutations, on the contrary, Lynch syndrome arises from MMR genes germline mutations and lacks mutated BRAF [117,118].
Clinically, BRAF-mutated CRC correlates with poor prognosis and overall survival (OS) in comparison to BRAF wild-type disease [119]. Furthermore, a study showed that BRAF-mutated CRC patients had worst OS as compared to patients carrying RAS (KRAS and NRAS) mutations [120]. Also, mutation in BRAF is a negative prognostic factor in stage II and III disease [121]. On the contrary, a recent study, performed meta-analysis in 1164 MSI-H non-metastatic CRC patients and showed that BRAF V600E mutation is associated with worst OS, but not disease recurrence [122]. Moreover, another meta-analysis in patients undergoing resection of liver metastasis showed that following metastasectomy, OS was worst for BRAF-mutated metastatic CRC [123]. It has also been revealed that, non-V600E BRAF-mutated (BRAF codons 594 and 596) CRCs have better prognosis as compared to BRAF-mutated CRCs; BRAF 594 and 496 tumors are microsatellite stable, rectal, non-mucinous with no peritoneal spread and have a significantly longer OS as compared to V600E BRAF-mutated tumors [124]. Similarly, other studies in CRC patients showed, in comparison to V600E BRAF-mutated tumors, non-V600E BRAF-mutated tumors are present in the younger population, lower grade tumors and the median OS is significantly longer compared with both V600E BRAF-mutant and BRAF wild-type patients [125,126].

2.3. CpG Island Methylator Phenotype (CIMP) Pathway

DNA methylation is the addition of a methyl group to cytosine in the 5′-position that is catalyzed by DNA methyltransferases via covalent linkage within a CG dinucleotide sequence within the promoter region, termed CpG transcription [127,128]. In normal cells, the majority of the CpG sites are heavily methylated while CpG islands, usually located in the promoter regions of genes, are unmethylated. However, following cancer initiation, hypermethylation within the promoter region may lead to inactivation of tumor-suppressor genes, while global hypomethylation is associated with genomic instability and chromosomal aberrations [129]. Epigenetic instability in CRC is demonstrated as hypermethylation of loci that contain CpG islands and this is usually accompanied by global DNA hypomethylation. Alterations in methylation pattern can affect virtually all signaling pathways, including TP53, TGFβ/SMAD, Wnt, NOTCH and receptor tyrosine kinases involved in cell cycle regulation, transcription regulation, DNA stability, apoptosis, cell-cell adhesion, angiogenesis, cell invasion and metastasis [130,131,132].
Many genes are identified to be methylated and silenced in CRC, some commonly methylated ones include APC, MLH1, MGMT, SFRP1, SFRP2, CDKN2A, TIMP3, VIM, SEPT, CDH1 and HLTF. Additionally, there is a distinct subset of CRCs, known as the CpG island methylator phenotype (CIMP) [133]; CIMP tumors frequently carry BRAFV600E mutations [134]. CIMP is further subclassified based on integrated genetic and epigenetic instability into CIMP2, CIMP-low, and CIMP-high [135,136]. DNA methylation profiling revealed that approximately 20% of CRCs are CIMP tumors; CIMP tumors significantly correlated with age, female sex, proximal colon location, as well as MSI, KRAS and BRAF mutations [87,137]. The most commonly used CIMP markers are MLH1, p16, MINT1, MINT2 and MINT31. Additional markers include CACNA1G, CRABP1, IGF2, NEUROG1, RUNX3, SOCS1, HIC1 and IGFBP3 for positive CIMP identification [138,139].
Although, upregulated expression of the DNA methyltransferases (DNMT3B or DNMT1) is associated with CIMP, the underlying mechanism(s) that promote CIMP are still unknown [140]. One of the plausible underlying mechanisms is based on the silencing of barriers that inhibit methylation of normally unmethylated CpG islands [141,142]. The other suggested mechanism is alterations in the chromatin structure and histone modification state of histone H3 lead to the detection of aberrant DNA methylation in loci that obtain this alteration [143,144]. While, PTEN, a TSG, shows reduced methylation rates, TWIST1 gene is silenced by promoter methylation in CRC. Aside from all of the above-mentioned mechanisms of DNA hypermethylation, global DNA hypomethylation frequently occurs at repetitive sequences including LINE-1 repeats, retrotransposons, introns and gene deserts [133,145].

EGFR-KRAS-BRAF Pathway

The epidermal growth factor receptor (EGFR) belongs to the ErbB/HER family and consists of four members; ErbB1 (EGFR/HER1), ErbB2 (Neu/HER2), ErbB3 (HER3), and ErbB4 (HER4) [146,147]. Activation of EGFR pathway, triggers several downstream intracellular signaling pathways, including the RAS/RAF/MEK/ERK, PI3K/AKT, and JAK/STAT3 pathways to regulate cell growth, survival, and migration [148,149,150]. Deregulated EGFR expression is present in various cancers including CRC [146]; increased EGFR expression is present in 25–77% of CRC cancers [151,152]. Activation of EGFR induces RAS-RAF activation, which leads to phosphorylation of mitogen-activated protein kinase (MAPK or MEK) and activation of extracellular signal-related kinase (ERK) [153,154].
The MAPK pathway includes KRAS and BRAF; the RAS/RAF/MAPK pathway is involved in regulating cell proliferation, differentiation, apoptosis and senescence [153]. Activation of MAPK includes RAS, RAF and MEK; RAS stimulates the signaling cascade via the phosphoinositol kinases (PI3K) as well as RAF [155,156]. PI3K activation inhibits apoptosis, activation of RAS provokes cellular proliferation, thus, promoting cell survival and tumor invasion and metastasis [156,157]. KRAS mutations include codons 12 and 13 on exon 2 and codon 61 on exon 3; codon 12 being the most frequently affected through missense mutations [158] including substitution of glycine for aspartate (p.G12D and p.G13D) [159], of which p.G13D account for 58% of the cases [160]. KRAS along with NRAS and HRAS are oncogenes belonging to the RAS family [161]. KRAS is commonly mutated in sporadic CRCs (35–45%) [158,162,163] and is associated with poor prognosis [164,165,166]; according to the adenocarcinoma sequence, KRAS mutations occur after APC mutations [10].
On the other hand, BRAF, a member of RAF family of serine/threonine kinases regulates cellular responses to growth signals through the RAS-RAF-MAP kinase pathway [167]. Activating mutations in BRAF are found in approximately 5–10% of metastatic CRC, however, they are rare in Lynch syndrome forms of CRC [167,168]. Moreover, BRAF mutations were identified in 40% of MSI-H and 4% of MSI-L tumors [169]. The majority of the BRAF mutations include the hotspot mutation, V600E (Val600Glu) and is found to correlate with poor prognosis in CRC patients [170,171].
Angiogenesis, the development of new blood vessels, is involved in tumor initiation, growth, and metastasis and involves several factors including vascular endothelial growth factors (VEGFs). In CRC, VEGF levels and VEGFR activity is enhanced and is associated with poor prognosis [172]. Elevated VEGF levels are seen in very early stages of colorectal neoplasia (adenoma); however, they were significantly elevated in a later stage of cancer (metastatic stage) [173]. Aberrant KRAS and TP53 as well as COX-2 expression regulate VEGF-VEGFR activity alteration, thus promoting cancer growth and migration [173,174]. The molecular pathways involved in the pathogenesis of CRC are depicted in Figure 2.

3. Therapeutic Strategies

Molecular profiling has helped in developing biomarkers that can potentially improve clinical outcomes in CRC patients and significantly enhance the survival of metastatic patients. Several molecular aberrations are defined for candidate biomarkers that have been tested in completed and ongoing trials in conjunction with targeted therapies and immunotherapies.

3.1. Targeting CIN Pathway

There are intense efforts to develop synthetic modulators of Wnt signaling including small molecules, peptides, and blocking antibodies to inhibit Wnt pathway [175]. As approved by US Food and Drug Administration (FDA), lithium chloride is already in clinical use and it is found to stimulate CTNNB1 by inhibiting GSK3. Moreover, non-steroidal anti-inflammatory drugs (NSAIDs) and celecoxib, the selective COX-2, block CTNNB1-dependent transcription in CRC [176,177] and lessen polyp formation in FAP patients as well in in vivo mice models of colon cancer [178,179]. Recently, two small molecules, XAV939 and pyrvinium, were identified using reporter-based screening approaches, while XAV939 increased AXIN stability by inhibiting tankyrase, pyrvinium stimulated CTNNB1 phosphorylation through casein kinase activation [180,181]. On the other hand, several Wnt-blocking antibodies blocked proliferation and induce apoptosis in different cancers [182]. In comparison to normal tissues, FZD7-specific antibodies selectively target colon cancer and HCC cells [183,184]. In vivo studies have shown that intraperitoneal injections of WNT3A-neutralizing antibodies reduce proliferation and increase apoptosis of prostate cancer in mouse models [185]. In addition, studies have indicated that use of Wnt-modulatory peptides, SFRP1, or SFRP1-derived peptides reduce HCT116 xenograft tumor development in nude mice [186,187].
On the other hand, no definite clinical role for TGF-β signaling pathway has been identified thus far; however, studies show that SMAD4 expression levels correlate with prognosis and response to 5-fluorouracil (5-FU) [188,189,190]. Moreover, current clinical trials (e.g., NCT00217737, also designated ECOG 5202) are studying the benefit of 18qLOH for directing the use of specific adjuvant therapies [72]. They also found that chronic use of NSAIDs, mainly Ibuprofen, stimulates a dose-dependent reduction of Notch pathway activity, thus confirming the protective effects of NSAIDs in CRC [72]. Remarkably, a few clinical trials aimed at targets including IGF-1R, Wnt, Notch, Hedgehog, and TGF-β are in progress; however, no definite results have appeared thus far. Phase II trials of the γ-secretase inhibitor (RO4929097) in Notch blockade therapy and the Hedgehog pathway inhibitor vismodegib can be considered promising [191,192].

3.2. Targeting MSI Pathway

MSI status has been demonstrated to be a reliable biomarker of immunotherapy response in the metastatic setting. Currently, checkpoint inhibitors are investigated in various solid tumors with promising responses [193,194]. Pembrolizumab, a humanized IgG4 antibody, was the first PD-1 blocker approved by FDA for treatment of metastatic CRC [193]. Another phase I trial in patients with MSI-H CRC found antitumor activity of pembrolizumab [194]. Although combination therapy of pembrolizumab and ipilimumab showed comparable efficacy in melanoma patients, the effect of this combination in CRC is still nascent [195]. The other humanized monoclonal IgG4-based PD-1 antibody, nivolumab, gained FDA approval for MSI-H metastatic CRC based on the results obtained in the CheckMate-142 clinical trial [196]. Remarkably, a combined therapy of nivolumab and ipilimumab improved the outcome of the patients with dMMR or MSI-H CRC who had previously received chemotherapy [197]. Other plausible candidate PD-1/PD-L1 inhibitors (atezolizumab, avelumab, and durvalumab) are under phase I trials for several tumors including CRC [198]. Furthermore, new immune checkpoint targets including TIM-3, TIGIT, T cell Ig, and T cell-derived LAG-3, which promote CRC progression, are also in phase I trials [199,200].

3.3. Targeting CIMP Pathway

Targeting the EGFR pathway generally involves use of anti-EGFR monoclonal antibodies and tyrosine kinase inhibitors (TKIs) aimed at intracellular kinases. Cetuximab was the first monoclonal antibody introduced to target EGFR. Cetuximab provokes EGFR internalization and degradation once bound to the external domain of EGFR [201]. Multiple studies have confirmed the beneficial effects of the cetuximab on CRC patients’ outcome [202]; in addition, a combined therapy of cetuximab with other existing chemotherapies also showed favorable results [203]. However, tumors carrying RAS, BRAF, or PIK3CA mutations, loss of PTEN, HER-2 amplification, and altered VEGF and VEGFR signaling are resistant to anti-EGFR therapy due to continuous activation of EGFR downstream signaling pathways [204]. In these cases, the patient is not eligible for anti-EGFR treatment with cetuximab. The other EGFR target, panitumumab, is a fully humanized antibody and, in comparison to cetuximab, does not provoke antibody-dependent cell-mediated cytotoxicity [205]. In the PRIME trial, efficacy of FOLFOX (folinic acid, fluorouracil, and oxaliplatin) alone as well as in combination with panitumumab was analyzed in patients with metastatic CRC. The combined therapy showed a higher OS and PFS than for FOLFOX alone [206,207]. PRIME and PEAK trials further analyzed the effects of panitumumab and fluorouracil/leucovorin (5-FU/LV) after panitumumab plus FOLFOX. The trials showed significant improvement in PFS and OS compared with panitumumab treatment alone [208]. In addition, the VALENTINO trial demonstrated synergistic effects of panitumumab with 5-FU/LV and better effects on outcome compared with treatment with panitumumab alone [209]. Although both cetuximab and panitumumab are FDA-approved drugs, panitumumab is economically more effective than cetuximab [210].
On the other hand, there are no approved specific targeted therapies for KRAS-mutated CRC. However, a novel agent, AMG 510, which is a small molecule that targets KRASG12 C mutation, was introduced. AMG 510 specifically and irreversibly blocks KRASG12 C by locking it in the inactive GDP-bound state [211]. In addition, other KRAS-modulating agents targeting G12C (MRTX849, LY3499446, or ARS-1620) and G12D mutations are also under investigations [212,213].
Although BRAF mutations are more frequent in melanoma and papillary thyroid carcinoma in comparison with CRC, a few studies analyzed the effects of BRAF inhibitors (vemurafenib or dabrafenib and trametinib) in patients with metastatic CRC. Although downstream MAPK activity was inhibited, the PFS or OS of patients did not improve [214,215,216]. However, the synergistic effect of BRAF and EGFR inhibitors in trials using vemurafenib in combination with IRI and cetuximab in BRAF-mutant CRC patients showed positive results [217,218,219,220]. Additionally, in a phase II trial, using encorafenib (a BRAF inhibitor) along with cetuximab, with or without alpelisib (ALP), improved the OS and PFS in advanced BRAF-mutant CRC patients [221]. Furthermore, a study by Corcoran et al. [218] found that patients with BRAF-V600E-mutated CRC, when treated with a triplet regimen (dabrafenib, trametinib, and panitumumab), produced a better response rate in comparison with the double regimens (dabrafenib + panitumumab or trametinib + panitumumab). The ongoing BEACON trial, which employs the triplet regimen of encorafenib, binimetinib (a MEK inhibitor), and cetuximab, has shown less toxicity and higher survival rates [222].
Anti-VEGF/VEGFR therapies are necessary to target steps in tumor metastasis. Bevacizumab (Avistin), a humanized IgG monoclonal antibody, was the first FDA-approved VEGF-targeted agent for metastatic CRC. According to several trials, bevacizumab showed to improve both progression-free survival and OS in metastatic CRC [223,224]. Furthermore, KRAS-mutant patients were also found to benefit from bevacizumab [225,226]. Moreover, combination of FOLFOX and bevacizumab improved progression-free survival and OS in CRC patients as compared to treatment with FOLFOX alone [227].

4. Conclusions

This brief review provides information about candidate biomarkers that can aid in improving the diagnosis and help with the early detection of CRC cases, thus ameliorating the prognosis of CRC patients. Although there are marked advances in CRC investigations, the role of molecular classification in therapeutic interventions remains unclear. The use of molecular alterations in predicting risk for CRC shows promise, while further studies are needed to determine if aberrantly methylated CpGs or other molecular alterations can be used as reliable and accurate indicators of risk for polyps or CRCs. Moreover, it is important to analyze the efficacy of multi-kinase/BRAF-inhibitor, sorafenib, in addition to other specific inhibitors of the EGFR as well as PI3K signaling pathway, in the treatment of CRC to further identify novel therapeutic targets. On the other hand, we believe that using new drug combinations and specific (personalized) targets will be the future avenue for efficient treatment of CRC patients with limited risk of toxicity and adverse side effects. Thus, understanding the underlying mechanisms of CRC cell genetic alteration and subsequent consequences can help pave the way for the development of novel diagnostic and therapeutic strategies.

Author Contributions

Conceptualization, A.M. and R.A.E.; writing—original draft preparation, I.G., A.M., A.A., and R.A.E.; writing—review and editing, A.-E.A.M., S.V., and A.M. All authors have read and agreed to the published version of the manuscript.

Funding

Our lab is supported by grants from Qatar University#: QUHI-CMED-19/20-1 and QUCG-CMED-20/21-2.

Conflicts of Interest

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

Abbreviations

5-FUFluorouracil
ACVR2Activin receptor type 2
AFAPAttenuated familial adenomatous polyposis
ALPAlpelisib
AMPKAdenosine monophosphate-activated protein kinase
APCAdenomatous polyposis coli
BRAFSerine/threonine-protein kinase B-Raf
BubBudding uninhibited by benzimidazoles
CDC20Cell division cycle 20
CDKNCyclin-dependent kinases
CIMPCpG island methylator phenotype
CINChromosomal instability
CMSConsensus molecular subtypes
COXCyclooxygenase
CRCColorectal cancer
CTNNB1Catenin beta-1
DCCDeleted in colorectal cancer
dMMRDefective mismatch repair
EDMExonuclease domain
EGFREpidermal growth factor receptor
ERKExtracellular signal related kinase
FAPFamilial adenomatous polyposis
FDAFood and Drug Administration
FOLFOXFolinic acid, fluorouracil, and oxaliplatin
GSHGlutathione
GSK3-βGlycogen synthase kinase 3-β
HNPCCHereditary non-polyposis colorectal cancer
IBDInflammatory bowel disease; Ig: Immunoglobin
IGFBP7Insulin-like growth factor-binding protein 7
JAKJanus kinase
KRASKirsten ras
LOHLoss of heterozygosity
LVLeucovorin
MAPMUTYH-associated polyposis
MAP/MEKMitogen-activated protein kinase
MLH/MSHMutL homolog
MSIMicrosatellite instability
MSSMicrosatellite stable
NF-κBNuclear factor kappa
NSAIDsNon-steroidal anti-inflammatory drugs
OSOverall survival
PD-1/PD-L1Programmed death ligand-1
PI3KPhosphatidylinositol 3-kinase
PI3KCAPhosphatidylinositol-4,5-Bisphosphate 3-Kinase Catalytic Subunit Alpha
POLE/POLD1Polymerase epsilon or delta 1
RAC1Rac family small GTPase 1
ROSReactive oxygen species
SCNAsSomatic copy number alterations
SMADSmall mothers against decapentaplegic
STATSignal transducer and activator of transcription
TCGAThe Cancer Genome Atlas
TGF-βTransforming growth factor-β
TIGART53-induced glycolysis regulatory phosphatase
TNMTumor-nodes-metastasis
TP53Tumor protein 53
TSGTumor suppressor gene
VEGFVascular endothelial growth factor
VEGFRVascular endothelial growth factor receptor
WntWingless

References

  1. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Mathers, C.; Parkin, D.M.; Piñeros, M.; Znaor, A.; Bray, F. Estimating the global cancer incidence and mortality in 2018: GLOBOCAN sources and methods. Int. J. Cancer 2019, 144, 1941–1953. [Google Scholar] [CrossRef] [Green Version]
  2. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [Green Version]
  3. Cancer Genome Atlas Network. Comprehensive molecular characterization of human colon and rectal cancer. Nature 2012, 487, 330–337. [Google Scholar] [CrossRef] [Green Version]
  4. Guinney, J.; Dienstmann, R.; Wang, X.; de Reyniès, A.; Schlicker, A.; Soneson, C.; Marisa, L.; Roepman, P.; Nyamundanda, G.; Angelino, P.; et al. The consensus molecular subtypes of colorectal cancer. Nat. Med. 2015, 21, 1350–1356. [Google Scholar] [CrossRef]
  5. O’Connell, J.B.; Maggard, M.A.; Ko, C.Y. Colon cancer survival rates with the new American Joint Committee on Cancer sixth edition staging. J. Natl. Cancer Inst. 2004, 96, 1420–1425. [Google Scholar] [CrossRef] [PubMed]
  6. Bogaert, J.; Prenen, H. Molecular genetics of colorectal cancer. Ann. Gastroenterol. 2014, 27, 9–14. [Google Scholar] [PubMed]
  7. Le Marchand, L.; Wilkens, L.R.; Hankin, J.H.; Kolonel, L.N.; Lyu, L.-C. A case-control study of diet and colorectal cancer in a multiethnic population in Hawaii (United States): Lipids and foods of animal origin. Cancer Causes Control 1997, 8, 637–648. [Google Scholar] [CrossRef] [PubMed]
  8. Wirtzfeld, D.A.; Petrelli, N.J.; Rodriguez-Bigas, M.A. Hamartomatous polyposis syndromes: Molecular genetics, neoplastic risk, and surveillance recommendations. Ann. Surg. Oncol. 2001, 8, 319–327. [Google Scholar] [CrossRef]
  9. Xie, J.; Itzkowitz, S.H. Cancer in inflammatory bowel disease. World J. Gastroenterol. 2008, 14, 378–389. [Google Scholar] [CrossRef]
  10. Fearon, E.R.; Vogelstein, B. A genetic model for colorectal tumorigenesis. Cell 1990, 61, 759–767. [Google Scholar] [CrossRef]
  11. Grady, W.M. Epigenetic events in the colorectum and in colon cancer. Biochem. Soc. Trans. 2005, 33, 684–688. [Google Scholar] [CrossRef] [PubMed]
  12. Grady, W.M.; Markowitz, S.D. Genetic and Epigenetic Alterations in Colon Cancer. Ann. Rev. Genom. Hum. Genet. 2002, 3, 101. [Google Scholar] [CrossRef] [PubMed]
  13. Lin, J.K.; Chang, S.C.; Yang, Y.C.; Li, A.F. Loss of heterozygosity and DNA aneuploidy in colorectal adenocarcinoma. Ann. Surg. Oncol. 2003, 10, 1086–1094. [Google Scholar] [CrossRef] [PubMed]
  14. Leary, R.J.; Lin, J.C.; Cummins, J.; Boca, S.; Wood, L.D.; Parsons, D.W.; Jones, S.; Sjöblom, T.; Park, B.H.; Parsons, R.; et al. Integrated analysis of homozygous deletions, focal amplifications, and sequence alterations in breast and colorectal cancers. Proc. Natl. Acad. Sci. USA 2008, 105, 16224–16229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Pino, M.S.; Chung, D.C. The chromosomal instability pathway in colon cancer. Gastroenterology 2010, 138, 2059–2072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Markowitz, S.D.; Bertagnolli, M.M. Molecular Basis of Colorectal Cancer. N. Engl. J. Med. 2009, 361, 2449–2460. [Google Scholar] [CrossRef] [Green Version]
  17. Tsang, A.H.; Cheng, K.H.; Wong, A.S.; Ng, S.S.; Ma, B.B.; Chan, C.M.; Tsui, N.B.; Chan, L.W.; Yung, B.Y.; Wong, S.C. Current and future molecular diagnostics in colorectal cancer and colorectal adenoma. World J. Gastroenterol. 2014, 20, 3847–3857. [Google Scholar] [CrossRef]
  18. Baker, S.J.; Fearon, E.R.; Nigro, J.M.; Hamilton, S.R.; Preisinger, A.C.; Jessup, J.M.; van Tuinen, P.; Ledbetter, D.H.; Barker, D.F.; Nakamura, Y.; et al. Chromosome 17 deletions and p53 gene mutations in colorectal carcinomas. Science 1989, 244, 217–221. [Google Scholar] [CrossRef]
  19. Thiagalingam, S.; Lengauer, C.; Leach, F.S.; Schutte, M.; Hahn, S.A.; Overhauser, J.; Willson, J.K.; Markowitz, S.; Hamilton, S.R.; Kern, S.E.; et al. Evaluation of candidate tumour suppressor genes on chromosome 18 in colorectal cancers. Nat. Genet. 1996, 13, 343–346. [Google Scholar] [CrossRef]
  20. Markowitz, S.; Wang, J.; Myeroff, L.; Parsons, R.; Sun, L.; Lutterbaugh, J.; Fan, R.S.; Zborowska, E.; Kinzler, K.W.; Vogelstein, B.; et al. Inactivation of the type II TGF-beta receptor in colon cancer cells with microsatellite instability. Science 1995, 268, 1336–1338. [Google Scholar] [CrossRef]
  21. Samuels, Y.; Wang, Z.; Bardelli, A.; Silliman, N.; Ptak, J.; Szabo, S.; Yan, H.; Gazdar, A.; Powell, S.M.; Riggins, G.J.; et al. High frequency of mutations of the PIK3CA gene in human cancers. Science 2004, 304, 554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Diep, C.B.; Kleivi, K.; Ribeiro, F.R.; Teixeira, M.R.; Lindgjaerde, O.C.; Lothe, R.A. The order of genetic events associated with colorectal cancer progression inferred from meta-analysis of copy number changes. Genes Chromosom. Cancer 2006, 45, 31–41. [Google Scholar] [CrossRef] [PubMed]
  23. Jasmine, F.; Rahaman, R.; Dodsworth, C.; Roy, S.; Paul, R.; Raza, M.; Paul-Brutus, R.; Kamal, M.; Ahsan, H.; Kibriya, M.G. A genome-wide study of cytogenetic changes in colorectal cancer using SNP microarrays: Opportunities for future personalized treatment. PLoS ONE 2012, 7, e31968. [Google Scholar] [CrossRef] [PubMed]
  24. Baudis, M. Genomic imbalances in 5918 malignant epithelial tumors: An explorative meta-analysis of chromosomal CGH data. BMC Cancer 2007, 7, 226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Nowell, P.C. The clonal evolution of tumor cell populations. Science 1976, 194, 23–28. [Google Scholar] [CrossRef]
  26. Béroud, C.; Soussi, T. APC gene: Database of germline and somatic mutations in human tumors and cell lines. Nucleic Acids Res. 1996, 24, 121–124. [Google Scholar] [CrossRef] [Green Version]
  27. Aoki, K.; Taketo, M.M. Adenomatous polyposis coli (APC): A multi-functional tumor suppressor gene. J. Cell Sci. 2007, 120, 3327–3335. [Google Scholar] [CrossRef] [Green Version]
  28. Kaplan, K.B.; Burds, A.A.; Swedlow, J.R.; Bekir, S.S.; Sorger, P.K.; Näthke, I.S. A role for the Adenomatous Polyposis Coli protein in chromosome segregation. Nat. Cell Biol. 2001, 3, 429–432. [Google Scholar] [CrossRef]
  29. Näthke, I.S.; Adams, C.L.; Polakis, P.; Sellin, J.H.; Nelson, W.J. The adenomatous polyposis coli tumor suppressor protein localizes to plasma membrane sites involved in active cell migration. J. Cell Biol. 1996, 134, 165–179. [Google Scholar] [CrossRef]
  30. Browne, S.J.; MacFarlane, M.; Cohen, G.M.; Paraskeva, C. The adenomatous polyposis coli protein and retinoblastoma protein are cleaved early in apoptosis and are potential substrates for caspases. Cell Death Differ. 1998, 5, 206–213. [Google Scholar] [CrossRef]
  31. Sansom, O.J.; Reed, K.R.; Hayes, A.J.; Ireland, H.; Brinkmann, H.; Newton, I.P.; Batlle, E.; Simon-Assmann, P.; Clevers, H.; Nathke, I.S.; et al. Loss of Apc in vivo immediately perturbs Wnt signaling, differentiation, and migration. Genes Dev. 2004, 18, 1385–1390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Baeg, G.H.; Matsumine, A.; Kuroda, T.; Bhattacharjee, R.N.; Miyashiro, I.; Toyoshima, K.; Akiyama, T. The tumour suppressor gene product APC blocks cell cycle progression from G0/G1 to S phase. EMBO J. 1995, 14, 5618–5625. [Google Scholar] [CrossRef] [PubMed]
  33. Grossi, V.; Fasano, C.; Celestini, V.; Lepore Signorile, M.; Sanese, P.; Simone, C. Chasing the FOXO3: Insights into Its New Mitochondrial Lair in Colorectal Cancer Landscape. Cancers 2019, 11, 414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Miyaki, E.; Hiraga, N.; Imamura, M.; Uchida, T.; Kan, H.; Tsuge, M.; Abe-Chayama, H.; Hayes, C.N.; Makokha, G.N.; Serikawa, M.; et al. Interferon alpha treatment stimulates interferon gamma expression in type I NKT cells and enhances their antiviral effect against hepatitis C virus. PLoS ONE 2017, 12, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Franovic, A.; Holterman, C.E.; Payette, J.; Lee, S. Human cancers converge at the HIF-2α oncogenic axis. Proc. Natl. Acad. Sci. USA 2009, 106, 21306–21311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Chittenden, T.W.; Howe, E.A.; Culhane, A.C.; Sultana, R.; Taylor, J.M.; Holmes, C.; Quackenbush, J. Functional classification analysis of somatically mutated genes in human breast and colorectal cancers. Genomics 2008, 91, 508–511. [Google Scholar] [CrossRef] [Green Version]
  37. Kinzler, K.W.; Vogelstein, B. Lessons from hereditary colorectal cancer. Cell 1996, 87, 159–170. [Google Scholar] [CrossRef] [Green Version]
  38. Bonk, T.; Humeny, A.; Sutter, C.; Gebert, J.; von Knebel Doeberitz, M.; Becker, C.-M. Molecular diagnosis of familial adenomatous polyposis (FAP): Genotyping of adenomatous polyposis coli (APC) alleles by MALDI-TOF mass spectrometry. Clin. Biochem. 2002, 35, 87. [Google Scholar] [CrossRef]
  39. Powell, S.M.; Petersen, G.M.; Krush, A.J.; Booker, S.; Jen, J.; Giardiello, F.M.; Hamilton, S.R.; Vogelstein, B.; Kinzler, K.W. Molecular Diagnosis of Familial Adenomatous Polyposis. N. Engl. J. Med. 1993, 329, 1982–1987. [Google Scholar] [CrossRef]
  40. Ishiguro, K.; Yoshida, T.; Yagishita, H.; Numata, Y.; Okayasu, T. Epithelial and stromal genetic instability contributes to genesis of colorectal adenomas. Gut 2006, 55, 695–702. [Google Scholar] [CrossRef] [Green Version]
  41. Donger, Z.; Liu, Y.; Liangtao, Z.; Weiting, G.; Dan, L.; Yong, Z.; Xueda, H.; Zhibo, G.; Jinghong, X.; Yanqin, H.; et al. Exome Capture Sequencing of Adenoma Reveals Genetic Alterations in Multiple Cellular Pathways at the Early Stage of Colorectal Tumorigenesis. PLoS ONE 2013, 8, 1–8. [Google Scholar] [CrossRef]
  42. Cheng, T.H.T.; Gorman, M.; Martin, L.; Barclay, E.; Casey, G.; Saunders, B.; Thomas, H.; Clark, S.; Tomlinson, I. Common colorectal cancer risk alleles contribute to the multiple colorectal adenoma phenotype, but do not influence colonic polyposis in FAP. Eur. J. Hum. Genet. 2015, 23, 260–263. [Google Scholar] [CrossRef] [PubMed]
  43. Vaqué, J.P.; Martínez, N.; Varela, I.; Fernández, F.; Mayorga, M.; Derdak, S.; Beltrán, S.; Moreno, T.; Almaraz, C.; De las Heras, G.; et al. Colorectal Adenomas Contain Multiple Somatic Mutations That Do Not Coincide with Synchronous Adenocarcinoma Specimens. PLoS ONE 2015, 10, 1–12. [Google Scholar] [CrossRef]
  44. Armaghany, T.; Wilson, J.D.; Chu, Q.; Mills, G. Genetic alterations in colorectal cancer. Gastrointest. Cancer Res. 2012, 5, 19–27. [Google Scholar] [PubMed]
  45. Sieber, O.M.; Lipton, L.; Crabtree, M.; Heinimann, K.; Fidalgo, P.; Phillips, R.K.S.; Bisgaard, M.-L.; Orntoft, T.F.; Aaltonen, L.A.; Hodgson, S.V.; et al. Multiple Colorectal Adenomas, Classic Adenomatous Polyposis, and Germ-Line Mutations in MYH. N. Engl. J. Med. 2003, 348, 791–799. [Google Scholar] [CrossRef] [PubMed]
  46. Fearnhead, N.S.; Britton, M.P.; Bodmer, W.F. The ABC of APC. Hum. Mol. Genet. 2001, 10, 721–733. [Google Scholar] [CrossRef] [Green Version]
  47. Rowan, A.J.; Lamlum, H.; Ilyas, M.; Wheeler, J.; Straub, J.; Papadopoulou, A.; Bicknell, D.; Bodmer, W.F.; Tomlinson, I.P. APC mutations in sporadic colorectal tumors: A mutational “hotspot” and interdependence of the “two hits”. Proc. Natl. Acad. Sci. USA 2000, 97, 3352–3357. [Google Scholar] [CrossRef] [Green Version]
  48. Cottrell, S.; Bicknell, D.; Kaklamanis, L.; Bodmer, W.F. Molecular analysis of APC mutations in familial adenomatous polyposis and sporadic colon carcinomas. Lancet 1992, 340, 626–630. [Google Scholar] [CrossRef]
  49. Christie, M.; Jorissen, R.N.; Mouradov, D.; Sakthianandeswaren, A.; Li, S.; Day, F.; Tsui, C.; Lipton, L.; Desai, J.; Jones, I.T.; et al. Different APC genotypes in proximal and distal sporadic colorectal cancers suggest distinct WNT/β-catenin signalling thresholds for tumourigenesis. Oncogene 2013, 32, 4675–4682. [Google Scholar] [CrossRef]
  50. Miller, J.R.; Moon, R.T. Signal transduction through beta-catenin and specification of cell fate during embryogenesis. Genes Dev. 1996, 10, 2527–2539. [Google Scholar] [CrossRef] [Green Version]
  51. Klaus, A.; Birchmeier, W. Wnt signalling and its impact on development and cancer. Nat. Rev. Cancer 2008, 8, 387–398. [Google Scholar] [CrossRef] [PubMed]
  52. Rubinfeld, B.; Albert, I.; Porfiri, E.; Fiol, C.; Munemitsu, S.; Polakis, P. Binding of GSK3beta to the APC-beta-catenin complex and regulation of complex assembly. Science 1996, 272, 1023–1026. [Google Scholar] [CrossRef] [PubMed]
  53. Pronobis, M.I.; Rusan, N.M.; Peifer, M. A novel GSK3-regulated APC:Axin interaction regulates Wnt signaling by driving a catalytic cycle of efficient βcatenin destruction. Elife 2015, 4, e08022. [Google Scholar] [CrossRef] [PubMed]
  54. Morin, P.J.; Sparks, A.B. Activation of beta-catenin-Tcf signaling in colon cancer by mutations in beta-catenin or APC. Science 1997, 275, 1787–1790. [Google Scholar] [CrossRef] [Green Version]
  55. Sekiya, T.; Adachi, S.; Kohu, K.; Yamada, T.; Higuchi, O.; Furukawa, Y.; Nakamura, Y.; Nakamura, T.; Tashiro, K.; Kuhara, S.; et al. Identification of BMP and Activin Membrane-bound Inhibitor (BAMBI), an Inhibitor of Transforming Growth Factor-β Signaling, as a Target of the β-Catenin Pathway in Colorectal Tumor Cells. J. Biol. Chem. 2004, 279, 6840–6846. [Google Scholar] [CrossRef] [Green Version]
  56. He, L.; Lu, N.; Dai, Q.; Zhao, Y.; Zhao, L.; Wang, H.; Li, Z.; You, Q.; Guo, Q. Wogonin induced G1 cell cycle arrest by regulating Wnt/β-catenin signaling pathway and inactivating CDK8 in human colorectal cancer carcinoma cells. Toxicology 2013, 312, 36–47. [Google Scholar] [CrossRef]
  57. Bolanos-Garcia, V.M.; Blundell, T.L. BUB1 and BUBR1: Multifaceted kinases of the cell cycle. Trends Biochem. Sci. 2011, 36, 141–150. [Google Scholar] [CrossRef] [Green Version]
  58. Tasaki, T.; Sohr, R.; Zanxian, X.; Heliweg, R.; Hörtnagi, H.; Varshavsky, A.; Yong Tae, K. Biochemical and Genetic Studies of UBR3, a Ubiquitin Ligase with a Function in Olfactory and Other Sensory Systems. J. Biol. Chem. 2007, 282, 18510–18520. [Google Scholar] [CrossRef] [Green Version]
  59. Nilsson, J.; Yekezare, M.; Minshull, J.; Pines, J. The APC/C maintains the spindle assembly checkpoint by targeting Cdc20 for destruction. Nat. Cell Biol. 2008, 10, 1411–1420. [Google Scholar] [CrossRef]
  60. Shin, H.-J.; Baek, K.-H.; Jeon, A.-H.; Park, M.-T.; Lee, S.-J.; Kang, C.-M.; Lee, H.-S.; Yoo, S.-H.; Chung, D.-H.; Sung, Y.-C.; et al. Dual roles of human BubR1, a mitotic checkpoint kinase, in the monitoring of chromosomal instability. Cancer Cell 2003, 4, 483. [Google Scholar] [CrossRef] [Green Version]
  61. Takayama, O.; Yamamoto, H.; Damdinsuren, B.; Sugita, Y.; Ngan, C.Y.; Xu, X.; Tsujino, T.; Takemasa, I.; Ikeda, M.; Sekimoto, M.; et al. Expression of PPARδ in multistage carcinogenesis of the colorectum: Implications of malignant cancer morphology. Br. J. Cancer 2006, 95, 889–895. [Google Scholar] [CrossRef] [PubMed]
  62. Renehan, A.G.; O’Dwyer, S.T.; Haboubi, N.J.; Potten, C.S. Early cellular events in colorectal carcinogenesis. Colorectal Dis. 2002, 4, 76–89. [Google Scholar] [CrossRef] [PubMed]
  63. Luebeck, E.G.; Moolgavkar, S.H. Multistage carcinogenesis and the incidence of colorectal cancer. Proc. Natl. Acad. Sci. USA 2002, 99, 15095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. An, N.; Zhao, C.; Yu, Z.; Yang, X. Identification of prognostic genes in colorectal cancer through transcription profiling of multi-stage carcinogenesis. Oncol. Lett. 2019, 17, 432–441. [Google Scholar] [CrossRef] [Green Version]
  65. Tetsu, O.; McCormick, F. Beta-catenin regulates expression of cyclin D1 in colon carcinoma cells. Nature 1999, 398, 422–426. [Google Scholar] [CrossRef]
  66. He, T.C.; Sparks, A.B.; Rago, C.; Hermeking, H.; Zawel, L.; da Costa, L.T.; Morin, P.J.; Vogelstein, B.; Kinzler, K.W. Identification of c-MYC as a target of the APC pathway. Science 1998, 281, 1509–1512. [Google Scholar] [CrossRef]
  67. Soohyun, P.; Jie, C.; Wangsheng, Y.; Ling, W.; Carmon, K.S.; Qingyun, J.L. Differential activities and mechanisms of the four R-spondins in potentiating Wnt/β-catenin signaling. J. Biol. Chem. 2018, 293, 9759–9769. [Google Scholar] [CrossRef] [Green Version]
  68. Heinen, C.D.; Goss, K.H.; Cornelius, J.R.; Babcock, G.F.; Knudsen, E.S.; Kowalik, T.; Groden, J. The APC tumor suppressor controls entry into S-phase through its ability to regulate the cyclin D/RB pathway. Gastroenterology 2002, 123, 751–763. [Google Scholar] [CrossRef]
  69. Arber, N.; Hibshoosh, H.; Moss, S.F.; Sutter, T.; Zhang, Y.; Begg, M.; Wang, S.; Weinstein, I.B.; Holt, P.R. Increased expression of cyclin D1 is an early event in multistage colorectal carcinogenesis. Gastroenterology 1996, 110, 669–674. [Google Scholar] [CrossRef]
  70. Van Es, J.H.; van Gijn, M.E.; Riccio, O.; van den Born, M.; Vooijs, M.; Begthel, H.; Cozijnsen, M.; Robine, S.; Winton, D.J.; Radtke, F.; et al. Notch/γ-secretase inhibition turns proliferative cells in intestinal crypts and adenomas into goblet cells. Nature 2005, 435, 959–963. [Google Scholar] [CrossRef]
  71. Hai, Z.; Pritchard, D.M.; Xiangdong, Y.; Bennett, E.; Gang, L.; Chunming, L.; Ai, W. KLF4 gene expression is inhibited by the notch signaling pathway that controls goblet cell differentiation in mouse gastrointestinal tract. Am. J. Physiol. Gastrointest. Liver Physiol. 2009, 296, G490–G498. [Google Scholar] [CrossRef] [Green Version]
  72. Kwon, C.; Cheng, P.; King, I.N.; Andersen, P.; Shenje, L.; Nigam, V.; Srivastava, D. Notch post-translationally regulates ?-catenin protein in stem and progenitor cells. Nat. Cell Biol. 2011, 13, 1244–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Firestein, R.; Bass, A.J.; Young, K.S.; Dunn, I.F.; Silver, S.J.; Guney, I.; Freed, E.; Ligon, A.H.; Vena, N.; Ogino, S.; et al. CDK8 is a colorectal cancer oncogene that regulates β-catenin activity. Nature 2008, 455, 547–551. [Google Scholar] [CrossRef] [PubMed]
  74. Fryer, C.J.; White, J.B.; Jones, K.A. Mastermind Recruits CycC:CDK8 to Phosphorylate the Notch ICD and Coordinate Activation with Turnover. Mol. Cell 2004, 16, 509–520. [Google Scholar] [CrossRef] [PubMed]
  75. Levine, A.J. p53, the cellular gatekeeper for growth and division. Cell 1997, 88, 323–331. [Google Scholar] [CrossRef] [Green Version]
  76. Smith, G.; Carey, F.A.; Beattie, J.; Wilkie, M.J.V.; Lightfoot, T.J.; Coxhead, J.; Garner, R.C.; Steele, R.J.C.; Wolf, C.R. Mutations in APC, Kirsten-ras, and p53—Alternative genetic pathways to colorectal cancer. Proc. Natl. Acad. Sci. USA 2002, 99, 9433. [Google Scholar] [CrossRef] [Green Version]
  77. Vogelstein, B.; Fearon, E.R.; Hamilton, S.R.; Kern, S.E.; Preisinger, A.C.; Leppert, M.; Nakamura, Y.; White, R.; Smits, A.M.; Bos, J.L. Genetic alterations during colorectal-tumor development. N. Engl. J. Med. 1988, 319, 525–532. [Google Scholar] [CrossRef] [Green Version]
  78. Sigal, A.; Rotter, V. Oncogenic mutations of the p53 tumor suppressor: The demons of the guardian of the genome. Cancer Res. 2000, 60, 6788–6793. [Google Scholar]
  79. Liu, Y.; Bodmer, W.F. Analysis of P53 mutations and their expression in 56 colorectal cancer cell lines. Proc. Natl. Acad. Sci. USA 2006, 103, 976–981. [Google Scholar] [CrossRef] [Green Version]
  80. Oikawa, T.; Okuda, M.; Ma, Z.; Goorha, R.; Tsujimoto, H.; Inokuma, H.; Fukasawa, K. Transcriptional control of BubR1 by p53 and suppression of centrosome amplification by BubR1. Mol. Cell. Biol. 2005, 25, 4046–4061. [Google Scholar] [CrossRef] [Green Version]
  81. Colussi, D.; Brandi, G.; Bazzoli, F.; Ricciardiello, L. Molecular Pathways Involved in Colorectal Cancer: Implications for Disease Behavior and Prevention. Int. J. Mol. Sci. 2013, 14, 16365–16385. [Google Scholar] [CrossRef] [Green Version]
  82. Zhang, J.; Su, G.; Lin, Y.; Meng, W.; Lai, J.K.L.; Qiao, L.; Li, X.; Xie, X. Targeting cyclin-dependent kinases in gastrointestinal cancer therapy. Discov. Med. 2019, 27, 27–36. [Google Scholar] [PubMed]
  83. Jones, R.G.; Plas, D.R.; Kubek, S.; Buzzai, M.; Mu, J.; Xu, Y.; Birnbaum, M.J.; Thompson, C.B. AMP-activated protein kinase induces a p53-dependent metabolic checkpoint. Mol. Cell 2005, 18, 283–293. [Google Scholar] [CrossRef] [PubMed]
  84. Pasz-Walczak, G.; Kordek, R.; Faflik, M. P21 (WAF1) Expression in Colorectal Cancer: Correlation with P53 and Cyclin D1 Expression, Clinicopathological Parameters and Prognosis. Pathol. Res. Pract. 2001, 197, 683–689. [Google Scholar] [CrossRef] [PubMed]
  85. Abukhdeir, A.M.; Park, B.H. P21 and p27: Roles in carcinogenesis and drug resistance. Expert Rev. Mol. Med. 2008, 10, e19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Nosho, K.; Kawasaki, T.; Chan, A.T.; Ohnishi, M.; Suemoto, Y.; Kirkner, G.J.; Fuchs, C.S.; Ogino, S. Cyclin D1 is frequently overexpressed in microsatellite unstable colorectal cancer, independent of CpG island methylator phenotype. Histopathology 2008, 53, 588–598. [Google Scholar] [CrossRef] [Green Version]
  87. Ogino, S.; Brahmandam, M.; Kawasaki, T.; Kirkner, G.J.; Loda, M.; Fuchs, C.S. Combined Analysis of COX-2 and p53 Expressions Reveals Synergistic Inverse Correlations with Microsatellite Instability and CpG Island Methylator Phenotype in Colorectal Cancer. Neoplasia 2006, 8, 458–464. [Google Scholar] [CrossRef] [Green Version]
  88. Venkatesan, P.; Bhutia, S.K.; Singh, A.K.; Das, S.K.; Dash, R.; Chaudhury, K.; Sarkar, D.; Fisher, P.B.; Mandal, M. AEE788 potentiates celecoxib-induced growth inhibition and apoptosis in human colon cancer cells. Life Sci. 2012, 91, 789–799. [Google Scholar] [CrossRef]
  89. Park, G.-B.; Jin, D.-H.; Kim, D. Sequential treatment with celecoxib and bortezomib enhances the ER stress-mediated autophagy-associated cell death of colon cancer cells. Oncol. Lett. 2018, 16, 4526–4536. [Google Scholar] [CrossRef] [Green Version]
  90. Schwarzenbach, H. Loss of Heterozygosity. In Brenner’s Encyclopedia of Genetics, 2nd ed.; Maloy, S., Hughes, K., Eds.; Academic Press: San Diego, CA, USA, 2013; pp. 271–273. [Google Scholar] [CrossRef]
  91. Sheffer, M.; Bacolod, M.D.; Zuk, O.; Giardina, S.F.; Pincas, H.; Barany, F.; Paty, P.B.; Gerald, W.L.; Notterman, D.A.; Domany, E. Association of survival and disease progression with chromosomal instability: A genomic exploration of colorectal cancer. Proc. Natl. Acad. Sci. USA 2009, 106, 7131–7136. [Google Scholar] [CrossRef] [Green Version]
  92. Jen, J.; Kim, H.; Piantadosi, S.; Liu, Z.F.; Levitt, R.C.; Sistonen, P.; Kinzler, K.W.; Vogelstein, B.; Hamilton, S.R. Allelic loss of chromosome 18q and prognosis in colorectal cancer. N. Engl. J. Med. 1994, 331, 213–221. [Google Scholar] [CrossRef]
  93. Fearon, E.R.; Cho, K.R.; Nigro, J.M.; Kern, S.E.; Simons, J.W.; Ruppert, J.M.; Hamilton, S.R.; Preisinger, A.C.; Thomas, G.; Kinzler, K.W.; et al. Identification of a chromosome 18q gene that is altered in colorectal cancers. Science 1990, 247, 49–56. [Google Scholar] [CrossRef] [PubMed]
  94. Mehlen, P.; Fearon, E.R. Role of the dependence receptor DCC in colorectal cancer pathogenesis. J. Clin. Oncol. 2004, 22, 3420–3428. [Google Scholar] [CrossRef] [PubMed]
  95. Mazelin, L.; Bernet, A.; Bonod-Bidaud, C.; Pays, L.; Arnaud, S.; Gespach, C.; Bredesen, D.E.; Scoazec, J.Y.; Mehlen, P. Netrin-1 controls colorectal tumorigenesis by regulating apoptosis. Nature 2004, 431, 80–84. [Google Scholar] [CrossRef]
  96. Takaku, K.; Oshima, M.; Miyoshi, H.; Matsui, M.; Seldin, M.F.; Taketo, M.M. Intestinal tumorigenesis in compound mutant mice of both Dpc4 (SMAD4) and Apc genes. Cell 1998, 92, 645. [Google Scholar] [CrossRef] [Green Version]
  97. Ritterhouse, L.L.; Wu, E.Y.; Kim, W.G.; Dillon, D.A.; Hirsch, M.S.; Sholl, L.M.; Agoston, A.T.; Setia, N.; Lauwers, G.Y.; Park, D.Y.; et al. Loss of SMAD4 protein expression in gastrointestinal and extra-gastrointestinal carcinomas. Histopathology 2019, 75, 546–551. [Google Scholar] [CrossRef] [PubMed]
  98. Fleming, N.I.; Jorissen, R.N.; Mouradov, D.; Christie, M.; Sakthianandeswaren, A.; Palmieri, M.; Day, F.; Li, S.; Tsui, C.; Lipton, L.; et al. SMAD2, SMAD3 and SMAD4 mutations in colorectal cancer. Cancer Res. 2013, 73, 725–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Shi, Y.; Hata, A.; Lo, R.S.; Massagué, J.; Pavletich, N.P. A structural basis for mutational inactivation of the tumour suppressor Smad4. Nature 1997, 388, 87–93. [Google Scholar] [CrossRef]
  100. Eppert, K.; Scherer, S.W. MADR2 maps to 18q21 and encodes a TGFbeta-regulated MAD-related protein that is functionally. Cell 1996, 86, 544. [Google Scholar] [CrossRef] [Green Version]
  101. Grady, D. How to Halve the Death Rate from Colon Cancer. New York Times 2007, 156, H5. [Google Scholar]
  102. Thibodeau, S.N.; Bren, G.; Schaid, D. Microsatellite instability in cancer of the proximal colon. Science 1993, 260, 816–819. [Google Scholar] [CrossRef] [PubMed]
  103. Geiersbach, K.B.; Samowitz, W.S. Microsatellite instability and colorectal cancer. Arch. Pathol. Lab. Med. 2011, 135, 1269–1277. [Google Scholar] [CrossRef] [PubMed]
  104. Ogino, S.; Nosho, K.; Irahara, N.; Shima, K.; Baba, Y.; Kirkner, G.J.; Meyerhardt, J.A.; Fuchs, C.S. Prognostic significance and molecular associations of 18q loss of heterozygosity: A cohort study of microsatellite stable colorectal cancers. J. Clin. Oncol. 2009, 27, 4591–4598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Perucho, M. Cancer of the microsatellite mutator phenotype. Biol. Chem. 1996, 377, 675–684. [Google Scholar]
  106. Mori, Y.; Yin, J.; Rashid, A.; Leggett, B.A.; Young, J.; Simms, L.; Kuehl, P.M.; Langenberg, P.; Meltzer, S.J.; Stine, O.C. Instabilotyping: Comprehensive identification of frameshift mutations caused by coding region microsatellite instability. Cancer Res. 2001, 61, 6046–6049. [Google Scholar]
  107. Parsons, R.; Myeroff, L.L.; Liu, B.; Willson, J.K.; Markowitz, S.D.; Kinzler, K.W.; Vogelstein, B. Microsatellite instability and mutations of the transforming growth factor beta type II receptor gene in colorectal cancer. Cancer Res. 1995, 55, 5548–5550. [Google Scholar]
  108. Boland, C.R.; Goel, A. Microsatellite instability in colorectal cancer. Gastroenterology 2010, 138, 2073–2087. [Google Scholar] [CrossRef]
  109. Velho, S.; Moutinho, C.; Cirnes, L.; Albuquerque, C.; Hamelin, R.; Schmitt, F.; Carneiro, F.; Oliveira, C.; Seruca, R. BRAF, KRAS and PIK3CA mutations in colorectal serrated polyps and cancer: Primary or secondary genetic events in colorectal carcinogenesis? BMC Cancer 2008, 8, 1–6. [Google Scholar] [CrossRef] [Green Version]
  110. Kim, J.H.; Rhee, Y.Y.; Kim, K.J.; Cho, N.Y.; Lee, H.S.; Kang, G.H. Annexin A10 expression correlates with serrated pathway features in colorectal carcinoma with microsatellite instability. APMIS 2014, 122, 1187–1195. [Google Scholar] [CrossRef]
  111. Meyer, L.A.; Broaddus, R.R.; Lu, K.H. Endometrial cancer and Lynch syndrome: Clinical and pathologic considerations. Cancer Control 2009, 16, 14–22. [Google Scholar] [CrossRef] [Green Version]
  112. Kaiser, J.C.; Meckbach, R.; Jacob, P. Genomic Instability and Radiation Risk in Molecular Pathways to Colon Cancer. PLoS ONE 2014, 9, 1–12. [Google Scholar] [CrossRef] [PubMed]
  113. Huaizeng, C.; Dafeng, Y.; Xing, X.; Weiguo, L.; Changkun, Z.; Xiaodong, C. Mismatch repair gene promoter methylation and expression in hydatidiform moles. Arch. Gynecol. Obstet. 2005, 272, 35–39. [Google Scholar] [CrossRef]
  114. Manes, M.; Garcia-Gomes, M.d.S.A.; Sandini, T.M.; Zaccarelli-Magalhães, J.; Florio, J.C.; Alexandre-Ribeiro, S.R.; Wadt, D.; Bernardi, M.M.; Massironi, S.M.G.; Mori, C.M.C. Behavioral and neurochemical characterization of the mlh mutant mice lacking otoconia. Behav. Brain Res. 2019, 359, 958–966. [Google Scholar] [CrossRef] [PubMed]
  115. Takehara, Y.; Nagasaka, T.; Nyuya, A.; Haruma, T.; Haraga, J.; Mori, Y.; Nakamura, K.; Fujiwara, T.; Boland, C.R.; Goel, A. Accuracy of four mononucleotide-repeat markers for the identification of DNA mismatch-repair deficiency in solid tumors. J. Transl. Med. 2018, 16, 1-N.PAG. [Google Scholar] [CrossRef] [Green Version]
  116. O’Brien, M.J.; Yang, S.; Mack, C.; Xu, H.; Huang, C.S.; Mulcahy, E.; Amorosino, M.; Farraye, F.A. Comparison of microsatellite instability, CpG island methylation phenotype, BRAF and KRAS status in serrated polyps and traditional adenomas indicates separate pathways to distinct colorectal carcinoma end points. Am. J. Surg. Pathol. 2006, 30, 1491–1501. [Google Scholar] [CrossRef]
  117. Domingo, E.; Niessen, R.C.; Oliveira, C.; Alhopuro, P.; Moutinho, C.; Espín, E.; Armengol, M.; Sijmons, R.H.; Kleibeuker, J.H.; Seruca, R.; et al. BRAF-V600E is not involved in the colorectal tumorigenesis of HNPCC in patients with functional MLH1 and MSH2 genes. Oncogene 2005, 24, 3995–3998. [Google Scholar] [CrossRef] [Green Version]
  118. Frouws, M.A.; Reimers, M.S.; Swets, M.; Bastiaannet, E.; Prinse, B.; van Eijk, R.; Lemmens, V.E.P.P.; van Herk-Sukel, M.P.P.; van Wezel, T.; Kuppen, P.J.K.; et al. The Influence of BRAF and KRAS Mutation Status on the Association between Aspirin Use and Survival after Colon Cancer Diagnosis. PLoS ONE 2017, 12, 1–12. [Google Scholar] [CrossRef]
  119. Venderbosch, S.; Nagtegaal, I.D.; Maughan, T.S.; Smith, C.G.; Cheadle, J.P.; Fisher, D.; Kaplan, R.; Quirke, P.; Seymour, M.T.; Richman, S.D.; et al. Mismatch repair status and BRAF mutation status in metastatic colorectal cancer patients: A pooled analysis of the CAIRO, CAIRO2, COIN, and FOCUS studies. Clin. Cancer Res. 2014, 20, 5322–5330. [Google Scholar] [CrossRef] [Green Version]
  120. Wang, Y.; Loree, J.M.; Yu, C.; Tschautscher, M.; Briggler, A.M.; Overman, M.J.; Broaddus, R.; Meric-Bernstam, F.; Jones, J.C.; Balcom, J.; et al. Distinct impacts of KRAS, NRAS and BRAF mutations on survival of patients with metastatic colorectal cancer. J. Clin. Oncol. 2018, 36, 3513. [Google Scholar] [CrossRef]
  121. Roth, A.D.; Tejpar, S.; Delorenzi, M.; Yan, P.; Fiocca, R.; Klingbiel, D.; Dietrich, D.; Biesmans, B.; Bodoky, G.; Barone, C.; et al. Prognostic role of KRAS and BRAF in stage II and III resected colon cancer: Results of the translational study on the PETACC-3, EORTC 40993, SAKK 60-00 trial. J. Clin. Oncol. 2010, 28, 466–474. [Google Scholar] [CrossRef]
  122. Manthravadi, S.; Sun, W.; Saeed, A. Prognostic impact of BRAF V600E mutation in patients with non-metastatic colorectal cancer with microsatellite instability: A systematic review and meta-analysis. J. Clin. Oncol. 2018, 36, 3597. [Google Scholar] [CrossRef]
  123. Tosi, F.; Magni, E.; Amatu, A.; Mauri, G.; Bencardino, K.; Truini, M.; Veronese, S.; De Carlis, L.; Ferrari, G.; Nichelatti, M.; et al. Effect of KRAS and BRAF Mutations on Survival of Metastatic Colorectal Cancer After Liver Resection: A Systematic Review and Meta-Analysis. Clin. Colorectal Cancer 2017, 16, e153–e163. [Google Scholar] [CrossRef] [PubMed]
  124. Cremolini, C.; Di Bartolomeo, M.; Amatu, A.; Antoniotti, C.; Moretto, R.; Berenato, R.; Perrone, F.; Tamborini, E.; Aprile, G.; Lonardi, S.; et al. BRAF codons 594 and 596 mutations identify a new molecular subtype of metastatic colorectal cancer at favorable prognosis. Ann. Oncol. 2015, 26, 2092–2097. [Google Scholar] [CrossRef] [PubMed]
  125. Shimada, Y.; Tajima, Y.; Nagahashi, M.; Ichikawa, H.; Oyanagi, H.; Okuda, S.; Takabe, K.; Wakai, T. Clinical Significance of BRAF Non-V600E Mutations in Colorectal Cancer: A Retrospective Study of Two Institutions. J. Surg. Res. 2018, 232, 72–81. [Google Scholar] [CrossRef]
  126. Jones, J.C.; Renfro, L.A.; Al-Shamsi, H.O.; Schrock, A.B.; Rankin, A.; Zhang, B.Y.; Kasi, P.M.; Voss, J.S.; Leal, A.D.; Sun, J.; et al. (Non-V600) BRAF Mutations Define a Clinically Distinct Molecular Subtype of Metastatic Colorectal Cancer. J. Clin. Oncol. 2017, 35, 2624–2630. [Google Scholar] [CrossRef]
  127. Wang, Y.; Liu, D.; Jin, X.; Song, H.; Lou, G. Genome-wide characterization of aberrant DNA methylation patterns and the potential clinical implications in patients with endometrial cancer. Pathol. Res. Pract. 2019, 215, 137–143. [Google Scholar] [CrossRef]
  128. González-Ramírez, I.; García-Cuellar, C.; Sánchez-Pérez, Y.; Granados-García, M. DNA methylation in oral squamous cell carcinoma: Molecular mechanisms and clinical implications. Oral Dis. 2011, 17, 771–778. [Google Scholar] [CrossRef]
  129. Magzoub, M.M.; Prunello, M.; Brennan, K.; Gevaert, O. The impact of DNA methylation on the cancer proteome. PLoS Comput. Biol. 2019, 15, 1–19. [Google Scholar] [CrossRef] [Green Version]
  130. Bastian, P.J.; Ellinger, J.; Heukamp, L.C.; Kahl, P.; Müller, S.C.; von Rücker, A. Prognostic Value of CpG Island Hypermethylation at PTGS2, RAR-beta, EDNRB, and Other Gene Loci in Patients Undergoing Radical Prostatectomy. Eur. Urol. 2007, 51, 665–674. [Google Scholar] [CrossRef]
  131. Hesson, L.B.; Wilson, R.; Morton, D.; Adams, C.; Walker, M.; Maher, E.R.; Latif, F. CpG island promoter hypermethylation of a novel Ras-effector gene RASSF2A is an early event in colon carcinogenesis and correlates inversely with K-ras mutations. Oncogene 2005, 24, 3987–3994. [Google Scholar] [CrossRef] [Green Version]
  132. Cohen, Y.; Merhavi-Shoham, E.; Avraham, R.B.; Frenkel, S.; Pe’er, J.; Goldenberg-Cohen, N. Hypermethylation of CpG island loci of multiple tumor suppressor genes in retinoblastoma. Exp. Eye Res. 2008, 86, 201–206. [Google Scholar] [CrossRef] [PubMed]
  133. Puccini, A.; Berger, M.D.; Naseem, M.; Tokunaga, R.; Battaglin, F.; Cao, S.; Hanna, D.L.; McSkane, M.; Soni, S.; Zhang, W.; et al. Colorectal cancer: Epigenetic alterations and their clinical implications. BBA Rev. Cancer 2017, 1868, 439–448. [Google Scholar] [CrossRef] [PubMed]
  134. Weisenberger, D.J.; Siegmund, K.D.; Campan, M.; Young, J.; Long, T.I.; Faasse, M.A.; Kang, G.H.; Widschwendter, M.; Weener, D.; Buchanan, D.; et al. CpG island methylator phenotype underlies sporadic microsatellite instability and is tightly associated with BRAF mutation in colorectal cancer. Nat. Genet. 2006, 38, 787–793. [Google Scholar] [CrossRef] [PubMed]
  135. Ogino, S.; Kawasaki, T.; Kirkner, G.J.; Suemoto, Y.; Meyerhardt, J.A.; Fuchs, C.S. Molecular correlates with MGMT promoter methylation and silencing support CpG island methylator phenotype-low (CIMP-low) in colorectal cancer. Gut 2007, 56, 1564–1571. [Google Scholar] [CrossRef] [Green Version]
  136. Shen, L.; Toyota, M.; Kondo, Y.; Lin, E.; Zhang, L.; Guo, Y.; Hernandez, N.S.; Chen, X.; Ahmed, S.; Konishi, K.; et al. Integrated genetic and epigenetic analysis identifies three different subclasses of colon cancer. Proc. Natl. Acad. Sci. USA 2007, 104, 18654–18659. [Google Scholar] [CrossRef] [Green Version]
  137. Ogino, S.; Cantor, M.; Kawasaki, I.; Brahmandam, M.; Kirkner, G.J.; Weisenberger, D.J.; Campan, M.; Laird, P.W.; Loda, M.; Fuchs, C.S. CpG island methylator phenotype (CIMP) of colorectal cancer is best characterised by quantitative DNA methylation analysis and prospective cohort studies. Gut 2006, 55, 1000–1006. [Google Scholar] [CrossRef]
  138. Suzuki, H.; Yamamoto, E.; Maruyama, R.; Niinuma, T.; Kai, M. Biological significance of the CpG island methylator phenotype. Biochem. Biophys. Res. Commun. 2014, 455, 35–42. [Google Scholar] [CrossRef]
  139. Suzuki, H.; Igarashi, S.; Nojima, M.; Maruyama, R.; Yamamoto, E.; Kai, M.; Akashi, H.; Watanabe, Y.; Yamamoto, H.; Sasaki, Y.; et al. IGFBP7 is a p53-responsive gene specifically silenced in colorectal cancer with CpG island methylator phenotype. Carcinogenesis 2010, 31, 342–349. [Google Scholar] [CrossRef] [Green Version]
  140. Hanon, B.M.; Al-Mohaimen Mohammad, N.A.; Mahmood, A.S. CpG Island Methylator Phenotype (CIMP) Correlation with Clinical and Morphological Feature of Colorectal Cancer in Iraq patients. Pan Arab J. Oncol. 2015, 8, 6–13. [Google Scholar]
  141. Shen, L.; Kondo, Y.; Yi, G.; Jiexin, Z.; Li, Z.; Ahmed, S.; Jingmin, S.; Xinli, C.; Waterland, R.A.; Issa, J.-P.J. Genome-Wide Profiling of DNA Methylation Reveals a Class of Normally Methylated CpG Island Promoters. PLoS Genet. 2007, 3, e181–e2036. [Google Scholar] [CrossRef] [Green Version]
  142. Suzuki, M.; Sato, S.; Arai, Y.; Shinohara, T.; Tanaka, S.; Greally, J.M.; Hattori, N.; Shiota, K. A new class of tissue-specifically methylated regions involving entire CpG islands in the mouse. Genes Cells 2007, 12, 1305–1314. [Google Scholar] [CrossRef] [PubMed]
  143. Rigal, M.; Kevei, Z.; Pélissier, T.; Mathieu, O. DNA methylation in an intron of the IBM1 histone demethylase gene stabilizes chromatin modification patterns. EMBO J. 2012, 31, 2981–2993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Zee, B.M.; Dibona, A.B.; Alekseyenko, A.A.; French, C.A.; Kuroda, M.I. The Oncoprotein BRD4-NUT Generates Aberrant Histone Modification Patterns. PLoS ONE 2016, 11, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Fuminori, S.; Yoshikuni, I.; Masamichi, H.; Yasuhiro, K.; Shuji, N. Epigenetic modulation associated with carcinogenesis and prognosis of human gastric cancer (Review). Oncol. Lett. 2017, 13, 3363–3368. [Google Scholar] [CrossRef] [Green Version]
  146. Arteaga, C.L.; Engelman, J.A. ERBB receptors: From oncogene discovery to basic science to mechanism-based cancer therapeutics. Cancer Cell 2014, 25, 282–303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Tebbutt, N.; Pedersen, M.W.; Johns, T.G. Targeting the ERBB family in cancer: Couples therapy. Nat. Rev. Cancer 2013, 13, 663–673. [Google Scholar] [CrossRef]
  148. Roskoski, R. Small molecule inhibitors targeting the EGFR/ErbB family of protein-tyrosine kinases in human cancers. Pharmacolog. Res. 2019, 139, 395–411. [Google Scholar] [CrossRef]
  149. Vecchione, L.; Jacobs, B.; Normanno, N.; Ciardiello, F.; Tejpar, S. EGFR-targeted therapy. Exp. Cell Res. 2011, 317, 2765–2771. [Google Scholar] [CrossRef]
  150. Wang, Z. ErbB Receptors and Cancer. Methods Mol. Biol. 2017, 1652, 3–35. [Google Scholar] [CrossRef]
  151. Hsu, J.L.; Hung, M.C. The role of HER2, EGFR, and other receptor tyrosine kinases in breast cancer. Cancer Metastasis Rev. 2016, 35, 575–588. [Google Scholar] [CrossRef] [Green Version]
  152. Roskoski, R. The ErbB/HER family of protein-tyrosine kinases and cancer. Pharmacolog. Res. 2014, 79, 34–74. [Google Scholar] [CrossRef] [PubMed]
  153. Pearson, G.; Robinson, F.; Beers Gibson, T.; Xu, B.E.; Karandikar, M.; Berman, K.; Cobb, M.H. Mitogen-activated protein (MAP) kinase pathways: Regulation and physiological functions. Endocr. Rev. 2001, 22, 153–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Wu, P.; Wee, P.; Jiang, J.; Chen, X.; Wang, Z. Differential regulation of transcription factors by location-specific EGF receptor signaling via a spatio-temporal interplay of ERK activation. PLoS ONE 2012, 7, e41354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Okano, J.; Gaslightwala, I.; Birnbaum, M.J.; Rustgi, A.K.; Nakagawa, H. Akt/protein kinase B isoforms are differentially regulated by epidermal growth factor stimulation. J. Biol. Chem. 2000, 275, 30934–30942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Schubbert, S.; Shannon, K.; Bollag, G. Hyperactive Ras in developmental disorders and cancer. Nat. Rev. Cancer 2007, 7, 295–308. [Google Scholar] [CrossRef]
  157. Zhang, X.; Tang, N.; Hadden, T.J.; Rishi, A.K. Akt, FoxO and regulation of apoptosis. Biochim. Biophys. Acta 2011, 1813, 1978–1986. [Google Scholar] [CrossRef] [Green Version]
  158. Bos, J.L.; Fearon, E.R.; Hamilton, S.R.; Verlaan-de Vries, M.; van Boom, J.H.; van der Eb, A.J.; Vogelstein, B. Prevalence of ras gene mutations in human colorectal cancers. Nature 1987, 327, 293–297. [Google Scholar] [CrossRef]
  159. Neumann, J.; Zeindl-Eberhart, E.; Kirchner, T.; Jung, A. Frequency and type of KRAS mutations in routine diagnostic analysis of metastatic colorectal cancer. Pathol. Res. Pract. 2009, 205, 858–862. [Google Scholar] [CrossRef]
  160. Kosmidou, V.; Oikonomou, E.; Vlassi, M.; Avlonitis, S.; Katseli, A.; Tsipras, I.; Mourtzoukou, D.; Kontogeorgos, G.; Zografos, G.; Pintzas, A. Tumor heterogeneity revealed by KRAS, BRAF, and PIK3CA pyrosequencing: KRAS and PIK3CA intratumor mutation profile differences and their therapeutic implications. Hum. Mutat. 2014, 35, 329–340. [Google Scholar] [CrossRef]
  161. Wennerberg, K.; Rossman, K.L.; Der, C.J. The Ras superfamily at a glance. J. Cell Sci. 2005, 118, 843–846. [Google Scholar] [CrossRef] [Green Version]
  162. Forrester, K.; Almoguera, C.; Han, K.; Grizzle, W.E.; Perucho, M. Detection of high incidence of K-ras oncogenes during human colon tumorigenesis. Nature 1987, 327, 298–303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Fernández-Medarde, A.; Santos, E. Ras in cancer and developmental diseases. Genes Cancer 2011, 2, 344–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Imamura, Y.; Morikawa, T.; Liao, X.; Lochhead, P.; Kuchiba, A.; Yamauchi, M.; Qian, Z.R.; Nishihara, R.; Meyerhardt, J.A.; Haigis, K.M.; et al. Specific mutations in KRAS codons 12 and 13, and patient prognosis in 1075 BRAF wild-type colorectal cancers. Clin. Cancer Res. 2012, 18, 4753–4763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Conlin, A.; Smith, G.; Carey, F.A.; Wolf, C.R.; Steele, R.J.C. The prognostic significance of K-ras, p53, and APC mutations in colorectal carcinoma. Gut 2005, 54, 1283–1286. [Google Scholar] [CrossRef] [Green Version]
  166. Phipps, A.I.; Buchanan, D.D.; Makar, K.W.; Win, A.K.; Baron, J.A.; Lindor, N.M.; Potter, J.D.; Newcomb, P.A. KRAS-mutation status in relation to colorectal cancer survival: The joint impact of correlated tumour markers. Br. J. Cancer 2013, 108, 1757–1764. [Google Scholar] [CrossRef] [Green Version]
  167. Samatar, A.A.; Poulikakos, P.I. Targeting RAS-ERK signalling in cancer: Promises and challenges. Nat. Rev. Drug Discov. 2014, 13, 928–942. [Google Scholar] [CrossRef]
  168. Davies, H.; Bignell, G.R.; Cox, C.; Stephens, P.; Edkins, S.; Clegg, S.; Teague, J.; Woffendin, H.; Garnett, M.J.; Bottomley, W.; et al. Mutations of the BRAF gene in human cancer. Nature 2002, 417, 949–954. [Google Scholar] [CrossRef]
  169. Iacopetta, B.; Li, W.Q.; Grieu, F.; Ruszkiewicz, A.; Kawakami, K. BRAF mutation and gene methylation frequencies of colorectal tumours with microsatellite instability increase markedly with patient age. Gut 2006, 55, 1213–1214. [Google Scholar] [CrossRef] [Green Version]
  170. Caputo, F.; Santini, C.; Bardasi, C.; Cerma, K.; Casadei-Gardini, A.; Spallanzani, A.; Andrikou, K.; Cascinu, S.; Gelsomino, F. BRAF-Mutated Colorectal Cancer: Clinical and Molecular Insights. Int. J. Mol. Sci. 2019, 20, 5369. [Google Scholar] [CrossRef] [Green Version]
  171. Prahallad, A.; Sun, C.; Huang, S.; Di Nicolantonio, F.; Salazar, R.; Zecchin, D.; Beijersbergen, R.L.; Bardelli, A.; Bernards, R. Unresponsiveness of colon cancer to BRAF(V600E) inhibition through feedback activation of EGFR. Nature 2012, 483, 100–103. [Google Scholar] [CrossRef] [Green Version]
  172. Lopez, A.; Harada, K.; Vasilakopoulou, M.; Shanbhag, N.; Ajani, J.A. Targeting Angiogenesis in Colorectal Carcinoma. Drugs 2019, 79, 63–74. [Google Scholar] [CrossRef] [PubMed]
  173. Guba, M.; Seeliger, H.; Kleespies, A.; Jauch, K.W.; Bruns, C. Vascular endothelial growth factor in colorectal cancer. Int. J. Colorectal Dis. 2004, 19, 510–517. [Google Scholar] [CrossRef]
  174. Amelio, I.; Melino, G. The p53 family and the hypoxia-inducible factors (HIFs): Determinants of cancer progression. Trends Biochem. Sci. 2015, 40, 425–434. [Google Scholar] [CrossRef] [PubMed]
  175. Ghosh, N.; Hossain, U.; Mandal, A.; Sil, P.C. The Wnt signaling pathway: A potential therapeutic target against cancer. Ann. N. Y. Acad. Sci. 2019, 1443, 54–74. [Google Scholar] [CrossRef] [PubMed]
  176. Dihlmann, S.; Siermann, A.; von Knebel Doeberitz, M. The nonsteroidal anti-inflammatory drugs aspirin and indomethacin attenuate β-catenin/TCF-4 signaling. Oncogene 2001, 20, 645–653. [Google Scholar] [CrossRef] [Green Version]
  177. Tuynman, J.B.; Vermeulen, L.; Boon, E.M.; Kemper, K.; Zwinderman, A.H.; Peppelenbosch, M.P.; Richel, D.J. Cyclooxygenase-2 inhibition inhibits c-Met kinase activity and Wnt activity in colon cancer. Cancer Res. 2008, 68, 1213–1220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Giardiello, F.M.; Hamilton, S.R.; Krush, A.J.; Piantadosi, S.; Hylind, L.M.; Celano, P.; Booker, S.V.; Robinson, C.R.; Offerhaus, G.J. Treatment of colonic and rectal adenomas with sulindac in familial adenomatous polyposis. N. Engl. J. Med. 1993, 328, 1313–1316. [Google Scholar] [CrossRef] [PubMed]
  179. Steinbach, G.; Lynch, P.M.; Phillips, R.K.; Wallace, M.H.; Hawk, E.; Gordon, G.B.; Wakabayashi, N.; Saunders, B.; Shen, Y.; Fujimura, T.; et al. The effect of celecoxib, a cyclooxygenase-2 inhibitor, in familial adenomatous polyposis. N. Engl. J. Med. 2000, 342, 1946–1952. [Google Scholar] [CrossRef]
  180. Thorne, C.A.; Hanson, A.J.; Schneider, J.; Tahinci, E.; Orton, D.; Cselenyi, C.S.; Jernigan, K.K.; Meyers, K.C.; Hang, B.I.; Waterson, A.G.; et al. Small-molecule inhibition of Wnt signaling through activation of casein kinase 1α. Nat. Chem. Biol. 2010, 6, 829–836. [Google Scholar] [CrossRef] [Green Version]
  181. Huang, S.-M.A.; Mishina, Y.M.; Liu, S.; Cheung, A.; Stegmeier, F.; Michaud, G.A.; Charlat, O.; Wiellette, E.; Zhang, Y.; Wiessner, S.; et al. Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature 2009, 461, 614–620. [Google Scholar] [CrossRef]
  182. Chen, S.; Guttridge, D.C.; You, Z.; Zhang, Z.; Fribley, A.; Mayo, M.W.; Kitajewski, J.; Wang, C.Y. Wnt-1 signaling inhibits apoptosis by activating beta-catenin/T cell factor-mediated transcription. J. Cell Biol. 2001, 152, 87–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Wei, W.; Chua, M.S.; Grepper, S.; So, S.K. Soluble Frizzled-7 receptor inhibits Wnt signaling and sensitizes hepatocellular carcinoma cells towards doxorubicin. Mol. Cancer 2011, 10, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Pode-Shakked, N.; Harari-Steinberg, O.; Haberman-Ziv, Y.; Rom-Gross, E.; Bahar, S.; Omer, D.; Metsuyanim, S.; Buzhor, E.; Jacob-Hirsch, J.; Goldstein, R.S.; et al. Resistance or sensitivity of Wilms’ tumor to anti-FZD7 antibody highlights the Wnt pathway as a possible therapeutic target. Oncogene 2011, 30, 1664–1680. [Google Scholar] [CrossRef] [PubMed]
  185. Hooper, C.; Killick, R.; Fernandes, C.; Sugden, D.; Lovestone, S. Transcriptomic profiles of Wnt3a and insulin in primary cultured rat cortical neurones. J. Neurochem. 2011, 118, 512–520. [Google Scholar] [CrossRef]
  186. Jeremy, J.Y.; Thompson, C.S.; Mikhailidis, D.P.; Dandona, P. Experimental diabetes mellitus inhibits prostacyclin synthesis by the rat penis: Pathological implications. Diabetologia 1985, 28, 365–368. [Google Scholar] [CrossRef] [Green Version]
  187. Lavergne, E.; Hendaoui, I.; Coulouarn, C.; Ribault, C.; Leseur, J.; Eliat, P.A.; Mebarki, S.; Corlu, A.; Clément, B.; Musso, O. Blocking Wnt signaling by SFRP-like molecules inhibits in vivo cell proliferation and tumor growth in cells carrying active β-catenin. Oncogene 2011, 30, 423–433. [Google Scholar] [CrossRef] [Green Version]
  188. Boulay, J.L.; Mild, G.; Lowy, A.; Reuter, J.; Lagrange, M.; Terracciano, L.; Laffer, U.; Herrmann, R.; Rochlitz, C. SMAD4 is a predictive marker for 5-fluorouracil-based chemotherapy in patients with colorectal cancer. Br. J. Cancer 2002, 87, 630. [Google Scholar] [CrossRef] [Green Version]
  189. Lin, Z.; Zhang, L.; Zhou, J.; Zheng, J. Silencing Smad4 attenuates sensitivity of colorectal cancer cells to cetuximab by promoting epithelial-mesenchymal transition. Mol. Med. Res. 2019, 20, 3735–3745. [Google Scholar] [CrossRef] [Green Version]
  190. Oyanagi, H.; Shimada, Y.; Nagahashi, M.; Ichikawa, H.; Tajima, Y.; Abe, K.; Nakano, M.; Kameyama, H.; Takii, Y.; Kawasaki, T.; et al. SMAD4 alteration associates with invasive-front pathological markers and poor prognosis in colorectal cancer. Histopathology 2019, 74, 873–882. [Google Scholar] [CrossRef] [Green Version]
  191. Strosberg, J.R.; Yeatman, T.; Weber, J.; Coppola, D.; Schell, M.J.; Han, G.; Almhanna, K.; Kim, R.; Valone, T.; Jump, H.; et al. A phase II study of RO4929097 in metastatic colorectal cancer. Eur. J. Cancer 2012, 48, 997–1003. [Google Scholar] [CrossRef] [Green Version]
  192. Berlin, J.; Bendell, J.C.; Hart, L.L.; Firdaus, I.; Gore, I.; Hermann, R.C.; Mulcahy, M.F.; Zalupski, M.M.; Mackey, H.M.; Yauch, R.L.; et al. A randomized phase II trial of vismodegib versus placebo with FOLFOX or FOLFIRI and bevacizumab in patients with previously untreated metastatic colorectal cancer. Clin. Cancer Res. 2013, 19, 258–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Le, D.T.; Uram, J.N.; Wang, H.; Bartlett, B.R.; Kemberling, H.; Eyring, A.D.; Skora, A.D.; Luber, B.S.; Azad, N.S.; Laheru, D.; et al. PD-1 Blockade in Tumors with Mismatch-Repair Deficiency. N. Engl. J. Med. 2015, 372, 2509–2520. [Google Scholar] [CrossRef] [PubMed]
  194. O’Neil, B.H.; Wallmark, J.M.; Lorente, D.; Elez, E.; Raimbourg, J.; Gomez-Roca, C.; Ejadi, S.; Piha-Paul, S.A.; Stein, M.N.; Abdul Razak, A.R.; et al. Safety and antitumor activity of the anti-PD-1 antibody pembrolizumab in patients with advanced colorectal carcinoma. PLoS ONE 2017, 12, e0189848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Long, G.V.; Atkinson, V.; Cebon, J.S.; Jameson, M.B.; Fitzharris, B.M.; McNeil, C.M.; Hill, A.G.; Ribas, A.; Atkins, M.B.; Thompson, J.A.; et al. Standard-dose pembrolizumab in combination with reduced-dose ipilimumab for patients with advanced melanoma (KEYNOTE-029): An open-label, phase 1b trial. Lancet Oncol. 2017, 18, 1202–1210. [Google Scholar] [CrossRef] [Green Version]
  196. Overman, M.J.; McDermott, R.; Leach, J.L.; Lonardi, S.; Lenz, H.J.; Morse, M.A.; Desai, J.; Hill, A.; Axelson, M.; Moss, R.A.; et al. Nivolumab in patients with metastatic DNA mismatch repair-deficient or microsatellite instability-high colorectal cancer (CheckMate 142): An open-label, multicentre, phase 2 study. Lancet Oncol. 2017, 18, 1182–1191. [Google Scholar] [CrossRef]
  197. Lenz, H.J.J.; Van Cutsem, E.; Limon, M.L.; Wong, K.Y.; Hendlisz, A.; Aglietta, M.; Garcia-Alfonso, P.; Neyns, B.; Luppi, G.; Cardin, D.; et al. Durable clinical benefit with nivolumab (NIVO) plus low-dose ipilimumab (IPI) as first-line therapy in microsatellite instability-high/mismatch repair deficient (MSI-H/dMMR) metastatic colorectal cancer (mCRC). Ann. Oncol. 2018, 29, viii714. [Google Scholar] [CrossRef]
  198. Xie, Y.-H.; Chen, Y.-X.; Fang, J.-Y. Comprehensive review of targeted therapy for colorectal cancer. Signal Transduct. Target. Ther. 2020, 5, 22. [Google Scholar] [CrossRef]
  199. Zhou, E.; Huang, Q.; Wang, J.; Fang, C.; Yang, L.; Zhu, M.; Chen, J.; Chen, L.; Dong, M. Up-regulation of Tim-3 is associated with poor prognosis of patients with colon cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 8018–8027. [Google Scholar]
  200. Yu, X.; Huang, X.; Chen, X.; Liu, J.; Wu, C.; Pu, Q.; Wang, Y.; Kang, X.; Zhou, L. Characterization of a novel anti-human lymphocyte activation gene 3 (LAG-3) antibody for cancer immunotherapy. mABs 2019, 11, 1139–1148. [Google Scholar] [CrossRef]
  201. Mendelsohn, J.; Prewett, M.; Rockwell, P.; Goldstein, N.I. CCR 20th anniversary commentary: A chimeric antibody, C225, inhibits EGFR activation and tumor growth. Clin. Cancer Res. 2015, 21, 227–229. [Google Scholar] [CrossRef] [Green Version]
  202. Jonker, D.J.; O’Callaghan, C.J.; Karapetis, C.S.; Zalcberg, J.R.; Tu, D.; Au, H.J.; Berry, S.R.; Krahn, M.; Price, T.; Simes, R.J.; et al. Cetuximab for the treatment of colorectal cancer. N. Engl. J. Med. 2007, 357, 2040–2048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Van Cutsem, E.; Köhne, C.H.; Hitre, E.; Zaluski, J.; Chang Chien, C.R.; Makhson, A.; D’Haens, G.; Pintér, T.; Lim, R.; Bodoky, G.; et al. Cetuximab and chemotherapy as initial treatment for metastatic colorectal cancer. N. Engl. J. Med. 2009, 360, 1408–1417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zhao, B.; Wang, L.; Qiu, H.; Zhang, M.; Sun, L.; Peng, P.; Yu, Q.; Yuan, X. Mechanisms of resistance to anti-EGFR therapy in colorectal cancer. Oncotarget 2017, 8, 3980–4000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Yarom, N.; Jonker, D.J. The role of the epidermal growth factor receptor in the mechanism and treatment of colorectal cancer. Discov. Med. 2011, 11, 95–105. [Google Scholar]
  206. Douillard, J.Y.; Siena, S.; Cassidy, J.; Tabernero, J.; Burkes, R.; Barugel, M.; Humblet, Y.; Bodoky, G.; Cunningham, D.; Jassem, J.; et al. Randomized, phase III trial of panitumumab with infusional fluorouracil, leucovorin, and oxaliplatin (FOLFOX4) versus FOLFOX4 alone as first-line treatment in patients with previously untreated metastatic colorectal cancer: The PRIME study. J. Clin. Oncol. 2010, 28, 4697–4705. [Google Scholar] [CrossRef]
  207. Douillard, J.Y.; Siena, S.; Cassidy, J.; Tabernero, J.; Burkes, R.; Barugel, M.; Humblet, Y.; Bodoky, G.; Cunningham, D.; Jassem, J.; et al. Final results from PRIME: Randomized phase III study of panitumumab with FOLFOX4 for first-line treatment of metastatic colorectal cancer. Ann. Oncol. 2014, 25, 1346–1355. [Google Scholar] [CrossRef]
  208. Modest, D.P.; Rivera, F.; Bachet, J.B.; de Braud, F.; Pietrantonio, F.; Koukakis, R.; Demonty, G.; Douillard, J.Y. Panitumumab-based maintenance after oxaliplatin discontinuation in metastatic colorectal cancer: A retrospective analysis of two randomised trials. Int. J. Cancer 2019, 145, 576–585. [Google Scholar] [CrossRef]
  209. Pietrantonio, F.; Morano, F.; Corallo, S.; Lonardi, S.; Cremolini, C.; Rimassa, L.; Sartore-Bianchi, A.; Tampellini, M.; Bustreo, S.; Clavarezza, M.; et al. First-line FOLFOX plus panitumumab (Pan) followed by 5FU/LV plus Pan or single-agent Pan as maintenance therapy in patients with RAS wild-type metastatic colorectal cancer (mCRC): The VALENTINO study. J. Clin. Oncol. 2018, 36, 3505. [Google Scholar] [CrossRef]
  210. Price, T.J.; Peeters, M.; Kim, T.W.; Li, J.; Cascinu, S.; Ruff, P.; Suresh, A.S.; Thomas, A.; Tjulandin, S.; Zhang, K.; et al. Panitumumab versus cetuximab in patients with chemotherapy-refractory wild-type KRAS exon 2 metastatic colorectal cancer (ASPECCT): A randomised, multicentre, open-label, non-inferiority phase 3 study. Lancet Oncol. 2014, 15, 569–579. [Google Scholar] [CrossRef] [Green Version]
  211. Canon, J.; Rex, K.; Saiki, A.Y.; Mohr, C.; Cooke, K.; Bagal, D.; Gaida, K.; Holt, T.; Knutson, C.G.; Koppada, N.; et al. The clinical KRAS(G12C) inhibitor AMG 510 drives anti-tumour immunity. Nature 2019, 575, 217–223. [Google Scholar] [CrossRef]
  212. Hallin, J.; Engstrom, L.D.; Hargis, L.; Calinisan, A.; Aranda, R.; Briere, D.M.; Sudhakar, N.; Bowcut, V.; Baer, B.R.; Ballard, J.A.; et al. The KRAS(G12C) Inhibitor MRTX849 Provides Insight toward Therapeutic Susceptibility of KRAS-Mutant Cancers in Mouse Models and Patients. Cancer Discov. 2020, 10, 54–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Janes, M.R.; Zhang, J.; Li, L.S.; Hansen, R.; Peters, U.; Guo, X.; Chen, Y.; Babbar, A.; Firdaus, S.J.; Darjania, L.; et al. Targeting KRAS Mutant Cancers with a Covalent G12C-Specific Inhibitor. Cell 2018, 172, 578.e517–589.e517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Lito, P.; Pratilas, C.A.; Joseph, E.W.; Tadi, M.; Halilovic, E.; Zubrowski, M.; Huang, A.; Wong, W.L.; Callahan, M.K.; Merghoub, T.; et al. Relief of profound feedback inhibition of mitogenic signaling by RAF inhibitors attenuates their activity in BRAFV600E melanomas. Cancer Cell 2012, 22, 668–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Corcoran, R.B.; Ebi, H.; Turke, A.B.; Coffee, E.M.; Nishino, M.; Cogdill, A.P.; Brown, R.D.; Della Pelle, P.; Dias-Santagata, D.; Hung, K.E.; et al. EGFR-mediated re-activation of MAPK signaling contributes to insensitivity of BRAF mutant colorectal cancers to RAF inhibition with vemurafenib. Cancer Discov. 2012, 2, 227–235. [Google Scholar] [CrossRef] [Green Version]
  216. Corcoran, R.B.; André, T.; Yoshino, T.; Bendell, J.C.; Atreya, C.E.; Schellens, J.H.M.; Ducreux, M.P.; McRee, A.; Siena, S.; Middleton, G.; et al. Efficacy and circulating tumor DNA (ctDNA) analysis of the BRAF inhibitor dabrafenib (D), MEK inhibitor trametinib (T), and anti-EGFR antibody panitumumab (P) in patients (pts) with BRAF V600E–mutated (BRAFm) metastatic colorectal cancer (mCRC). Ann. Oncol. 2016, 27, vi150. [Google Scholar] [CrossRef]
  217. Van Geel, R.; Tabernero, J.; Elez, E.; Bendell, J.C.; Spreafico, A.; Schuler, M.; Yoshino, T.; Delord, J.P.; Yamada, Y.; Lolkema, M.P.; et al. A Phase Ib Dose-Escalation Study of Encorafenib and Cetuximab with or without Alpelisib in Metastatic BRAF-Mutant Colorectal Cancer. Cancer Discov. 2017, 7, 610–619. [Google Scholar] [CrossRef] [Green Version]
  218. Corcoran, R.B.; André, T.; Atreya, C.E.; Schellens, J.H.M.; Yoshino, T.; Bendell, J.C.; Hollebecque, A.; McRee, A.J.; Siena, S.; Middleton, G.; et al. Combined BRAF, EGFR, and MEK Inhibition in Patients with BRAF(V600E)-Mutant Colorectal Cancer. Cancer Discov. 2018, 8, 428–443. [Google Scholar] [CrossRef] [Green Version]
  219. Hong, D.S.; Morris, V.K.; El Osta, B.; Sorokin, A.V.; Janku, F.; Fu, S.; Overman, M.J.; Piha-Paul, S.; Subbiah, V.; Kee, B.; et al. Phase IB Study of Vemurafenib in Combination with Irinotecan and Cetuximab in Patients with Metastatic Colorectal Cancer with BRAFV600E Mutation. Cancer Discov. 2016, 6, 1352–1365. [Google Scholar] [CrossRef] [Green Version]
  220. Kopetz, S.; McDonough, S.L.; Morris, V.K.; Lenz, H.-J.; Magliocco, A.M.; Atreya, C.E.; Diaz, L.A.; Allegra, C.J.; Wang, S.E.; Lieu, C.H.; et al. Randomized trial of irinotecan and cetuximab with or without vemurafenib in BRAF-mutant metastatic colorectal cancer (SWOG 1406). J. Clin. Oncol. 2017, 35, 520. [Google Scholar] [CrossRef] [Green Version]
  221. Tabernero, J.; Geel, R.V.; Guren, T.K.; Yaeger, R.D.; Spreafico, A.; Faris, J.E.; Yoshino, T.; Yamada, Y.; Kim, T.W.; Bendell, J.C.; et al. Phase 2 results: Encorafenib (ENCO) and cetuximab (CETUX) with or without alpelisib (ALP) in patients with advanced BRAF-mutant colorectal cancer (BRAFm CRC). J. Clin. Oncol. 2016, 34, 3544. [Google Scholar] [CrossRef] [Green Version]
  222. Kopetz, S.; Grothey, A.; Yaeger, R.; Van Cutsem, E.; Desai, J.; Yoshino, T.; Wasan, H.; Ciardiello, F.; Loupakis, F.; Hong, Y.S.; et al. Encorafenib, Binimetinib, and Cetuximab in BRAF V600E-Mutated Colorectal Cancer. N. Engl. J. Med. 2019, 381, 1632–1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Passardi, A.; Nanni, O.; Tassinari, D.; Turci, D.; Cavanna, L.; Fontana, A.; Ruscelli, S.; Mucciarini, C.; Lorusso, V.; Ragazzini, A.; et al. Effectiveness of bevacizumab added to standard chemotherapy in metastatic colorectal cancer: Final results for first-line treatment from the ITACa randomized clinical trial. Ann. Oncol. 2015, 26, 1201–1207. [Google Scholar] [CrossRef]
  224. Simkens, L.H.; van Tinteren, H.; May, A.; Tije, T.A.J.; Creemers, G.J.; Loosveld, O.J.; de Jongh, F.E.; Erdkamp, F.L.; Erjavec, Z.; van der Torren, A.M.; et al. Maintenance treatment with capecitabine and bevacizumab in metastatic colorectal cancer (CAIRO3): A phase 3 randomised controlled trial of the Dutch Colorectal Cancer Group. Lancet 2015, 385, 1843–1852. [Google Scholar] [CrossRef]
  225. Schwartzberg, L.S.; Rivera, F.; Karthaus, M.; Fasola, G.; Canon, J.L.; Hecht, J.R.; Yu, H.; Oliner, K.S.; Go, W.Y. PEAK: A randomized, multicenter phase II study of panitumumab plus modified fluorouracil, leucovorin, and oxaliplatin (mFOLFOX6) or bevacizumab plus mFOLFOX6 in patients with previously untreated, unresectable, wild-type KRAS exon 2 metastatic colorectal cancer. J. Clin. Oncol. 2014, 32, 2240–2247. [Google Scholar] [CrossRef] [PubMed]
  226. Venook, A.P.; Niedzwiecki, D.; Lenz, H.J.; Innocenti, F.; Fruth, B.; Meyerhardt, J.A.; Schrag, D.; Greene, C.; O’Neil, B.H.; Atkins, J.N.; et al. Effect of First-Line Chemotherapy Combined With Cetuximab or Bevacizumab on Overall Survival in Patients With KRAS Wild-Type Advanced or Metastatic Colorectal Cancer: A Randomized Clinical Trial. JAMA 2017, 317, 2392–2401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Giantonio, B.J.; Catalano, P.J.; Meropol, N.J.; O’Dwyer, P.J.; Mitchell, E.P.; Alberts, S.R.; Schwartz, M.A.; Benson, A.B., 3rd. Bevacizumab in combination with oxaliplatin, fluorouracil, and leucovorin (FOLFOX4) for previously treated metastatic colorectal cancer: Results from the Eastern Cooperative Oncology Group Study E3200. J. Clin. Oncol. 2007, 25, 1539–1544. [Google Scholar] [CrossRef]
Figure 1. Multistep genetic model for colorectal adenocarcinoma sequence. There are three pathways regulating the adenocarcinoma sequence: chromosomal instability (CIN), microsatellite instability (MSI), and CpG island methylator phenotype (CIMP) hypermethylation.
Figure 1. Multistep genetic model for colorectal adenocarcinoma sequence. There are three pathways regulating the adenocarcinoma sequence: chromosomal instability (CIN), microsatellite instability (MSI), and CpG island methylator phenotype (CIMP) hypermethylation.
Ijms 22 00130 g001
Figure 2. Schematic representation of the molecular pathways involved in CRC pathogenesis.
Figure 2. Schematic representation of the molecular pathways involved in CRC pathogenesis.
Ijms 22 00130 g002
Table 1. Biological differences in the gene expression-based molecular subtypes of colorectal cancer (CRC).
Table 1. Biological differences in the gene expression-based molecular subtypes of colorectal cancer (CRC).
SubtypeGene ExpressionPrognosisReference
CMS1
(MSI immune)
  • Deregulated DNA mismatch repair, MSI and MLH1 silencing
  • CIMP-high with common Serine/threonine-protein kinase B-Raf (BRAF) mutations and low number of SCNAs
  • Immune infiltration and activation
Very poor survival rate after relapse[4]
CMS2
(Canonical)
  • Express epithelial signatures with Wingless (Wnt) and MYC signaling activation
  • Frequently exhibit loss of TSGs and overexpression of oncogenes than the other subtypes
Better survival rate after relapse in comparison to other subtypes
CMS3
(Metabolic)
  • Fewer SCNAs as compared to CMS2 and CMS4
  • Epithelial signatures
  • Metabolic dysregulation in a variety of pathways with frequent KRAS mutations
  • Slightly higher presence of CIMP-low
Better survival rate after relapse in comparison to other subtypes
CMS4
(Mesenchymal)
  • Activation of Transforming growth factor-β (TGF-β)
  • Upregulated expression of EMT genes
  • Enhanced expression of genes regulating inflammation, matrix remodeling, stromal invasion and angiogenesis
Worse overall and relapse-free survival as compared to other subtypes
CIMP: CpG island methylator phenotype; EMT: epithelial-mesenchymal transition; MSI: microsatellite instability; SCNA: somatic copy number alterations.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. 2021, 22, 130. https://doi.org/10.3390/ijms22010130

AMA Style

Malki A, ElRuz RA, Gupta I, Allouch A, Vranic S, Al Moustafa A-E. Molecular Mechanisms of Colon Cancer Progression and Metastasis: Recent Insights and Advancements. International Journal of Molecular Sciences. 2021; 22(1):130. https://doi.org/10.3390/ijms22010130

Chicago/Turabian Style

Malki, Ahmed, Rasha Abu ElRuz, Ishita Gupta, Asma Allouch, Semir Vranic, and Ala-Eddin Al Moustafa. 2021. "Molecular Mechanisms of Colon Cancer Progression and Metastasis: Recent Insights and Advancements" International Journal of Molecular Sciences 22, no. 1: 130. https://doi.org/10.3390/ijms22010130

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