Open access peer-reviewed chapter

Epigenetic Biomarkers and Their Therapeutic Applications in Colorectal Cancer

Written By

Antja-Voy Hartley, Matthew Martin and Tao Lu

Submitted: 27 August 2018 Reviewed: 16 November 2018 Published: 09 January 2019

DOI: 10.5772/intechopen.82572

From the Edited Volume

Advances in the Molecular Understanding of Colorectal Cancer

Edited by Eva Segelov

Chapter metrics overview

1,500 Chapter Downloads

View Full Metrics

Abstract

Colorectal cancer (CRC) is one of the most aggressive cancers worldwide and is known to develop through a stepwise process involving the accumulation of several genetic and epigenetic alterations. Furthermore, numerous studies have highlighted the significant role that certain epigenetic enzymes play in CRC pathogenesis, particularly those that govern chromatin components in the promoter regions of tumor suppressors and oncogenes. Here, we delineate the relationship between CRC-associated epigenetic marks, their modifying enzymes, and the classification of CRC into distinct molecular pathways or subtypes. Moreover, we discuss some of the most prominent methyltransferases, demethylases, acetyltransferases, and deacetylases, which have been targeted for preclinical and clinical CRC treatment. Notably, inhibitors against these epigenetic enzymes are a promising new class of anticancer drugs, with several obtaining Food and Drug Administration (FDA) approval for the treatment of blood and solid tumors. By highlighting the epigenetic molecular pathways leading to CRC development as well as providing an update on current CRC epigenetic therapies, this chapter sheds fresh insight into new and emerging avenues for future therapeutics.

Keywords

  • checkpoint
  • CIMP
  • CIN
  • CRC
  • demethylase
  • DNA methylation
  • DNMT
  • epigenetics
  • HAT
  • HDAC
  • methyltransferase
  • MSI

1. Introduction

Colorectal cancer (CRC) is the third leading cause of cancer-related deaths globally and is expected to be responsible for an estimated 1.1 million deaths by 2030 [1]. With this growing global burden, prevention and treatment of CRC remains a significant public health challenge. CRC is thought to originate from sequential accumulation of genetic and epigenetic aberrations [2]. Of the identified genetic mutations, approximately 15 have been characterized as “driver mutations” and are thought to be functionally important during CRC initiation and progression [3, 4]. These include genes affecting critical cellular pathways such as those governing proliferation, apoptosis, migration, adhesion, and DNA damage and repair [3]. Importantly, it is now well established that epigenetic alterations can also serve as major driver events in the pathogenesis of CRC [5, 6, 7]. However, unlike genetic mutations, epigenetic modifications consist of heritable changes in gene expression without DNA sequence changes and are intrinsically reversible by nature. These epigenetic events include alterations in DNA methylation, histone modifications, and non-coding RNAs. Moreover, the reversibility of these modifications makes them attractive molecular targets for anticancer therapeutic interventions [3, 8].

CRC is a highly heterogeneous disease and can be classified into molecularly and pathologically distinct pathways and subtypes [9]. Moreover, these classifications have significantly influenced patient stratification, prognosis, and therapeutic response [9, 10]. In this chapter, we focus on three epigenetic-related primary molecular pathways, namely the microsatellite instability (MSI) phenotype, the chromosomal instability (CIN) phenotype, and the CpG island methylator phenotype (CIMP). Importantly, each pathway reflects the underlying mechanisms of carcinogenesis as marked by certain aberrations such as a defective DNA mismatch repair (MMR) system, which is associated with MSI CRCs [11], or by widespread promoter DNA methylation within CpG islands as is the case with CIMP tumors [11, 12]. On the other hand, the CIN pathway, which manifests in majority of CRC cases (~85%), arises through widespread chromosomal imbalances [9, 13, 14]. We also make mention of the relationship between these defined pathways and the four consensus molecular subtyping classifications, with emphasis on the frequent overlap observed between two or more of the aforementioned pathways.

In the past few decades, several studies have analyzed epigenetic marks, the enzymes mediating these marks, and the extent of their active contribution to CRC tumor development and progression [2, 3]. For instance, several methylation-related enzymes have been found to be clinically relevant to CRC [15]. Among these, some of the most prominent histone methyltransferases (HMTs) that have been targeted for preclinical and clinical treatment of CRC are discussed in Section 4 of this chapter [15, 16]. On the other hand, a comparatively less number of histone demethylases (HDMs) have been validated as pertinent to CRC pathogenesis. As important regulators of colon cell transformation, histone deacetylases (HDAC) have also emerged as prominent markers of early carcinogenic events due to their unique role in maintaining higher-order chromatin structure [17].

In this chapter, we also highlight a few chemical inhibitors relevant to epigenetic therapy. However, we also noted that among the CRC-associated epigenetic enzymes, only a few of them have potent inhibitors available [15]. This suggests that the knowledge concerning targeting these enzymes for CRC is still insufficient and needs further evaluation. For example, only a few DNA methyltransferase (DNMT) and HMT inhibitors have been used in CRC cells [2], and a handful of Food and Drug Administration (FDA)-approved HDAC inhibitors are currently being explored for the treatment of solid tumors including CRC [18].

Unfortunately, the use of such epigenetic-based inhibitors has not been without limitations. Major drawbacks, such as adverse side effects and lack of clinical efficacy, have limited their use as single agents. Therefore, many inhibitors show more promise in combination therapy with chemotherapies suggesting that the full therapeutic potential of epigenetic therapy will perhaps be best realized in combination with other anti-cancer agents [19, 20]. This is also complemented by the recent understanding that there is a strong interplay between immune and cancer cells within the tumor microenvironment [21]. Recent studies in CRC cells have shown promising combinations of epigenetic and immunomodulatory drugs. By reversing expression changes of genes involved in immune suppression and thus enhancing expression of tumor-associated antigens, cancer cells potentially become more sensitive to immune checkpoint inhibitors [22]. These and other discoveries have established a highly promising basis for studies using combined epigenetic and immunotherapeutic agents for treating CRC.

Advertisement

2. Epigenetic modifications in CRC

2.1 Histone modifications

Over the past decade, significant advances in our understanding of the CRC “epigenome” have revealed that most CRC cases harbor alterations in their histone modification states, particularly regarding aberrant histone methylation and acetylation [6, 15, 23]. Importantly, these abnormal histone marks are highly recurrent and have recently been used as biomarkers to predict the clinical outcome in CRC patients [2]. These include changes in the global patterns of specific histone modifications. For example, Tamagawa et al. showed that global changes in histone H3K4me2, H3K9ac, and H3K9me2 in metachronous liver metastasis correlated to overall survival of CRC patients [24]. Specifically, low H3K4me2 levels were shown to correlate with overall poor prognosis [24]. Likewise, other studies have identified reduced levels of H3K9me3 and H4K20me3 as diagnostic biomarkers for CRC in circulating nucleosomes which correlated with poor patient outcome [25]. Conversely, high H4K20me3 and H3K9me3, as well as low nuclear expression of H3K4me3, were associated with a better prognosis for early-stage CRC patients [26].

Furthermore, since reduction or enrichment of these marks frequently occurs at the promoters of key CRC-related oncogenes and tumor suppressors, this results in detrimental changes in gene expression that form the basis of tumorigenesis [15, 27]. For instance, H3K4me3, when found to be elevated in CRC primary tumors and cell lines, resulted in activated Wingless-type (WNT) signaling and target gene expression via interaction between SET domain-containing protein 1A (SETD1A) and β-catenin [28]. Meanwhile, another study revealed that low H3K4me1/2/3 levels were associated with hypoxia-induced silencing of MLH1 in SW480 cells, which is a key event in the DNA mismatch defects linked to the development of sporadic CRC [29]. Yokoyama’s group also demonstrated a role for the well-recognized repressive mark H3K9me3, revealing that its increased levels in metastatic CRC patient-derived cells correlated with enhanced cell motility [30]. Interestingly, this coincided with repression of Ataxia-telangiectasia mutated (ATM) and p53-associated KZNF protein (APAK), leading to a defect in p53-dependent apoptosis [30]. Moreover, enrichment of another repressive mark, H3K27me3, was associated with poor CRC patient prognosis while elevated H3K79me2 was shown to enhance interleukin (IL)-22-induced stemness in CRC cells [31, 32]. Intriguingly, more recent studies have also shown that mutations in specific methylation sites could promote CRC development. For instance, the Shah and Lu groups identified histone 3 lysine 36-to-methionine (H3K36M) substitution mutations in CRC patient samples, which promoted more undifferentiated sarcomas in vivo [33, 34]. This suggests that H3K36 methylation potentially constitutes a major tumor suppressive epigenetic mark.

In addition to abnormal methylation, disruption of histone acetylation patterns also contributes to CRC pathogenesis, particularly relating to transcriptional inactivation of tumor suppressor genes and, sometimes, activation of oncogenes. For example, Richon et al. showed that hypoacetylation at the promoter of the tumor suppressor p21WAF1 led to its repression, an effect that was reversed by inhibition of HDAC activity [35]. Conversely, mass spectrometry-based analyses used to quantify global alterations of histone modifications in CRC samples identified H3K27ac as a modification frequently upregulated in CRC [36]. In fact, one study highlighted the effects of aspirin in reducing the enrichment of H3K27ac in the promoters of inducible nitric oxide synthase (iNOS), tumor necrosis factor alpha (TNF-α), and IL-6 [37]. This in turn corresponded to the dramatic reduction of the mRNA and protein levels of these genes, which suppressed inflammatory colitis symptoms and CRC tumor burden [37]. Taken together, these studies emphasize the differential abundance of key repressive and activating histone methylation and acetylation marks in CRC and suggests their role in regulating genes associated with CRC development and progression.

2.2 DNA methylation

DNA methylation constitutes the first recognized epigenetic alteration in CRC [38]. Usually, global DNA hypomethylation is frequently seen, which occurs gradually and early in the process of CRC carcinogenesis [38]. More precisely, global DNA hypomethylation mainly takes place on cytosine guanine (CpG) dinucleotides within pericentromeric regions. Initially, this hypomethylation in CRC was hypothesized to be associated mainly with widespread oncogene activation but has now been linked predominantly to increased genomic instability [3]. This increased accumulation of chromosome breakage and overall chromosomal instability contributes to a prevalent subtype of CRCs known as the CIN phenotype as we briefly described in Section 1 [39].

Notably, hypomethylation typically occurs in concert with systematic and discrete DNA hypermethylation events at the promoters of genes involved in DNA repair, apoptosis, proliferation, angiogenesis, adhesion, and invasion [38, 40]. DNA hypermethylation is the most extensively characterized epigenetic alteration in CRC, occurring at CpG dinucleotide-dense regions, called CpG islands, which are present in about 60% of genes [6, 41]. Apart from CpG islands, DNA hypermethylation has also been extensively observed within the first exonic/intronic regions of some genes and generally results in transcriptional silencing [42]. Some of the most frequently hypermethylated genes in CRC include Adenomatous polyposis coli (APC), Cyclin-dependent kinase inhibitor 2A (p16INK4a/CDKN2A), Tissue inhibitor of metalloproteinases 3 (TIMP3), O-6-Methylguanine-DNA methyltransferase (MGMT), Secreted frizzled related protein 1 (SFRP1), Transmembrane protein with epidermal growth factor (EGF) like and two follistatin like domains 2 (TMEFF2), Heparan sulfate-glucosamine 3-sulfotransferase 2 (HS3ST2/3OST2), Ras association domain family member 1 (RASSF1A), and GATA binding protein 4 (GATA4) [43].

Another subtype of CRCs with extensive patterns of promoter methylation, known as the CIMP phenotype as described in Section 1, is also characterized by aberrant DNA methylation at genes with roles in CRC initiation or progression [44]. For instance, using a qPCR-based technique, one group identified genes with the highest percentage of methylation in CRC patients including Runt related transcription factor 3 (RUNX3), Protocadherin 10 (PCDH10), Secreted frizzled related protein 5 (SFRP5), Insulin-like growth factor 2 (IGF2), and Hepatocyte nuclear factor 1 homeobox A (HNF1) homeobox B (Hnf1b) [45]. Moreover, these genes were observed to have the most promising biomarker potential because of the frequent gene repression patterns [45]. Other commonly hypermethylated genes, such as Sex-determining region Y (SRY)-related HMG-box (SOX17) and Apoptosis-associated speck-like protein containing a CARD (ASC)/target of methylation-induced silencing (TMS1), were differentially methylated based on the staging of the disease [46, 47, 48], whereas MLH1, p16, Death-associated protein kinase 1 (DAP-kinase), Ras association domain family member 2 (RASSF2A), and WNT inhibitory factor 1 (Wif-1) were regarded as plasma or serum detection markers for CRC [49].

In summary, these data strongly support the promising utility of DNA methylation as a critical diagnostic marker for CRC. Unfortunately, this has not necessarily translated into their prognostic or predictive use in clinical practice [50]. This can be attributed to significant variability in sensitivity, specificity, and reproducibility between diverse patient cohorts and gene expression platforms, which ultimately impacts the prognostic value of many tests. Currently, two FDA-approved commercial tests, Epi proColon® and Cologuard, have been used for screening alterations in methylation of common genes, such as SEPT9, NDRG4, and BMP3, for early detection of CRC [51]. However, they also generally lack prognostic value and require improvements in terms of sensitivity and specificity. Several other methylation biomarker assays have also been suggested, but validation in independent and large population screening studies is still needed [52].

2.3 Noncoding RNAs

Another epigenetic regulatory mechanism frequently deregulated in CRC involves the role of noncoding RNA (ncRNAs). Specifically, aberrations of microRNAs (miRNAs) expression, a major class of ncRNAs, are often observed in CRC and are considered to play a major role in tumorigenesis and CRC progression [23, 53]. These observations are consistent with the fact that miRNAs tend to exert oncogenic or tumor-suppressive effects. For example, miRNAs, such as miR-141, miR-200c, miR-145, miR-373, miR-520c, miR-135a, and miR-135b, have all been shown to affect CRC by regulating epithelial differentiation, WNT signaling, and CRC cell migratory and invasive potential [54]. Other miRNAs implicated in CRC include miRNA-124a and miRNA-34b/34c, which were shown to regulate the cell cycle and TP53 pathway, respectively [55]. Several miRNAs are also associated with epithelial-to-mesenchymal transition (EMT) in CRC. miR-15/16, miR-140, and miR-200 family members were shown to be associated with suppression of EMT and tumor cell metastatic potential while miR-21 enhanced this process [56].

Other studies have highlighted that alterations in the expression pattern of miRNAs in CRC were considered diagnostic, prognostic, or chemosensitivity markers [57]. For instance, high levels of miR-320 and miR-498 were correlated with progression-free survival in stage II CRC [58], while miR-21 abundance was associated with poor patient response to 5-fluorouracil adjuvant chemotherapy [59]. On the other hand, induced suppression of miR-21 promoted the sensitization of CRC cells to chemotherapeutic regimens [60, 61]. Another study by Toiyama et al. demonstrated a correlation between elevated serum miR-200c levels and stage IV CRC compared to earlier stages. Furthermore, high serum miR-200c showed a significantly positive correlation with lymph node metastasis, distant metastasis, and prognosis [62]. A comprehensive look of miRNAs as CRC biomarkers is reviewed by several other sources [60].

There is also ample evidence of miRNAs being downregulated in CRC and thus playing tumor-suppressive roles [63]. Arndt et al. showed that reduced levels of miR-133a as well as enrichment of miR-224 were associated with CRC initiation [64]. Moreover, this study and others also revealed that CRC patients at the adenomatous and carcinoma stages consistently exhibited reduced steady-state levels of miR-143 and miR-145 [63, 64]. Another classic example includes the miR-34 family. Transfection of miR-34a into CRC cells led to induction of apoptosis and inhibition of cell proliferation in part by amplifying the p53-mediated apoptotic response [65, 66]. Intriguingly, p53 has been shown to regulate miR-34a, suggesting a positive feedback loop between the two in which miR-34a could partly mediate the tumor-suppressive roles of p53 [55].

Interestingly, miRNA deregulation can also induce aberrant activity of many of the components of the epigenetic machinery [67]. Take DNMT3A for example, which has been identified as a miR-143 target and is associated with CRC via downregulation of miR-143 and subsequent increase in DNMT3A expression levels [68]. Other examples include miR-140 and miR-449, which have been shown to target and downregulate HDAC1 and 4, respectively, thus exerting their tumor-suppressive effects [69, 70]. Taken together, these findings underscore the importance of miRNAs in exerting both oncogenic and antitumor roles in CRC, which may serve as the basis for the development of novel prognostic and therapeutic markers.

Advertisement

3. Classification of CRC pathways and subtypes using epigenetic features

The development of CRC occurs via aberrations in multiple genetic and epigenetic pathways. These pathways can be defined by three principal molecular phenotypes (Figure 1). As aforementioned in Section 1, these include the MSI phenotype, which is characterized by mutations in DNA MMR genes; the CIN phenotype characterized by mutations in APC/Wnt/β-catenin pathway; and a third CIMP, defined by global CpG island hypermethylation, which results in widespread silencing of tumor suppressor genes [9, 71]. Notably, each pathway is characterized by distinct epigenetically related pathological features that drive the process of tumor initiation and development. In this section, we will describe various epigenetic aspects of these CRC phenotypes as well as how the molecular aspects of each pathway have been employed as useful diagnostic and prognostic tools to guide the clinical management of CRC patients. Finally, we also briefly acknowledge how these pathways may overlap within broader systems of subtype classification and highlight some of the current challenges in precisely defining CRC subtypes.

Figure 1.

Schematic depicting epigenetic alterations in CRC and their association with CRC molecular subtypes. The CRC “epigenome” harbors alterations in their histone modification states, particularly regarding aberrant histone methylation (blue dot) and acetylation (yellow dot). These global histone aberrations serve as biomarkers to predict the clinical outcome in CRC patients. DNA methylation is another major epigenetic alteration seen in CRC. Usually, global DNA hypomethylation occurs gradually and early in the process of CRC carcinogenesis. This occurs in concert with systematic and discrete CpG island hypermethylation events at the promoters of genes involved in DNA repair, apoptosis, proliferation, angiogenesis, adhesion, and invasion. This generally results in transcriptional silencing. Aberrations of micro-RNAs (miRNAs) expression, a major class of ncRNAs, are also often observed in CRC and are considered to play a major role in tumorigenesis and CRC progression. This figure also illustrates three main epigenetic-related molecular subtypes of CRC, namely the microsatellite instability (MSI) phenotype, the chromosomal instability (CIN) phenotype, and the CpG island methylator phenotype (CIMP). MSI CRCs arise from a defective DNA mismatch repair (MMR) system, which is associated with frameshift mutations and base pair substitutions in genes. CIMP CRCs are characterized by widespread promoter CpG island hypermethylation, whereas the CIN CRCs arise through widespread imbalances in chromosome number (aneuploidy), global hypomethylation, and loss of heterozygosity (LOH).

3.1 Microsatellite instability phenotype (MSI)

Approximately 15–20% of all CRC cases harbor MSI [72]. Microsatellites are defined as repetitive one to six base pair DNA sequences distributed along coding and noncoding regions of the genome. Importantly, the repetitive nature of these regions makes them particularly susceptible to mismatch errors [72]. Tumors with the MSI phenotype are therefore driven by the inactivation of MMR genes, which are involved in repairing DNA recombination and replication errors as well cellular responses to DNA damage [12]. The net effect of defective MMR machinery is accumulation of single base-pair mismatches, which results in a hypermutable cellular state [72].

MSI tumors are typically classified as MSI-high (MSI-H) or MSI-low (MSI-L). In MSI-H CRCs, usually two or more of the five microsatellite markers show instability, whereas in MSI-L tumors, only one of the five markers shows instability [12]. If none of the markers show instability, however, these are classified as microsatellite stable (MSS) CRC tumors, which account for approximately 80–85% of CRC patients [12]. Although majority of MSI-H tumors sporadically arise, a few are also linked to a familial hereditary syndrome known as Lynch syndrome or hereditary non-polyposis CRC (HNPCC) which account for about 3–5% of all CRC cases [12]. Sporadic MSI tumors are generally affected by epigenetic inactivation at the MLH1 promoter via CpG island hypermethylation, whereas Lynch syndrome is caused by germline mutations in MLH1, PMS2, MSH6, or MSH2 [12].

Importantly, unlike MSS, MSI tumors are poorly differentiated and are more often located in the proximal colon. MSI tumors also harbor a mucinous or signet ring type histology and increased number of tumor-infiltrating lymphocytes [72]. Additionally, when MSI was first identified in CRC patients, it was shown that MSI-H patients tended to have a better patient prognosis compared with MSS tumors and had an overall lower tumor stage at diagnosis [12]. Moreover, randomized phase III clinical trials along with several prospective studies have shown that the MMR status of these patients is also predictive of response to adjuvant 5-FU-based chemotherapy [12, 73]. The consensus was that patients with MSI-H tumors did not benefit from adjuvant 5-FU therapy compared to their non-MSI-H counterparts [73]. Furthermore, these data are consistent with other studies that revealed that human CRC cell lines with MSI-H phenotypes displayed resistance to DNA damaging agents, such as 5-FU, which could be overcome by the restoration of normal MMR function [74, 75].

3.2 Chromosomal instability (CIN) phenotype

The acquisition of genomic or chromosomal instability is a key feature in CRC development [39]. In fact, CIN has been found in approximately 85% of CRC cases and is characterized by increased chromosomal losses and gains as well as increased loss of heterozygosity [13, 14]. Although the exact mechanisms underlying CIN remain incompletely understood, it has been attributed to defects in genes related to the DNA damage response, telomere stability, and chromosomal segregation [39]. Unfortunately, standardizing the precise quantitative criteria that define a “CIN-positive” tumor has been challenging due to difficulties in the detection approaches of chromosomal instability [39]. The approaches currently in use include cytometry, karyotyping, and loss of heterozygosity analyses [76].

Along with the typical chromosomal abnormalities, accumulation of a characteristic set of mutations in specific tumor suppressor genes and oncogenes is also a prominent feature observed in CIN tumors [39]. These include mutations in APC, p53, Cyclooxygenase-2 (COX-2), and KRAS as well as 18q alterations [39]. Interestingly, many studies have sought to determine the prognostic value of KRAS, TP53, or 18q alterations. So far, evidence of increased risk of relapse or mortality in CRC patients with KRAS mutations has been presented, but other studies have failed to confirm this correlation [77]. Consequently, some of these putative individual prognostic markers are still undergoing rigorous study. However, several compelling studies indicate that the overall CIN phenotype is associated with a less favorable outcome in patients than those with the MSI phenotype, and unlike MSI tumors, it is not significantly influenced by adjuvant therapy in patients with stage II–III CRC [78].

Several ongoing phase I and II clinical trials are underway to therapeutically target pathways that directly or indirectly initiate and perpetuate CIN. Some of these include small-molecule inhibitors of COX-2, Polo-Like Kinases (Plks), Eg5, and Centromere protein E (CENP-E) [39]. Swanton et al. also showed that CIN-positive tumors are intrinsically resistant to taxanes due to the similarity between pathways that regulate the chromosomal segregation and those implicated in the taxane response [79]. These and other studies have collectively prompted the Chromosomal Instability and Anti-Tubulin Response Assessment (CINATRA) trial to assess whether patients with MSI-positive solid tumors derive benefit from EPO906 (new microtubule stabilizer) compared to patients with CIN-positive cancers [80]. Overall, phase I trials showed encouraging tumor control and response rates in patients with metastatic CRC (mCRC), although the trial was prematurely closed due to toxicity issues [80]. In summary, these data support the role of the CIN pathway in guiding patient stratification and the clinical management of CRC. However, more studies to better define the mechanisms underlying CIN and determine how CIN influences progression will be critical to advance our understanding of the most common form of genetic instability in CRC. Moreover, the feasibility of standardizing detection of CIN-positive tumors and thus be able to target chromosomally unstable cells, will be critical.

3.3 CpG island methylator phenotype (CIMP)

The term “CpG island methylator phenotype,” or CIMP, was coined in 1999 by Toyota and Issa to denote the CRC tumor subtypes characterized by widespread promoter DNA hypermethylation at certain tumor suppressor genes [81]. More than 50% of genes have promoters found within CpG islands [44]. Hence, the frequency of CIMP CRCs depends on which promoters are examined for methylation, with some promoters being more beneficial than others for identifying CIMP. Several studies have revealed that this methylation is common at the promoters of a diverse spectrum of genes, including Phosphatase and tensin homolog (PTEN), RUNX3, and Unc-5 netrin receptor C (UNC5C), making these key genes part of the expression signature profile in the evolution of CIMP CRCs [44]. Moreover, based on a panel of CIMP-specific markers coupled with the B-Raf proto-oncogene (BRAF) mutational status, CIMP tumors may be further classified according to the fraction of promoters that exceed a certain threshold of DNA methylation as being CIMP high, low, or negative [44]. Although CIMP-high and CIMP-low CRCs are significantly associated with biological sex as well as BRAF and KRAS mutational status, these classifications need additional refinement [44]. Nonetheless, it is reminiscent of the classification of CRCs based on degrees of MSI [82].

Notably, several of the clinicopathological characteristics of CIMP-high tumors have also been correlated to MSI tumors [82, 83]. For example, like MSI, CIMP tumors also represent a clinically distinct group characterized by epigenetic instability, distinct histological and pathological features, and discrete precursor lesions [84]. Pathologically, CIMP tumors also originate similarly to MSI tumors in the proximal colon, with a mucinous and poorly differentiated histological type and are frequently diagnosed in elderly and female patients [84, 85]. However, the determination of which specific methylated loci should be used to define CIMP remains a major challenge in the evaluation of CIMP tumors. Currently, several studies have proposed the classic panel containing the genes hMLH1 and p16 as well as Munc-18-interacting (MINT) proteins MINT1, MINT2, and MINT31. This panel has been further developed to contain the genes Calcium voltage-gated channel subunit alpha1 G (CACNA1G), Cellular retinoic acid binding protein 1 (CRABP1), IGF2, Neurogenin 1 (NEUROG1), RUNX3, Suppressor of cytokine signaling 1 (SOCS1), Hypermethylated in cancer 1 (HIC1), IGF-binding protein 3 (IGFBP3), and Werner syndrome ATP-dependent helicase (WRN) [86].

In recent years, the use of DNA hypermethylation of specific genes to predict CRC patient outcome and therapeutic approach has received much attention. Although further validation is warranted, many studies have found a correlation between MSS CIMP+ CRC patients and poor prognosis [87]. Furthermore, the correlation between CIMP status and response to chemotherapy has been investigated. CIMP status predicts poor survival in metastatic MSS CRC patients treated with chemotherapeutic agent 5-FU [88]. Overall, patients who did not receive chemotherapy treatment but had tumors identified as MSS and CIMP had a worse survival outcome [89]. In contrast, two separate studies reported better outcomes for patients with CIMP tumors, a conflict that could be attributed to differences in the criteria used across the studies to define CIMP status. Additionally, this also suggests that the heterogeneous nature of CIMP tumors may warrant further classification [44].

Intriguingly, many studies have also found an association between CIMP status and other important epidemiological factors [90]. For instance, reports of an association between cigarette smoking, obesity, and CIMP showed that the number of cigarettes smoked as well as body mass index (BMI) had a significant relationship to CIMP tumor development [90, 91]. Furthermore, associations of CIMP status with smoking and obesity were evident only for females [90]. Taken together, the above evidences support a critical role of the CIMP pathway in the pathogenesis of CRC, which has also become a significant part of the current management of CRC. In the future, it will also be essential to have a consensus on a standardized panel of loci to define CIMP, similar to that utilized to identify MSI CRCs.

3.4 Consensus molecular subtypes

The three molecular pathways described so far also fall within several consensus classification systems for CRC. These systems vary in terms of the number of proposed subtypes, which can range from three to six depending on the combinations of genetic, epigenetic, clinical, and histopathological parameters used as well as the extent of the overlap between the three molecular pathways. For instance, the Consensus Molecular Subtypes (CMS) consortium has been suggested as one of the most robust classification systems and describes four groups (CMS1–4) based on expression profiling data from multiple studies [92]. While the majority of MSI-H CRCs fall into the CMS1 category, CMS2–4 display higher CIN. However, CMS3 samples have a distinctive profile compared with other CIN tumors. They tend to have lower CIN, higher prevalence of CIMP and close to 30% of the tumors are hypermutated which confers significant overlap with MSI status tumors [93]. Additionally, the Cancer Genome Atlas study also demonstrated that CIMP overlaps with the MSI pathway because of the fact that sporadic MSI-H CRCs usually harbor CIMP-high clinicopathological features [94]. Meanwhile, CMS4 are defined as CIN-heterogeneous tumors with mesenchymal characteristics that occur in later disease stages [93].

It is also noteworthy that many CRC tumors demonstrate mixed characteristics compatible with two or more of these subtypes, which may represent a transition phenotype or intratumoral heterogeneity, while others cannot be precisely classified into any of these pre-defined subgroups [95]. Furthermore, these classifications often lack incorporation of the molecular markers used for traditional TNM staging of CRC [96]. Taken together, these challenges as well as the existing incongruity between the various systems illustrate the need to further refine these consensus classifications by developing more progressive and integrated approaches.

Advertisement

4. The role of major epigenetic enzymes in CRC and therapeutic strategies for targeting them

4.1 Histone methyltransferases and demethylases

As discussed so far, aberrant changes in epigenetic modifications can significantly contribute to CRC progression. It is therefore unsurprising that many of the epigenetic enzymes mediating these modifications are themselves deregulated during the initiation and progression of CRC. Here, we describe the significance of changes in the expression levels of two such families of enzymes that oppose each other in terms of function, namely histone methyltransferases (HMTs) and demethylases (HDMs). Although changes in the expression or activity levels of several methylation-related enzymes have been linked to CRC, in most cases only a limited knowledge regarding the molecular mechanisms by which these enzymes contribute to disease development exists [15]. We summarize current knowledge regarding some of the preclinical validated implications of these enzymes as proof of principle for the employment of epigenetic agents in CRC. We also briefly discuss potential mechanisms of action of these enzymes as well as the advantages of targeting them using combinatorial over monotherapy approaches.

Histone lysine methyltransferases (HKMTs) have been widely studied across multiple solid tumor types including CRC [97]. For instance, studies in a preclinical model of CRC found that increased expression and activity of SET and MYND domain containing 3 (SMYD3), a well-known HKMT, was strongly correlated with tumorigenesis. Moreover, RNAi-mediated depletion of SMYD3 significantly impaired CRC cell proliferation, indicating a crucial role of SMYD3 in maintaining CRC malignancy [98]. More recent studies suggest a putative mechanism by which this overexpression might occur by demonstrating that hypomethylation of the SMYD3 promoter was observed in CRC tumor tissues compared to adjacent normal tissues. Further subgroup clinicopathological analyses showed that this hypomethylation was observed with stage III and IV tumors as defined by moderate to well-differentiated histology and positive lymph node metastasis [99].

Another well-studied HKMT, enhancer of zeste 2 (EZH2), is also frequently deregulated in CRC. Both mRNA and protein levels of EZH2 were found to be significantly increased in CRC tissues compared to non-cancerous counterparts [16]. Additionally, increased EZH2 expression was directly correlated with tumor size, metastases, and overall worse disease-free survival of CRC patients [100]. He et al. also showed that siRNA-mediated depletion of EZH2 inhibited the proliferation and migration of SW620 CRC cells, while inducing apoptosis and G0/G1 cell cycle arrest [101]. Another mechanistic study also revealed that knockdown of EZH2 significantly reduced CRC cell invasion and matrix metalloproteinases 2/9 (MMP2/9) secretion in vitro while promoting increased overall survival and decreased lung metastasis in vivo [102]. Furthermore, this EZH2-induced CRC cell invasion was mediated by direct binding of Signal transducer and activator of transcription 3(STAT3) to the EZH2 promoter, resulting in downregulation of the vitamin D receptor (VDR) [102]. Interestingly, an association between a missense variant in EZH2 and risk of CRC was discovered by the Li group. They identified that the presence of the rs2302427 variant showed a significant association with increased CRC susceptibility [103]. Recent studies point to other mechanistic roles of HKMTs in CRC. For example, depletion of SETD1A, a member of the trithorax (TrxG) family of HMTs, inhibited CRC cell growth and colony formation in part by decreasing expression of approximately 50% of Wnt/β-catenin target genes [28]. Finally, in a mouse model, IL-22-mediated activation of disruptor of telomeric silencing 1-like (DOT1L) promoted CRC stemness and tumorigenic potential and was considered a predictor of poor survival outcome in CRC patients [32].

Protein arginine methyltransferases (PRMTs), although studied to a lesser extent, have also been shown to play critical roles in CRC malignancy via activation of Wnt/β-catenin and NF-ĸB signaling [104]. CARM1, for example, is an important positive modulator of Wnt/β-catenin transcription and was found to promote survival and anchorage-independent growth of CRC cells with aberrantly activated Wnt/β-catenin signaling [105]. Meanwhile, our lab and others have shown that protein arginine methyltransferase 5 (PRMT5) was overexpressed in CRC cells and patient-derived primary tumors, which correlated with increased cell growth, migration, invasion, and NF-ĸB activation as well decreased overall patient survival [106, 107, 108, 109]. The enzymes catalyzing removal of methylation marks, HDMs, are perhaps the least studied among the enzymes mentioned thus far and only a few have been implicated as playing tumor suppressive or oncogenic roles in CRC. LSD1, KDM4B, KDM4C, and KDM5B have all been shown to play pro-tumorigenic roles by promoting CRC cell growth and metastasis, whereas HDMs, such as JMJD3 and JMJD1B, have been implicated as tumor suppressors [15]. Taken together, these data provide strong support for the continued development of selective and potent small-molecule inhibitors against these methylation-modifying enzymes as promising therapeutic agents for CRC.

4.2 Targeting HMTs and HDMs in CRC

Disruption of epigenetic regulation in CRC mediated by deregulated HMTs, DNMTs, and HDMs has garnered increasing interest in recent years. In this section, we aim to review the current status on the development of therapeutic strategies to modulate histone methylation for CRC treatment. The current therapeutics including pre-clinical and clinical agents that target epigenetic enzymes in CRC are listed inTable 1. Thus far, more than 20 histone-methylation enzymes have been found to be clinically relevant to CRC, including 17 oncoproteins and 8 tumor suppressors, although their exact mechanisms of action are not fully understood [15]. Furthermore, more than 20 small-molecule inhibitors targeting HMTs, DMNTs, and HDMs have been employed for preclinical or clinical studies. For example, treatment of DLD1 colon cell line and primary CRC cells with a potent HKMT inhibitor EPZ004777 (anti-DOT1L) resulted in significant reduction in sphere formation in vitro, thus inhibiting cell growth [32]. Other HKMT inhibitors, such as BCI-121 and Chaetocin, have significantly suppressed CRC cell growth and migration by inhibiting SMYD3 and SUV39H1, respectively [98]. Notably, inhibitors against the HKMT EZH2 have yielded some of the most promising results for treating CRC. DZNep, an indirect EZH2 inhibitor, induced apoptosis in CRC cell lines and stem cells, while GSK346 impaired the migratory potential of CRC cells and reduced H3K27me3 levels in Colo205 and HT-29 cells (Table 1) [110].

Table 1.

Overview of pre-clinical and clinical drugs that target epigenetic enzymes in CRC.

Unlike HKMTs, the development of inhibitors against PRMTs has only recently gained prominence in the cancer field, and only a couple of these have made it to the clinical trial phase thus far. AMI-1, which inhibits PRMT1 and PRMT5, demonstrated antiproliferative activity in CRC cells and xenograft mouse models [106]. However, further in vivo validation studies are needed, and it has not entered clinical trial yet. Another promising PRMT5 inhibitor that recently made it to Phase I clinical trials is GSK3326595, which potently inhibited tumor growth in vitro and in vivo [111]. Trials with GSK3326595 are currently being conducted in adult subjects with relapsed and/or refractory solid tumors (NCT02783300). Additionally, inhibitors targeting HDMs are even fewer in number and have shown limited efficacy in suppressing CRC cell growth. For example, KDM4A/C inhibitors were ineffective in blocking HCT116 CRC cell growth when used in isolation [112]. However, they exhibited potent antiproliferative effects in combination with another HDM inhibitor, NCL-2, which targets LSD1 [113]. These data suggest a potential for synergy between the two classes of HDM inhibitors.

Finally, the use of DNMT inhibitors for CRC treatment has also shown some exciting promise. In studies using CRC cell lines, suppression of DNMT1 and DNMT3B resulted in significant reduction in methylation, which correlated with the re-expression of tumor suppressor genes. This also resulted in induction of apoptosis as well as reduced cell proliferation and stemness [114]. Notably, studies with the DNMT1 inhibitor, 5-aza-2′-deoxycytidine (decitabine), exhibited its ability to re-sensitize colorectal tumors to both irinotecan and 5-FU, thus becoming a major component of the treatment regimen for CRC in the clinic [19]. Another recent preclinical study showed that combination of the anti-EGFR inhibitor, gefitinib and decitabine showed highly synergistic inhibition of CRC cell proliferation and migration [115]. Additional combination regimens are outlined in Table 1.

4.3 Acetyltransferases and deacetylases

Acetylation of histones by acetyltransferases (HATs) and removal of these acetyl marks by HDACs are essential events for the maintenance of normal chromatin organization and function [116]. However, as is often the case in cancer, these enzymes are dysregulated, leading to increased chromosomal instability and aberrant gene expression changes [117]. To date, only a handful of HATs have been reported as contributing to the pathogenesis of CRC. Here, we describe the role of a few of these HATs namely p300/CREB-binding protein (p300/CBP), GCN5, N-Acetyltransferase 10 (Nat10), and Human males absent on the first (hMOF). Assessment of 262 CRC samples from patients receiving 5-FU treatment demonstrated that low expression of p300/CBP in CRC tissue was closely associated with poor clinical response to 5-FU based-chemotherapy [118]. Furthermore, low p300/CBP expression also correlated with poor disease-free survival and increased early disease progression in the same patients [118]. Mechanistic studies also uncovered that 5-FU induced degradation of p300/CBP which was dependent on chaperone-mediated autophagy involving heat-shock cognate protein 70 kDa (hsc70) and lysosomal-associated membrane protein 2A (LAMP2A). In short, degradation of p300/CBP was found to be relevant to chemoresistance to 5-FU, since blocking this degradation also enhanced 5-FU’s cytotoxicity in CRC cells [118].

Conversely, another HAT GCN5 has been implicated in promoting CRC cell growth via its upregulation rather than downregulation. One study found that GCN5 overexpression in human colon adenocarcinoma tissues was attributed to the activities of the transcription factors, c-Myc and E2F transcription factor 1 (E2F1) [119]. Depletion of c-Myc inhibited CRC cell proliferation mainly by downregulating GCN5 transcription, an effect that was rescued by ectopic expression of GCN5. However, ectopic overexpression of E2F1 had the opposite effect by suppressing GCN5 levels, thus inducing cell death. Furthermore, inhibition of GCN5 with CPTH2, a HAT inhibitor, also suppressed CRC cell growth, revealing an avenue of great therapeutic potential [119]. Other HATs implicated in CRC include Nat10 and hMOF, which were downregulated in CRC tissues. Particularly, recent studies showed that Nat10 downregulation and subcellular redistribution were associated with increased cellular motility and invasion in CRC cells [120]. Meanwhile, low expression of hMOF correlated with clinicopathological features of CRC such as lymph node metastasis and advanced tumor stage [121].

In CRC, HDACs are also frequently overexpressed and represent another attractive class of targets for anticancer therapy. HDAC1–3 and HDAC5–8 have emerged as some of the most relevant deacetylases in CRC. Although all are highly overexpressed in CRC, only few studies have explored the relevance of this overexpression to disease [23]. For example, knockdown of HDAC1, 2, and 3 reduced the growth of several CRC cells by largely unknown mechanisms [122]. Interestingly however, a mechanistic link between HDAC2 expression and sensitivity of CRC cells to other anticancer agents was recently established. Alzoubi et al. demonstrated that depletion of HDAC2 specifically enhanced the combined anti-tumor effect of the pan-HDAC inhibitor suberoylanilide hydroxamic acid (SAHA) and the DNA-damaging agents, 5-FU and oxaliplatin, in SW480 and HT29 cells. On the other hand, overexpression of HDAC2 conferred resistance to these agents, which were independent of the p53 mutational status [123]. In summary, these findings strongly suggest that HATs and HDACs are critical biomarkers for CRC and influence the sensitivity of CRC cells to certain therapeutics as evidenced by their frequent combination with other anticancer agents (Table 1).

4.4 Targeting HATs and HDACs in CRC

Like methylation, several studies have demonstrated that inhibitors targeting HATs and HDACs also induce epigenetic alterations that modulate the expression of genes or pathways critical for CRC treatment. One study showed that direct inhibition of p300/KAT3B histone acetyltransferase, a coactivator of β-Catenin with rimonabant, induced downregulation of Wnt/β-catenin target genes in HCT116 cells [124]. Furthermore, rimonabant also significantly reduced tumor growth in HCT116 xenografts [124]. The general HAT inhibitors such as garcinol and anacardic acid have also been shown to inhibit CRC tumor growth in mice as well as sensitize cancer cells to irradiation [125].

Compared to HATs, a far greater number of studies have been dedicated to investigating the efficacy of HDAC inhibitors at both the preclinical and clinical level. Overall, the use of HDAC inhibitors in preclinical solid tumor models has shown some early promise albeit their progress to the clinic has been hindered by serious limitations including ineffective concentrations and cardiac toxicity [126]. In CRC specifically, these inhibitors are mainly being administered as combination therapy with conventional chemotherapy or other agents [127]. In pre-clinical models for instance, treatment of irinotecan-resistant CRC cells with HDAC inhibitors conferred sensitization of these cells to irinotecan, whereas HDAC inhibitor CG2 showed an additive effect when used with irinotecan, 5-FU, or oxaliplatin in HCT116 xenografts [128]. Meanwhile, a small molecular inhibitor of HDAC6, ACY-1215, was able to enhance the anticancer activity of oxaliplatin by promoting apoptosis and blocking cell proliferation in CRC cells and xenograft models [129].

4.5 Benefits and pitfalls of epigenetic enzyme inhibitors

Despite their potential, a large gap still remains between the biological activity of epigenetic enzyme inhibitors in preclinical studies and their potential clinical utility. For example, the development of HAT inhibitors poses several challenges [130]. Because of their function in complexes consisting of many proteins which play multiple roles in HAT target specificity, this significantly limits inhibitor use in vivo [130]. Additionally, many undesired effects such as reactivity or lack of selectivity between different HAT subtypes are often associated with HAT inhibitors [130]. Hence, studies geared towards the development of more potent and selective inhibitors by carefully studying the catalytic mechanism and enzyme kinetics of various HATs are needed. As far as HDAC inhibitors are concerned, they have shown preferential efficacy against hematological malignancies, and therefore, drugs such as vorinostat (SAHA) and romidepsin (FK228) have achieved FDA approval for the treatment of cutaneous T cell lymphoma [131, 132]. Unfortunately, the evidence regarding HDAC inhibitors efficacy for solid tumors has not been as convincing and encouraging although they are well tolerated at low but not high doses. Currently, adverse side effects and inadequate clinical efficacy are the major limitations to their use, and more efforts are underway to generate specific HDAC inhibitors for solid tumors such as CRC [133]. Nonetheless, a few early phase clinical trials using vorinostat in combination with other chemotherapeutic agents have shown some early promise for mCRC patients. These include combinatorial regimens of vorinostat with 5-FU, leucovorin, and oxaliplatin (FOLFOX) as well as randomized phase II trial studies investigating the efficacy of vorinostat and hydroxychloroquine or regorafenib in refractory mCRC patients [134, 135]. Other regimens are outlined in Table 1. Similarly, while DNMT inhibitors have also met with some degree of success for treating blood cancers such as myelodysplastic syndrome (e.g., decitabine and 5-azacitidine), the major drawbacks of these compounds in solid tumors are harsh side effects and transient demethylation, which revert after drug removal [136, 137]. Interestingly, however, some studies have suggested that this transient demethylation that occurs with DNMT inhibition (e.g., 5-azadeoxycytidine) potentially creates a therapeutic window that can be leveraged for epigenetic reprogramming and/or combinatorial therapies with cytotoxic drugs [138].

Other general limitations regarding the use of epigenetic therapy in solid tumors deal with the unfavorable pharmacokinetic properties of these drugs, including instability, toxicity, and short half-life [137]. Some of these invariably contribute in some way to common toxicities associated with HDAC and DNMT inhibitors in CRC including thrombocytopenia, neutropenia, diarrhea, nausea, vomiting, and fatigue [139]. Furthermore, maintaining therapeutically relevant levels of the drugs necessary for clinical benefit is particularly difficult, and as of yet, no FDA-approved epigenetic treatments exist for CRC despite promising preclinical studies. This signifies the overall marginal clinically compelling responses to these agents in CRC patients. To overcome some of these limitations, newer formulations have been made to render these inhibitors more bioavailable, stable, and ultimately usable at lower doses with less toxicity and greater therapeutic efficacy. Examples of these include the oral HDAC inhibitor, entinostat, used in in vitro and in vivo models of CRC and an orally active formulation of 5-azacitidine, cc-486 [140].

4.6 Emerging immunomodulatory and epigenetic combinatorial therapies

Compared to MSS tumors, there are an exponentially higher number of mutations acquired in MSI-H CRCs. Interestingly, these mutations have the potential to elevate the production of neo-antigens [141]. The result is increased tumor immunogenicity, which is further complemented by the fact that these tumors also harbor a high number of tumor-infiltrating lymphocytes. Within this context, CRC patients with MSI-H represent a subgroup more likely to benefit from immune checkpoint inhibitors compared to those with MSS tumors. Immune checkpoint inhibitors have shown unprecedented benefit across multiple tumor types. These agents specifically target the proteins programmed death ligand-1 (e.g., durvalumab) and programmed death-1 (e.g., nivolumab, pembrolizumab) and are administered as monotherapies or in combination with other anticancer agents. At the present time, several ongoing early and late-phase II and III clinical trials investigating the efficacy of immune checkpoint inhibitors in MSI-H and MSS CRC patients are being extensively explored including pembrolizumab (Keytruda), which recently obtained FDA approval (e.g., NCT01876511 and NCT02060188) [142].

Moreover, the possibility of combining epigenetic therapy and immunotherapy has also been recently explored, and several ongoing clinical trials in CRC investigating the combination strategies of HDACi and DNMTi with checkpoint inhibitors have been undertaken. Specifically, these epigenetic therapies have been shown to augment the effect of checkpoint inhibitors and are currently in early and late phase clinical trials [143]. However, since MSS subtypes represent the larger fraction of CRC cases, the marginal activity displayed by drugs such as pembrolizumab for treating MSS CRCs has been less than encouraging [144]. Hence, overcoming the clinical ineffectiveness of this class of drugs for this subtype remains an important need. Intriguingly, however, recent studies showed that treatment with 5-azacitidine and entinostat in CRC cell lines conferred a shift towards a CIMP+ signature, which would predictively convert them into a more immunogenic state [145]. This increased sensitivity to immunotherapy has prompted a clinical trial evaluating this strategy, with the combination regimen of romidepsin (HDAC inhibitor) and cc-486 with pembrolizumab in MSS-CRC patients (NCT02512172) [145]. Finally, romidepsin was also found to potentiate 5-FU cytotoxicity in HCT-116, HT29, and SW48 cells by inducing apoptosis and cell cycle arrest [146]. Interestingly, MHC class II gene expression was also induced with this combination, once again supporting the possible cooperation of epigenetic therapy with immunomodulatory agents [146]. In summary, the above evidences support a cooperative role between epigenetic and immune therapies, although further efforts to optimize the epigenetic control of immune-related gene expression will be necessary to successfully translate these notions to the clinic.

Advertisement

5. Conclusion and perspectives

In this chapter, we have highlighted the pivotal contribution of epigenetic deregulation, specifically, DNA methylation, histone alterations, and miRNAs to the initiation, progression, and prognosis of CRC. We also underscored the relevance of these epigenetic mechanisms in terms of classifying CRC subtypes as well as their importance in guiding strategies for therapeutic intervention. Moreover, we emphasized the epigenetic enzymes that are involved in these aberrant pathways and presented some up-to-date findings on pre-clinical and clinical trials of epigenetic drugs used as single agents or in combination with conventional anticancer agents in CRC. Furthermore, mounting evidence demonstrates that epigenetic drugs are also capable of altering the immunogenicity of the CRC microenvironment and creating opportunities for potentiating the effects of immune checkpoint inhibitors.

Understandably, drugs targeting the cancer epigenome are also plagued with major challenges including lack of specificity, toxicity, and short half-life. Fortunately, these challenges have facilitated re-evaluation of the dosing and formulation strategies for epigenetic drugs, leading to superior therapeutic drugs with lower toxic profiles. Another underexplored avenue includes targeting less commonly manipulated epigenetic mechanisms such as the use of miRNA mimics [147]. Furthermore, in light of the advent of personalized therapies, more intricate studies are also needed to elucidate the relationship between individual driver genetic mutations and epigenetic alterations, thus providing a pathway-driven basis for developing selective therapeutic strategies. This may call for a more stringent control of gene expression in CRC cells via selective targeting of epigenetic regulatory enzymes. This includes the prospects of CRISPR/Cas9/Cas13-based genome and RNA editing, which may provide validated starting points for further development towards novel CRC therapeutic agents [148].

Advertisement

Acknowledgments

This publication is made possible, in part, with support from the Indiana Clinical and Translational Sciences Institute (CTSI) funded from the National Institutes of Health, National Center for Advancing Translational Sciences, Clinical and Translational Sciences Award (to TL), V foundation Kay Yow Cancer Fund (Grant 4486242 to TL), NIH-NIGMS Grant (#1R01GM120156-01A1 (to TL), and 100 VOH Grant (#2987613 to TL), as well as NIH-NCI Grant (#1R03CA223906-01).

Advertisement

Abbreviations

APAKATM and p53-associated KZNF protein
APCadenomatous polyposis coli
ASC/TMS1 or PYCARDapoptosis-associated speck-like protein containing a CARD
BETbromodomain and extra-terminal motif
BMIbody mass index
BRAFB-Raf proto-oncogene, serine/threonine kinase
CACNA1Gcalcium voltage-gated channel subunit alpha1 G
CARM1coactivator-associated arginine methyltransferase 1
CDKN2A/p16INK4acyclin-dependent kinase inhibitor 2A
CENP-Ecentromere protein E
CIMPCpG island methylator phenotype
CINchromosomal instability
COX-2cyclooxygenase-2
CRABP1cellular retinoic acid binding protein 1
CRCcolorectal cancer
DAP-kinasedeath-associated protein kinase 1
DNMTDNA methyltransferase
DNMT3ADNA methyltransferase 3A
DOT1Ldisruptor of telomeric silencing 1-like
E2F1E2F transcription factor 1
Eg5kinesin 5 family member
EMTepithelial-to-mesenchymal transition
EZH2enhancer of zeste 2
FDAFood and Drug Administration
5-FU5-fluorouracil
GATA4GATA-binding protein 4
HAThistone acetyltransferase
HDAChistone deacetylase
HDMhistone demethylase
HIC1hypermethylated in cancer 1
HKMThistone lysine methyltransferase
HKMTshistone lysine methyltransferases
hMLH1human mutL homolog 1
hMOFhuman males absent on the first
HMThistone methyltransferase
Hnf1bHNF1 homeobox B
HNPCChereditary non-polyposis colon cancer
HS3ST2 (3OST2)heparan sulfate-glucosamine 3-sulfotransferase 2
hsc70heat-shock cognate protein 70 kDa
IGF2insulin-like growth factor 2
IGFBP3insulin-like growth factor binding protein 3
IL-6interleukin 6
iNOSinducible nitric oxide synthase
KDMlysine demethylase
KRASKirsten rat sarcoma 2 viral oncogene
LAMP2Alysosomal-associated membrane protein 2A
LOHloss of heterozygosity
mCRCmetastatic colorectal cancer
MGMTO-6-methylguanine-DNA methyltransferase
MINTMunc-18-interacting
miRNAmicroRNA
MLH1mutL homolog 1
MSH2, 6MutS protein homolog 2, 6
MSImicrosatellite instability
MSSmicrosatellite stable
Nat10N-acetyltransferase 10
NEUROG1neurogenin 1
nRNAnoncoding RNA
p300/CBPp300/CREB-binding protein
p53 or TP53tumor protein 53
PCDH10protocadherin 10
PD-1programmed death-1
PDL-1programmed death ligand-1
Plkspolo-like kinases
PMS2PMS1 homolog 2, mismatch repair system component
PRMTprotein arginine methyltransferase
PRMT5protein arginine methyltransferase 5
PTENphosphatase and tensin homolog
RASSF1Aras association domain family member 1
RASSF2Aras association domain family member 2
RUNX3runt-related transcription factor 3
SAHAsuberoylanilide hydroxamic acid
SETD1ASET domain-containing protein 1A
SFRP1secreted frizzled related protein 1
SFRP5secreted frizzled related protein 5
SMYD3SET and MYND domain containing 3
SOCS1suppressor of cytokine signaling 1
SOX17SRY-Box 17
STAT3signal transducer and activator of transcription 3
SUV39H1suppressor of variegation 3–9 homolog 1
TIMP3tissue inhibitor of metalloproteinases 3
TMEFF2transmembrane protein with EGF-like and two follistatin-like domains 2
TNF-αtumor necrosis factor alpha
UNC5CUnc-5 netrin receptor C
Wif-1WNT inhibitory factor 1
WNTwingless type
WRNWerner syndrome ATP-dependent helicase

References

  1. 1. Arnold M et al. Global patterns and trends in colorectal cancer incidence and mortality. Gut. 2017;66(4):683-691
  2. 2. Vaiopoulos AG, Athanasoula KC, Papavassiliou AG. Epigenetic modifications in colorectal cancer: Molecular insights and therapeutic challenges. Biochimica et Biophysica Acta (BBA)—Molecular Basis of Disease. 2014;1842(7):971-980
  3. 3. Bardhan K, Liu K. Epigenetics and colorectal cancer pathogenesis. Cancers. 2013;5(2):676-713
  4. 4. Sjoblom T et al. The consensus coding sequences of human breast and colorectal cancers. Science. 2006;314(5797):268-274
  5. 5. Migheli F, Migliore L. Epigenetics of colorectal cancer. Clinical Genetics. 2012;81(4):312-318
  6. 6. Lao VV, Grady WM. Epigenetics and colorectal cancer. Nature Reviews. Gastroenterology & Hepatology. 2011;8(12):686-700
  7. 7. van Engeland M et al. Colorectal cancer epigenetics: Complex simplicity. Journal of Clinical Oncology. 2011;29(10):1382-1391
  8. 8. Datta K, Kolaja K. Epigenetics: Druggable targets for anti-cancer therapies. Current Opinion in Toxicology. 2017;6:79-86
  9. 9. Inamura K. Colorectal cancers: An update on their molecular pathology. Cancers. 2018;10(1):26
  10. 10. Gavin PG et al. Colon cancer mutation: Prognosis/prediction—Response. Clinical Cancer Research. 2013;19(5):1301
  11. 11. Kawakami H, Zaanan A, Sinicrope FA. Microsatellite instability testing and its role in the management of colorectal cancer. Current Treatment Options in Oncology. 2015;16(7):30
  12. 12. Kawakami H, Zaanan A, Sinicrope FA. MSI testing and its role in the management of colorectal cancer. Current Treatment Options in Oncology. 2015;16(7):30-30
  13. 13. Comprehensive molecular characterization of human colon and rectal cancer. Nature. Cancer Genome Atlas Network; 2012;487(7407):330-337
  14. 14. Carethers JM, Jung BH. Genetics and genetic biomarkers in sporadic colorectal cancer. Gastroenterology. 2015;149(5):1177-1190.e3
  15. 15. Huang T et al. Targeting histone methylation for colorectal cancer. Therapeutic Advances in Gastroenterology. 2017;10(1):114-131
  16. 16. Liu YL et al. Expression and clinicopathological significance of EED, SUZ12 and EZH2 mRNA in colorectal cancer. Journal of Cancer Research and Clinical Oncology. 2015;141(4):661-669
  17. 17. Stypula-Cyrus Y et al. HDAC up-regulation in early colon field carcinogenesis is involved in cell tumorigenicity through regulation of chromatin structure. PLoS One. 2013;8(5):e64600
  18. 18. Mottamal M et al. Histone deacetylase inhibitors in clinical studies as templates for new anticancer agents. Molecules. 2015;20(3):3898-3941
  19. 19. Flis S, Gnyszka A, Flis K. DNA methyltransferase inhibitors improve the effect of chemotherapeutic agents in SW48 and HT-29 colorectal cancer cells. PLoS One. 2014;9(3):e92305
  20. 20. Baretti M, Azad NS. The role of epigenetic therapies in colorectal cancer. Current Problems in Cancer. 2018;42(6):530-547
  21. 21. Bhome R et al. The colorectal cancer microenvironment: Strategies for studying the role of cancer-associated fibroblasts. Methods in Molecular Biology. 2018;1765:87-98
  22. 22. Dunn J, Rao S. Epigenetics and immunotherapy: The current state of play. Molecular Immunology. 2017;87:227-239
  23. 23. Jia Y, Guo M. Epigenetic changes in colorectal cancer. Chinese Journal of Cancer. 2013;32(1):21-30
  24. 24. Tamagawa H et al. The global histone modification pattern correlates with overall survival in metachronous liver metastasis of colorectal cancer. Oncology Reports. 2012;27(3):637-642
  25. 25. Gezer U et al. Characterization of H3K9me3- and H4K20me3-associated circulating nucleosomal DNA by high-throughput sequencing in colorectal cancer. Tumour Biology. 2013;34(1):329-336
  26. 26. Benard A et al. Histone trimethylation at H3K4, H3K9 and H4K20 correlates with patient survival and tumor recurrence in early-stage colon cancer. BMC Cancer. 2014;14:531
  27. 27. Kodach LL et al. The role of EZH2 and DNA methylation in the silencing of the tumour suppressor RUNX3 in colorectal cancer. Carcinogenesis. 2010;31(9):1567-1575
  28. 28. Salz T et al. hSETD1A regulates Wnt target genes and controls tumor growth of colorectal cancer cells. Cancer Research. 2014;74(3):775-786
  29. 29. Lu Y et al. Silencing of the DNA mismatch repair gene MLH1 induced by hypoxic stress in a pathway dependent on the histone demethylase LSD1. Cell Reports. 2014;8(2):501-513
  30. 30. Yokoyama Y et al. Cancer-associated upregulation of histone H3 lysine 9 trimethylation promotes cell motility in vitro and drives tumor formation in vivo. Cancer Science. 2013;104(7):889-895
  31. 31. Olcina MM et al. H3K9me3 facilitates hypoxia-induced p53-dependent apoptosis through repression of APAK. Oncogene. 2016;35(6):793-799
  32. 32. Kryczek I et al. IL-22(+)CD4(+) T cells promote colorectal cancer stemness via STAT3 transcription factor activation and induction of the methyltransferase DOT1L. Immunity. 2014;40(5):772-784
  33. 33. Shah MA et al. A global assessment of cancer genomic alterations in epigenetic mechanisms. Epigenetics & Chromatin. 2014;7(1):29
  34. 34. Li F et al. The histone mark H3K36me3 regulates human DNA mismatch repair through its interaction with MutSalpha. Cell. 2013;153(3):590-600
  35. 35. Richon VM et al. Histone deacetylase inhibitor selectively induces p21WAF1 expression and gene-associated histone acetylation. Proceedings of the National Academy of Sciences of the United States of America. 2000;97(18):10014-10019
  36. 36. Karczmarski J et al. Histone H3 lysine 27 acetylation is altered in colon cancer. Clinical Proteomics. 2014;11(1):24-24
  37. 37. Guo Y et al. The epigenetic effects of aspirin: The modification of histone H3 lysine 27 acetylation in the prevention of colon carcinogenesis in azoxymethane- and dextran sulfate sodium-treated CF-1 mice. Carcinogenesis. 2016;37(6):616-624
  38. 38. Suzuki K et al. Global DNA demethylation in gastrointestinal cancer is age dependent and precedes genomic damage. Cancer Cell. 2006;9(3):199-207
  39. 39. Pino MS, Chung DC. The chromosomal instability pathway IN colon cancer. Gastroenterology. 2010;138(6):2059-2072
  40. 40. Wu C, Bekaii-Saab T. CpG Island methylation, microsatellite instability, and BRAF mutations and their clinical application in the treatment of colon cancer. Chemotherapy Research and Practice. 2012;2012:359041
  41. 41. Hinoue T et al. Genome-scale analysis of aberrant DNA methylation in colorectal cancer. Genome Research. 2012;22(2):271-282
  42. 42. Brenet F et al. DNA methylation of the first exon is tightly linked to transcriptional silencing. PLoS One. 2011;6(1):e14524
  43. 43. Ashktorab H et al. Toward a comprehensive and systematic methylome signature in colorectal cancers. Epigenetics. 2013;8(8):807-815
  44. 44. Mojarad EN et al. The CpG island methylator phenotype (CIMP) in colorectal cancer. Gastroenterology and Hepatology From Bed to Bench. 2013;6(3):120-128
  45. 45. Silva TD et al. DNA methylation as an epigenetic biomarker in colorectal cancer. Oncology Letters. 2013;6(6):1687-1692
  46. 46. Galamb O et al. Aberrant DNA methylation of WNT pathway genes in the development and progression of CIMP-negative colorectal cancer. Epigenetics. 2016;11(8):588-602
  47. 47. Riojas MA et al. Methylation-induced silencing of ASC/TMS1, a pro-apoptotic gene, is a late-stage event in colorectal cancer. Cancer Biology & Therapy. 2007;6(11):1710-1716
  48. 48. Zhang W et al. Epigenetic inactivation of the canonical Wnt antagonist SRY-box containing gene 17 in colorectal cancer. Cancer Research. 2008;68(8):2764-2772
  49. 49. Lee BB et al. Aberrant methylation of APC, MGMT, RASSF2A, and Wif-1 genes in plasma as a biomarker for early detection of colorectal cancer. Clinical Cancer Research. 2009;15(19):6185-6191
  50. 50. Draht MXG et al. Prognostic DNA methylation markers for sporadic colorectal cancer: A systematic review. Clinical Epigenetics. 2018;10(1):35
  51. 51. Toth K et al. Detection of methylated SEPT9 in plasma is a reliable screening method for both left- and right-sided colon cancers. PLoS One. 2012;7(9):e46000
  52. 52. Mojtabanezhad Shariatpanahi A et al. The importance of stool DNA methylation in colorectal cancer diagnosis: A meta-analysis. PLoS One. 2018;13(7):e0200735
  53. 53. Orang AV, Barzegari A. MicroRNAs in colorectal cancer: From diagnosis to targeted therapy. Asian Pacific Journal of Cancer Prevention. 2014;15(17):6989-6999
  54. 54. Strubberg AM, Madison BB. MicroRNAs in the etiology of colorectal cancer: Pathways and clinical implications. Disease Models & Mechanisms. 2017;10(3):197-214
  55. 55. Okada N et al. A positive feedback between p53 and miR-34 miRNAs mediates tumor suppression. Genes & Development. 2014;28(5):438-450
  56. 56. Li J et al. MicroRNA-140 inhibits the epithelial-mesenchymal transition and metastasis in colorectal cancer. Molecular Therapy—Nucleic Acids. 2018;10:426-437
  57. 57. Hollis M et al. MicroRNAs potential utility in colon cancer: Early detection, prognosis, and chemosensitivity. World Journal of Gastroenterology. 2015;21(27):8284-8292
  58. 58. Earle JS et al. Association of microRNA expression with microsatellite instability status in colorectal adenocarcinoma. The Journal of Molecular Diagnostics. 2010;12(4):433-440
  59. 59. Schetter AJ et al. MicroRNA expression profiles associated with prognosis and therapeutic outcome in colon adenocarcinoma. Journal of the American Medical Association. 2008;299(4):425-436
  60. 60. Okugawa Y, Toiyama Y, Goel A. An update on microRNAs as colorectal cancer biomarkers: Where are we and what's next? Expert Review of Molecular Diagnostics. 2014;14(8):999-1021
  61. 61. Shibuya H et al. Clinicopathological and prognostic value of microRNA-21 and microRNA-155 in colorectal cancer. Oncology. 2010;79(3-4):313-320
  62. 62. Toiyama Y et al. Serum miR-200c is a novel prognostic and metastasis-predictive biomarker in patients with colorectal cancer. Annals of Surgery. 2014;259(4):735-743
  63. 63. Meng W-J et al. MicroRNA expression profile reveals miR-17-92 and miR-143-145 cluster in synchronous colorectal Cancer. Medicine. 2015;94(32):e1297
  64. 64. Arndt GM et al. Characterization of global microRNA expression reveals oncogenic potential of miR-145 in metastatic colorectal cancer. BMC Cancer. 2009;9:374-374
  65. 65. Zhang Q et al. miR-34a increases the sensitivity of colorectal cancer cells to 5-fluorouracil in vitro and in vivo. American Journal of Cancer Research. 2018;8(2):280-290
  66. 66. Slattery ML et al. The p53-signaling pathway and colorectal cancer: Interactions between downstream p53 target genes and miRNAs. Genomics. 2018. pii: S0888-7543(18)30036-3. DOI: 10.1016/j.ygeno.2018.05.006. [Epub ahead of print]
  67. 67. Wang F et al. Reciprocal regulation between microRNAs and epigenetic machinery in colorectal cancer. Oncology Letters. 2017;13(3):1048-1057
  68. 68. Ng EK et al. MicroRNA-143 targets DNA methyltransferases 3A in colorectal cancer. British Journal of Cancer. 2009;101(4):699-706
  69. 69. Yu L et al. microRNA-140-5p inhibits colorectal cancer invasion and metastasis by targeting ADAMTS5 and IGFBP5. Stem Cell Research & Therapy. 2016;7:180
  70. 70. Sun X et al. miR-449a inhibits colorectal cancer progression by targeting SATB2. Oncotarget. 2017;8(60):100975-100988
  71. 71. Roelands J et al. Immunogenomic classification of colorectal cancer and therapeutic implications. International Journal of Molecular Sciences. 2017;18(10):2229
  72. 72. Gatalica Z et al. High microsatellite instability (MSI-H) colorectal carcinoma: A brief review of predictive biomarkers in the era of personalized medicine. Familial Cancer. 2016;15:405-412
  73. 73. Samowitz WS et al. Microsatellite instability in sporadic colon cancer is associated with an improved prognosis at the population level. Cancer Epidemiology, Biomarkers & Prevention. 2001;10(9):917-923
  74. 74. Boland CR, Goel A. Microsatellite instability in colorectal cancer. Gastroenterology. 2010;138(6):2073-2087.e3
  75. 75. Arnold CN, Goel A, Boland CR. Role of hMLH1 promoter hypermethylation in drug resistance to 5-fluorouracil in colorectal cancer cell lines. International Journal of Cancer. 2003;106(1):66-73
  76. 76. Weber JC et al. Allelotyping analyses of synchronous primary and metastasis CIN colon cancers identified different subtypes. International Journal of Cancer. 2007;120(3):524-532
  77. 77. Wang C et al. Prognostic significance of microsatellite instability and Ki-ras mutation type in stage II colorectal cancer. Oncology. 2003;64(3):259-265
  78. 78. Walther A, Houlston R, Tomlinson I. Association between chromosomal instability and prognosis in colorectal cancer: A meta-analysis. Gut. 2008;57(7):941-950
  79. 79. Swanton C et al. Regulators of mitotic arrest and ceramide metabolism are determinants of sensitivity to paclitaxel and other chemotherapeutic drugs. Cancer Cell. 2007;11(6):498-512
  80. 80. Moorcraft SY et al. Patupilone in patients with pretreated metastatic/locally recurrent colorectal cancer: Results of the phase II CINATRA trial. Investigational New Drugs. 2013;31(5):1339-1344
  81. 81. Toyota M et al. CpG island methylator phenotype in colorectal cancer. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(15):8681-8686
  82. 82. Hawkins N et al. CpG island methylation in sporadic colorectal cancers and its relationship to microsatellite instability. Gastroenterology. 2002;122(5):1376-1387
  83. 83. Shen L et al. Integrated genetic and epigenetic analysis identifies three different subclasses of colon cancer. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(47):18654-18659
  84. 84. Kang GH. Four molecular subtypes of colorectal cancer and their precursor lesions. Archives of Pathology & Laboratory Medicine. 2011;135(6):698-703
  85. 85. Ogino S et al. CpG island methylator phenotype-low (CIMP-low) in colorectal cancer: Possible associations with male sex and KRAS mutations. The Journal of Molecular Diagnostics. 2006;8(5):582-588
  86. 86. Lee S et al. CpG island methylator phenotype in colorectal cancers: Comparison of the new and classic CpG island methylator phenotype marker panels. Archives of Pathology & Laboratory Medicine. 2008;132(10):1657-1665
  87. 87. Ward RL et al. Adverse prognostic effect of methylation in colorectal cancer is reversed by microsatellite instability. Journal of Clinical Oncology. 2003;21(20):3729-3736
  88. 88. Shen L et al. Association between DNA methylation and shortened survival in patients with advanced colorectal cancer treated with 5-fluorouracil based chemotherapy. Clinical Cancer Research. 2007;13(20):6093-6098
  89. 89. Van Rijnsoever M et al. CpG island methylator phenotype is an independent predictor of survival benefit from 5-fluorouracil in stage III colorectal cancer. Clinical Cancer Research. 2003;9(8):2898-2903
  90. 90. Weisenberger DJ et al. Association of the Colorectal CpG Island Methylator phenotype with molecular features, risk factors and family history. Cancer Epidemiology, Biomarkers & Prevention. A publication of the American Association for Cancer Research, cosponsored by the American Society of Preventive Oncology. 2015;24(3):512-519
  91. 91. Samowitz WS et al. Association of smoking, CpG island methylator phenotype, and V600E BRAF mutations in colon cancer. Journal of the National Cancer Institute. 2006;98(23):1731-1738
  92. 92. Müller MF, Ibrahim AEK, Arends MJ. Molecular pathological classification of colorectal cancer. Virchows Archiv: An International Journal of Pathology. 2016;469(2):125-134
  93. 93. Thanki K et al. Consensus molecular subtypes of colorectal cancer and their clinical implications. International Biological and Biomedical Journal. 2017;3(3):105-111
  94. 94. Kim JH et al. Distinct features between MLH1-methylated and unmethylated colorectal carcinomas with the CpG island methylator phenotype: Implications in the serrated neoplasia pathway. Oncotarget. 2016;7(12):14095-14111
  95. 95. Guinney J et al. The consensus molecular subtypes of colorectal cancer. Nature Medicine. 2015;21(11):1350-1356
  96. 96. Alwers E et al. Associations between molecular classifications of colorectal cancer and patient survival: A systematic review. Clinical Gastroenterology and Hepatology. 2018. pii: S1542-3565(17)31536-7. DOI: 10.1016/j.cgh.2017.12.038. [Epub ahead of print]
  97. 97. Albert M, Helin K. Histone methyltransferases in cancer. Seminars in Cell & Developmental Biology. 2010;21(2):209-220
  98. 98. Peserico A et al. A SMYD3 small-molecule inhibitor impairing cancer cell growth. Journal of Cellular Physiology. 2015;230(10):2447-2460
  99. 99. Li B et al. SMYD3 promoter hypomethylation is associated with the risk of colorectal cancer. Future Oncology. 2018;14(18):1825-1834
  100. 100. Wang CG et al. EZH2 and STAT6 expression profiles are correlated with colorectal cancer stage and prognosis. World Journal of Gastroenterology. 2010;16(19):2421-2427
  101. 101. He SB et al. Inhibition of EZH2 expression is associated with the proliferation, apoptosis, and migration of SW620 colorectal cancer cells in vitro. Experimental Biology and Medicine (Maywood, N.J.). 2015;240(4):458-466
  102. 102. Lin Y-W et al. Role of STAT3 and Vitamin D Receptor in EZH2-Mediated Invasion of Human Colorectal Cancer. 2013;230(3):277-290
  103. 103. Li H et al. A missense variant in EZH2 is associated with colorectal cancer risk in a Chinese population. Oncotarget. 2017;8(55):94738-94742
  104. 104. Chung J et al. Protein arginine methyltransferase 5 regulates WNT/β-catenin target gene expression in at multiple levels. Blood. 2016;128(22):4106-4106
  105. 105. Ou CY et al. A coactivator role of CARM1 in the dysregulation of beta-catenin activity in colorectal cancer cell growth and gene expression. Molecular Cancer Research. 2011;9(5):660-670
  106. 106. Zhang B et al. Targeting protein arginine methyltransferase 5 inhibits colorectal cancer growth by decreasing arginine methylation of eIF4E and FGFR3. Oncotarget. 2015;6(26): p. 22799-22811
  107. 107. Prabhu L et al. Adapting AlphaLISA high throughput screen to discover a novel small-molecule inhibitor targeting protein arginine methyltransferase 5 in pancreatic and colorectal cancers. Oncotarget. 2017;8(25):39963-39977
  108. 108. Wei H et al. Protein arginine methylation of non-histone proteins and its role in diseases. Cell Cycle. 2014;13(1):32-41
  109. 109. Wei H et al. PRMT5 dimethylates R30 of the p65 subunit to activate NF-kappaB. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(33):13516-13521
  110. 110. Verma SK et al. Identification of potent, selective, cell-active inhibitors of the histone lysine methyltransferase EZH2. ACS Medicinal Chemistry Letters. 2012;3(12):1091-1096
  111. 111. Wang Y, Hu W, Yuan Y. Protein arginine methyltransferase 5 (PRMT5) as an anticancer target and its inhibitor discovery. Journal of Medicinal Chemistry. 2018;61(21):9429-9441
  112. 112. Kim TD et al. Pro-growth role of the JMJD2C histone demethylase in HCT-116 colon cancer cells and identification of curcuminoids as JMJD2 inhibitors. American Journal of Translational Research. 2014;6(3):236-247
  113. 113. Lin H et al. Small molecule KDM4s inhibitors as anti-cancer agents. Journal of Enzyme Inhibition and Medicinal Chemistry. 2018;33(1):777-793
  114. 114. Li S et al. Inhibition of DNMT suppresses the stemness of colorectal cancer cells through down-regulating Wnt signaling pathway. Cellular Signalling. 2018;47:79-87
  115. 115. Lou YF et al. Combination of gefitinib and DNA methylation inhibitor decitabine exerts synergistic anti-cancer activity in colon cancer cells. PLoS One. 2014;9(5):e97719
  116. 116. Gong F, Miller KM. Mammalian DNA repair: HATs and HDACs make their mark through histone acetylation. Mutation Research. 2013;750(1-2):23-30
  117. 117. Yang XJ, Seto E. HATs and HDACs: From structure, function and regulation to novel strategies for therapy and prevention. Oncogene. 2007;26(37):5310-5318
  118. 118. Du C et al. 5-fluorouracil targets histone acetyltransferases p300/CBP in the treatment of colorectal cancer. Cancer Letters. 2017;400:183-193
  119. 119. Yin YW et al. The histone acetyltransferase GCN5 expression is elevated and regulated by c-Myc and E2F1 transcription factors in human colon cancer. Gene Expression. 2015;16(4):187-196
  120. 120. Zhang H et al. GSK-3beta-regulated N-acetyltransferase 10 is involved in colorectal cancer invasion. Clinical Cancer Research. 2014;20(17):4717-4729
  121. 121. Cao L et al. Correlation of low expression of hMOF with clinicopathological features of colorectal carcinoma, gastric cancer and renal cell carcinoma. International Journal of Oncology. 2014;44(4):1207-1214
  122. 122. Wilson AJ et al. Histone deacetylase 3 (HDAC3) and other class I HDACs regulate colon cell maturation and p21 expression and are deregulated in human colon cancer. The Journal of Biological Chemistry. 2006;281(19):13548-13558
  123. 123. Alzoubi S et al. Synergy between histone deacetylase inhibitors and DNA-damaging agents is mediated by histone deacetylase 2 in colorectal cancer. Oncotarget. 2016;7(28):44505-44521
  124. 124. Proto MC et al. Inhibition of Wnt/beta-catenin pathway and histone acetyltransferase activity by Rimonabant: A therapeutic target for colon cancer. Scientific Reports. 2017;7(1):11678
  125. 125. Ranjbarnejad T et al. Garcinol exhibits anti-proliferative activities by targeting microsomal prostaglandin E synthase-1 in human colon cancer cells. Human & Experimental Toxicology. 2017;36(7):692-700
  126. 126. Tampakis A, Christina Tampaki E, Kouraklis G. Histone deacetylase inhibitors and colorectal cancer: What is new? 2014;14(9):1220-1227
  127. 127. Suraweera A, O’Byrne KJ, Richard DJ. Combination therapy with histone deacetylase inhibitors (HDACi) for the treatment of cancer: Achieving the full therapeutic potential of HDACi. Frontiers in Oncology. 2018;8:92
  128. 128. Na YS et al. Effects of the HDAC inhibitor CG2 in combination with irinotecan, 5-fluorouracil, or oxaliplatin on HCT116 colon cancer cells and xenografts. Oncology Reports. 2010;24(6):1509-1514
  129. 129. Lee DH et al. The HDAC6 inhibitor ACY1215 enhances the anticancer activity of oxaliplatin in colorectal cancer cells. International Journal of Oncology. 2018;53(2):844-854
  130. 130. Wapenaar H, Dekker FJJCE. Histone acetyltransferases: Challenges in targeting bi-substrate enzymes. Clinical Epigenetics. 2016;8(1):59
  131. 131. Jain S, Zain J. Romidepsin in the treatment of cutaneous T-cell lymphoma. Journal of Blood Medicine. 2011;2:37-47
  132. 132. Geskin LJ. Vorinostat in combination with other agents for therapy of cutaneous T-cell lymphomas: A case series. Blood. 2007;110(11):4482
  133. 133. Li Y, Seto E. HDACs and HDAC inhibitors in cancer development and therapy. Cold Spring Harbor Perspectives in Medicine. 2016;6(10):a026831
  134. 134. Van Cutsem E et al. Regorafenib for patients with metastatic colorectal cancer who progressed after standard therapy: Results of the large, single-arm, open-label phase IIIb CONSIGN study. The Oncologist. 2018. https://doi.org/10.1634/theoncologist.2018-0072
  135. 135. Fakih MG et al. A phase I, pharmacokinetic, and pharmacodynamic study of vorinostat in combination with 5-fluorouracil, leucovorin, and oxaliplatin in patients with refractory colorectal Cancer. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 2009;15(9):3189-3195
  136. 136. Hervouet E et al. Specific or not specific recruitment of DNMTs for DNA methylation, an epigenetic dilemma. Clinical Epigenetics. 2018;10(1):17
  137. 137. Heerboth S et al. Use of epigenetic drugs in disease: An overview. Genetics & Epigenetics. 2014;6:9-19
  138. 138. Lyko F, Brown R. DNA methyltransferase inhibitors and the development of epigenetic cancer therapies. Journal of the National Cancer Institute. 2005;97(20):1498-1506
  139. 139. Nervi C, De Marinis E, Codacci-Pisanelli G. Epigenetic treatment of solid tumours: A review of clinical trials. Clinical Epigenetics. 2015;7:127
  140. 140. Azad NS et al. Combination epigenetic therapy in metastatic colorectal cancer (mCRC) with subcutaneous 5-azacitidine and entinostat: A phase 2 consortium/stand up 2 cancer study. Oncotarget. 2017;8(21):35326-35338
  141. 141. Buckowitz A et al. Microsatellite instability in colorectal cancer is associated with local lymphocyte infiltration and low frequency of distant metastases. British Journal of Cancer. 2005;92(9):1746-1753
  142. 142. Le DT et al. PD-1 blockade in tumors with mismatch-repair deficiency. The New England Journal of Medicine. 2015;372(26):2509-2520
  143. 143. Emambux S et al. Results and challenges of immune checkpoint inhibitors in colorectal cancer. Expert Opinion on Biological Therapy. 2018;18(5):561-573
  144. 144. Passardi A et al. Immune checkpoints as a target for colorectal cancer treatment. International Journal of Molecular Sciences. 2017;18(6). Artical ID: 1324
  145. 145. Abdelfatah E et al. Epigenetic therapy in gastrointestinal cancer: The right combination. Therapeutic Advances in Gastroenterology. 2016;9(4):560-579
  146. 146. Okada K et al. Combination of the histone deacetylase inhibitor depsipeptide and 5-fluorouracil upregulates major histocompatibility complex class II and p21 genes and activates caspase-3/7 in human colon cancer HCT-116 cells. Oncology Reports. 2016;36(4):1875-1885
  147. 147. Shah MY et al. microRNA therapeutics in cancer—An emerging concept. eBioMedicine. 2016;12:34-42
  148. 148. Klann TS et al. CRISPR-Cas9 epigenome editing enables high-throughput screening for functional regulatory elements in the human genome. Nature Biotechnology. 2017;35:561

Written By

Antja-Voy Hartley, Matthew Martin and Tao Lu

Submitted: 27 August 2018 Reviewed: 16 November 2018 Published: 09 January 2019