Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 3217 word(s) 3217 2021-09-27 04:09:02 |
2 Done Meta information modification 3217 2021-11-12 02:28:31 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Xiaolin, Z. Epigenetics and Cellular Metabolism. Encyclopedia. Available online: https://encyclopedia.pub/entry/15893 (accessed on 29 March 2024).
Xiaolin Z. Epigenetics and Cellular Metabolism. Encyclopedia. Available at: https://encyclopedia.pub/entry/15893. Accessed March 29, 2024.
Xiaolin, Zhang. "Epigenetics and Cellular Metabolism" Encyclopedia, https://encyclopedia.pub/entry/15893 (accessed March 29, 2024).
Xiaolin, Z. (2021, November 11). Epigenetics and Cellular Metabolism. In Encyclopedia. https://encyclopedia.pub/entry/15893
Xiaolin, Zhang. "Epigenetics and Cellular Metabolism." Encyclopedia. Web. 11 November, 2021.
Epigenetics and Cellular Metabolism
Edit

Epigenetics refers to the regulatory code that dictates gene expression or not and can be stably inherited in the absence of a constant genomic sequence. The current research content of epigenetics mainly includes DNA methylation and hydroxylmethylation, histone modifications, chromosome remodeling, and non-coding RNA regulation. In the early stage of CRC, DNA methylation status begins to change abnormally, mainly through the hypermethylation of some CpG islands leading to the down-regulation of gene expression and genome-wide hypomethylation, which cause genome instability to participate in tumorigenesis and development.

colorectal cancer epigenetics cellular metabolism targeted therapy tumorigenesis

1. Introduction

Colorectal cancer (CRC), a multi-step, multi-factor complex disease, is a common malignant tumor of the digestive tract. It has become the world’s third most common cancer and the second most common cause of death from cancer [1]. Both genetic and environmental risk factors play essential roles during the development of CRC [2]. In recent years, there have been large advances in cancer diagnosis technology. However, because the early symptoms are not dominant, most patients are diagnosed at an advanced stage, which affects the prognosis. Therefore, it is urgent to reduce the high incidence and mortality of CRC, and to develop advanced prevention and treatment strategies. For decades, people have elucidated many aspects of CRC biology, including genetics, epigenetics, metabolism and signaling pathways. Among them, epigenetic changes and cellular metabolic reprogramming play a key role in the occurrence and development of CRC, which is becoming more and more important.
Epigenetics refers to the regulatory code that dictates gene expression or not and can be stably inherited in the absence of a constant genomic sequence. The current research content of epigenetics mainly includes DNA methylation and hydroxylmethylation, histone modifications, chromosome remodeling, and non-coding RNA regulation. In the early stage of CRC, DNA methylation status begins to change abnormally, mainly through the hypermethylation of some CpG islands leading to the down-regulation of gene expression and genome-wide hypomethylation, which cause genome instability to participate in tumorigenesis and development [3][4]. Histone modifications are often closely related to DNA methylation and act together in the process of gene transcription. Covalent modifications in histones includes acetylation, methylation, phosphorylation, ADP-ribosylation, ubiquitination, succinylation and myristoylation, etc. Histone methylation modifications mainly include the monomethylation, dimethylation or trimethylation of histone H3 or H4, such as H3K4me2, H3K4me3, H3K9me3 and H3K27me3. Histone acetylation modifications mainly include histone H3 and H4 acetylation. Histone modification significantly affects gene expression; therefore, there are obvious histone modification abnormalities in CRC [5]. MicroRNA (miRNA) is a small single-stranded non-coding RNA that can regulate the expression of various oncogenes and tumor suppressor genes after transcription. The dysregulated expression of many miRNAs has been shown to mediate important signaling pathways in the multi-step carcinogenesis of CRC [6]. In recent years, long non-coding RNA (lncRNA) has also been shown to be highly associated with the tumorigenesis of CRC [7][8]. Even in mitochondrion, epigenetics also plays a fundamental role in energy hemostasis [9].
Cell division and proliferation requires a large amount of protein, lipid and nucleic acid as molecular raw materials and adenosine triphosphate (ATP) as energy, leading to the reorganization of anabolic flow in tumor cells. Metabolic reprogramming contributes to tumor progression and metastasis, which is considered to be an important hallmark of cancer [10]. The most dominant metabolic reprogramming is the aerobic glycolysis, also called the Warburg effect. Tumor cells increase glucose consumption to promote glycolysis, converting pyruvate to lactic acid even under aerobic conditions, which is different to that of the normal cells. However, glutamine metabolism and oxidative phosphorylation (OXPHOS), as well as lipometabolism and one-carbon pathways, are also altered in some tumor cells because of genetic heterogeneity and microenvironmental discrepancy [11][12][13][14][15]. Cellular metabolism is exquisitely modulated. Nonetheless, the regulatory axis in tumors are usually disorganized. This is the reason why dysregulated cell metabolism is observed in tumors. Cell metabolism is one of the essential factors contributing to carcinogenesis, and is also a result of malignancy [16][17]. Studies have found that metabolic reprogramming is widespread in CRC. In different types of cancers, including CRC, dysregulated metabolic pathways such as abnormal glycolysis, glutamate and lipid synthesis are often observed, leading to unlimited tumorigenesis [18].
It is worth noting that certain metabolic changes are known to occur at the epigenetic level, so that epigenetics and metabolism are highly intertwined in a mutually beneficial manner [19]. Metabolites produced in metabolic pathways, such as glycolysis cycles and OXPHOS, can be used as cofactors for epigenetic regulation [20]. Studies have also shown that epigenetics can regulate the expression of metabolism-related genes and affect the metabolic reprogramming of tumor cells [21][22][23]. Since many metabolic changes and epigenetic regulation are common in multiple cancer types, they become promising targets for tumor therapy. The purpose of this review is to describe the interactions between epigenetic changes and metabolic reprogramming, and to list the research progress of drugs targeting epigenetics and metabolism in CRC.

2. Epigenetics in CRC

2.1. DNA Methylation and Hydroxymethylation in CRC

2.1.1. DNA Methylation in CRC

DNA methylation is the covalent bonding of a methyl group at the 5′ carbon position of the cytosine of CpG dinucleotides under the action of DNA methyltransferases (DNMTs) to form 5-methylcytosine (5mC). DNMTs are mainly divided into two categories: DNA methylation maintenance enzyme DNMT1 and DNA de novo methylase DNMT3A, DNMT3B and DNMT3L [24]. In the human genome, CpG dinucleotides account for 10% and exist in two forms. One is dispersed in DNA and accounts for 70% to 80% of CpG dinucleotides, and this form is in the form of methyl. Another form, where CpG sites are highly clustered, is called CpG islands, which only account for 1% to 2% of the genome. They are mostly found in the 5′ promoter region of the gene, and can also extend to the exon region of the gene. Changes in normal DNA methylation patterns include DNA hypomethylation, which occurs in normal unmethylated regions of the genome, and DNA hypermethylation, which usually occurs on CpG islands [25].
Recent studies have shown that inactivation of multiple tumor-suppressor genes due to abnormal methylation of CpG islands plays an important role in the development of CRC [26]. DNA methylation has a unique subtype in CRC, which is called the CpG island methylation phenotype (CIMP) [27]. CIMP has been considered a significant direction for CRC research [28]. Evidence shows that this abnormal methylation has been found in the promoter regions of some vital tumor suppressor genes in CRC, including O6-methylguanine-DNA methyltransferase (MGMT), thrombospondin 1 (THBS1), tissue inhibitor of metalloproteinases 3 (TIMP3), p14ARF, p15INK4B (cyclin dependent kinase inhibitor 2B), p16INK4a, adenomatous polyposis coli (APC), deleted in colorectal carcinoma (DCC), MutL homolog 1 (MLH1), insulin-like growth factor 2 (IGF2), suppressor of cytokine signaling 1 (SOCS-1) and Runt-related transcription factor 3 (RUNX3) genes. The CpG island methylation in the promoter region of the above genes may play a crucial role in the occurrence and development of CRC, and the hypermethylation of the p16 promoter region is common in CRC. Moreover, hypermethylation of the p16 promoter region is widespread in CRC, and p16 methylation can be considered as an indicator of the prognosis of CRC [29].
Genome-wide hypomethylation is aberrant and involves early epigenetic changes during the development of CRC [30]. There are few studies into molecular events related to human genome hypomethylation, but there is evidence that gene-related specific hypomethylation changes are closely related to the development of CRC. In tumor cells, oncogenes are in a state of hypomethylation, and so they are activated. Abnormal hypomethylation of some genes, such as mothers against decapentaplegic homolog 3 (SMAD3), long interspersed nucleotide element-1 (LINE-1) and GDNF family receptor alpha 1 (GFRA1) genes, have been shown to be associated with the poor prognosis of CRC [31][32][33].

2.1.2. DNA Hydroxymethylation in CRC

During active and passive demethylation pathways, there are several kinds of intermediates, such as 5-hydroxymethylcytosine (5hmC), 5-formylcytosine (5fC), 5-carboxylcytosine (5caC) and 5-hydroxymethyluracil (5hmU) [9]. Recently, DNA 5hmC is also found to be a biomarker for tumors and may play an important role during tumorigenesis [34]. 5hmC regulates intestinal differentiation in adults, and its aberrant alterations lead to the disorder of gene expression in CRC [35]. Besides, DNA 5hmC is associated with malignant tumor behavior in patients and is considered to be an independent prognostic factor for overall survival, indicating that down-regulation of DNA 5hmC may be a useful biomarker for prognostic assessment of CRC [36]. Aberrant hydroxymethylation of arachidonate 15-lipoxygenase (ALOX15), growth hormone releasing hormone receptor (GHRHR), tissue factor pathway inhibitor 2 (TFPI2) and transketolase like 1 (TKTL1) at some genome position is correlated with CRC [37]. These lines of evidence indicate that 5hmC in genomes may be a promising biomarker for the diagnosis and prognosis of CRC.
Abnormal 5hmC modifications often occur at transcription start site (TSSs) regions and proximal regions closer to TSSs than that of the abnormal methylation [37]. Some studies have shown that promoters with 5hmC have a natural resistance to hypermethylation in CRC [38]. This result means that 5hmC also functions as an important regulator during transcription, and can be used as a therapeutic target. Vitamin C can promote decitabine or azacytidine-induced DNA 5hmC, leading to subsequent reactivation of p21WAF1 (CDKN1A), an epigenetically silenced tumor suppressor in CRC [39]. Treatment with anti-cancerous omega-3 polyunsaturated fatty acids in CRC shows an increase in the genomic DNA 5hmC [40].

2.2. Histone Post-Translational Modification

Histone modification is mainly the various modifications of the free N-terminal, including acetylation, methylation, phosphorylation, ubiquitination and so on. These modifications direct the fate and function of histones, which further alter the chromosomal substructures and subsequently gene transcription. Histone modification can catalyze the change of gene expression through histone modification enzymes, including writers, erasers and readers, which play important roles in the occurrence and development of malignant tumors. Therefore, the dysregulation of histone modification in CRC may change the expression of corresponding genes. Histone modifications can result in the activation or inhibition of transcription, depending on the location and type of modification [41]. The acetylation of lysine residues is related to transcriptional activation and is considered an “activation marker” [42]. Studies have shown that the acetylation level of H3K27 in CRC is significantly higher than that in normal tissues [43]. Ashktorab et al. demonstrated that H3K12 and H3K18 acetylation are significantly increased in moderately and well-differentiated colon cancer, and are decreased in poorly differentiated colon cancer [44]. However, lysine methylation can cause many consequences, such as promotion of transcriptional activity, regulation of transcriptional inhibition and so on. H3K4me3 is enriched in transcriptionally active gene promoters, while H3K9me3 and H3K27me3 are present in transcriptional repressive gene promoters. Benard et al. demonstrated that the high expression of activated histone modification H3K4me3 and the low expression of silent modification H3K9me3 may be related to the poor clinical outcome of CRC [45].
Histone modification is dynamically regulated by a variety of enzymes. For example, histone methyltransferases (HMTs) and histone demethylases (HDMs) are responsible for increasing and removing methyl groups, respectively. Histone acetyltransferases (HATs) can increase the acetyl groups of lysine residues, while histone deacetylases (HDACs) can remove acetyl groups. The dynamic balance of these two enzyme activities maintains the proper state of histone acetylation. Therefore, the dysregulation of histone modification-related enzyme activities is related to the development of cancer [46]. For instance, H3K9 methyltransferase G9a promotes gastric cancer progression and suppresses its autophagy by activating mTOR signaling [47]. H4K16 acetyltransferase MYST1/KAT8 contributes to glioblastoma progression by activating EGFR signaling [48]. H3K9 deacetylase SIRT6, one of the class III HDACs, are important for melanoma and gliomblatoma progression via inhibition of the transcription of glycolytic genes [49][50]. In CRC, epigenetics also functions as a fundamental role during carcinogenesis and subsequent tumor development and metastasis. For instance, H3K79 methyltransferase DOT1L supports CRC cancer progression via epigenetically promoting c-Myc expression [51]. Histone demethylase JMJD2D can interacts with β-catenin, activating the transcription of its target genes, which further supports cell proliferation, migration, and invasion of CRC [52]. Histone deacetylase SIRT1 inhibits CRC metastasis by transcriptional repression of miR-15b-5p [53]. These results indicate that post-translational modifications of the histone are important for CRC tumorigenesis and may be used as therapeutic targets for targeted therapy. Importantly, inhibitors targeting histone deacetylase have become promising drugs for the treatment of CRC [54].

2.3. Non-Coding RNAs in CRC

2.3.1. MicroRNAs in CRC

MiRNA is a type of non-coding single-stranded RNA molecule with a length of about 22 nucleotides encoded by endogenous gene. By binding to a specific site in the 3′ UTR region of the target gene’s mRNA, miRNA promotes the degradation of the target gene’s mRNA, thereby exercising the function of regulating gene expression. A large number of studies have shown that the abnormal regulation of miRNAs plays an important role in the development and metastasis of CRC, showing dual effects of promoting tumors or suppressing tumors [55]. Deng et al. proved that overexpression of miR-21 can promote cell proliferation and invasion, and can alleviate the inhibitory effect of the chemotherapy drug 5-fluorouracil (5-Fu), the first-line drug for the treatment of CRC, on the proliferation and invasion of HT-29 cells [56]. The low expression of miR-133b is related to the poor survival rate and metastasis of CRC [57]. In SW620 and HT-29 cell lines, the ectopic expression of miR-133b can effectively inhibit tumor cell proliferation and apoptosis in vitro and in vivo by directly targeting receptor tyrosine kinases [58]. It has been found that the expression of miR-143 and miR-145 is reduced in CRC. Transfection of miR-143 and miR-145 precursors into HCT116 or SW480 cell lines, respectively, can observe cell growth inhibition and drug sensitivity enhancement [59]. More and more evidence shows that miRNAs can be used as markers for prognostic evaluation and efficacy evaluation of CRC [6]. Therefore, inhibiting oncogenic miRNAs and restoring the tumor-suppressive miRNAs can be used as promising strategies for the treatment of CRC.

2.3.2. LncRNAs in CRC

Long non-coding RNAs (lncRNAs) are transcripts longer than 200 nucleotides that do not encode proteins. They can interact with almost all molecules in cells, including RNA, DNA, proteins, and even metabolites. Recently, lncRNAs such as cancer-susceptibility 15 (CACS15), cytoskeleton regulator RNA (CYTOR), HOX transcript antisense RNA (HOTAIR), metastasis associated lung adenocarcinoma transcript 1 (MALAT1), taurine upregulated 1 (TUG1), nuclear paraspeckle assembly transcript 1 (NEAT1), miR-17-92a-1 cluster host gene (MIR17HG) and so on, have been shown to be tightly correlated with the prognosis of CRC and function as competitive endogenous RNAs (ceRNAs) [60]. These ceRNA networks, formed by the lncRNA/miRNA/mRNA interactions, have been found in a wide range of biological processes in CRC, including liver metastasis, epithelial-to-mesenchymal transition (EMT), inflammation, and chemotherapy/radiation resistance [60]. Moreover, lncRNAs also function as protein partners to regulate protein stabilization. For instance, lncRNA nuclear paraspeckle assembly transcript 1 (NEAT1) directly interacts with Ddx5 protein, enhancing its stability, and sequentially activating Wnt signaling in CRC [61]. LncRNA colon cancer associated transcript 1-L (CCAT1-L), which is specifically transcribed from the upstream 515kb site of MYC in human colorectal carcinoma, promotes long-term chromatin looping, which further enhances MYC transcription regulation via long-range interactions between the MYC promoter and its enhancers [62]. These findings indicate that lncRNAs may be potential biomarkers and therapeutic targets for CRC.

3. Cellular Metabolism in CRC

3.1. Aerobic Glycolysis in CRC

Current studies have found that the Warburg effect is closely related to the occurrence, invasion and metastasis of CRC [63][64]. Compared with normal mucosa, many metabolic pathways of CRC have changed. The metabolic changes of CRC is the result of metabolic reprogramming guided by the activation of proto-oncogenes and the inactivation of tumor suppressor genes [65]. Many genes and proteins related to glucose uptake and glycolysis are dysregulated in CRC, including K-RAS, hypoxia-inducible factor (HIF), MYC, PI3K/AKT/mTOR axis and their related signaling pathways and tumor suppressor gene p53 [66]. In fact, the dysregulation of many of the above genes is related to tumor aggressiveness and poor prognosis of CRC [67]. For instance, K-RAS mutations occur in approximately 40% of CRCs [68][69]. Yun et al. found that glucose transporter 1 (GLUT1) levels were up-regulated in the transcripts of CRC cell lines in the K-RAS gene mutation state, and K-RAS mutant cells showed increased glucose uptake and glycolysis [70]. Another major gene is HIF, which is stimulated by the hypoxic microenvironment. Up-regulation of HIF-1α expression was found in 55% of CRC biopsies [71]. HIF can up-regulate the expression of hundreds of genes, such as GLUT1, c-Myc, vascular endothelial growth factor (VEGF), and glycolysis-related genes. P53 is also one of the most import and mutant genes in CRCs. More than 40% of CRCs carry mutations in the tumor suppressor gene p53, leading to loss or gain of function [72]. Ma et al. found that the glycolysis of wild-type p53 CRC HCT116 cells contributes about 40% to ATP, while the contribution of p53 mutant cell glycolysis to ATP is up-regulated to about 66% [73].

3.2. Glutamine Metabolism in CRC

In addition to glycolysis, glutaminolysis is also important in CRC. Glutamine is the fastest consumed amino acid in tumor cells. The growth of cancer cells is dependent on glutamine, and tumor cells cannot grow in a medium lacking glutamine. After glutamine enters the cell via the transporter alanine-serine-cysteine transporter 2 (ASCT2), it is hydrolyzed into glutamate and ammonia under the action of glutaminase (GLS), glutamine can be converted into α-ketoglutaric acid (α-KG) and enter the tricarboxylic acid (TCA) cycle, which provides intermediate metabolites and energy for cells [74]. The gene expression of glutamine metabolism-related enzymes in CRC is up-regulated. For example, the level of GLS1 mRNA in CRC is significantly higher than that in neighboring normal tissues [75]. These changes are regulated by oncogenes and tumor suppressor genes. Among them, the proto-oncogene c-Myc is the main transcription factor that promotes the metabolism of glutamine in cancer cells. ASCT2 and GLS1 can be transcriptionally activated by c-Myc, thereby promoting the uptake and metabolism of glutamine [76][77].

3.3. Biosynthetic Metabolism in CRC

In addition to energy catabolism, the biosynthetic metabolism of CRC is also very different from normal human tissues, which is mainly reflected in lipids, proteins and nucleotides. For instance, comparing CRC to normal colon tissue, an integrated analysis of miRNA and mRNA endorses a twenty miRNAs signature, which is tightly correlated with metabolic genes including long-chain acyl-CoA synthetase 6 (ACSL6) and phosphoribosyl pyrophosphate synthetase 1 and 2 (PRPS1/2) [78]. Nucleotides in tumors are mainly synthesized by de novo pathways, and their de novo synthetase activities are relatively high in CRC [79][80].
The lipid metabolism of normal cells is maintained at a low level, and the activity of enzymes related to lipid metabolism is also low. Abnormal lipid metabolism of tumor cells is mainly manifested by de novo fatty acid synthesis and enhanced lipid synthesis [81]. Enzymes related to fatty acid synthesis have been discovered, such as ATP citrate lyase (ACLY), acetyl-CoA carboxylase (ACC) and fatty acid synthetase (FASN) is highly expressed in CRC [82][83]. Nevertheless, several genes, such as acetyl-CoA acyltransferase 2 (ACAA2), acyl-CoA dehydrogenase short chain (ACADS), acetyl-CoA acetyltransferase 1 (ACAT1), acyl-CoA oxidase 1 (ACOX), carnitine palmitoyltransferase 1A (CPT1A) and 3-hydroxy-3-methylglutaryl-CoA synthase 2 (HMGCS2), were shown to be downregulated in CRC, compared with distant normal colon tissue (NTC) in a transcriptome study [84]. This makes it complex for the relationship of lipogenesis and tumorigenesis in CRC.
Protein homeostasis is also important for cell fate. Interestingly, protein synthesis is also accurately modulated in cells and deregulation of it may be an important cause for carcinogensis. It is showed that the PI3K/AKT/mTOR signaling pathway is abnormally activated in CRC, which not only enhances glycolysis, but also enhances protein synthesis [85]. Activated mTOR can regulate downstream pathways, such as eukaryotic initiation factor 4E binding protein 1 (4EBP1) and S6 kinase 1 (S6K1), which are key regulators of protein translation [86].

References

  1. Purushothaman, V.L.; Cuomo, R.E.; Garland, C.F.; Mackey, T.K. Could age increase the strength of inverse association between ultraviolet B exposure and colorectal cancer? BMC Public Health 2021, 21, 1238.
  2. Dekker, E.; Tanis, P.J.; Vleugels, J.L.A.; Kasi, P.M.; Wallace, M.B. Colorectal cancer. Lancet 2019, 394, 1467–1480.
  3. Jair, K.-W.; Bachman, K.E.; Suzuki, H.; Ting, A.H.; Rhee, I.; Yen, R.-W.C.; Baylin, S.B.; Schuebel, K.E. De novo CpG island methylation in human cancer cells. Cancer Res. 2006, 66, 682–692.
  4. Herman, J.G.; Baylin, S.B. Gene silencing in cancer in association with promoter hypermethylation. N. Engl. J. Med. 2003, 349, 2042–2054.
  5. Qin, J.; Wen, B.; Liang, Y.; Yu, W.; Li, H. Histone Modifications and their Role in Colorectal Cancer (Review). Pathol. Oncol. Res. 2020, 26, 2023–2033.
  6. To, K.K.; Tong, C.W.; Wu, M.; Cho, W.C. MicroRNAs in the prognosis and therapy of colorectal cancer: From bench to bedside. World J. Gastroenterol. 2018, 24, 2949–2973.
  7. Dastmalchi, N.; Safaralizadeh, R.; Nargesi, M.M. LncRNAs: Potential Novel Prognostic and Diagnostic Biomarkers in Colorectal Cancer. Curr. Med. Chem. 2020, 27, 5067–5077.
  8. Bermúdez, M.; Aguilar-Medina, M.; Lizárraga-Verdugo, E.; Avendaño-Félix, M.; Silva-Benítez, E.; López-Camarillo, C.; Ramos-Payán, R. LncRNAs as Regulators of Autophagy and Drug Resistance in Colorectal Cancer. Front. Oncol. 2019, 9, 1008.
  9. Dong, Z.; Pu, L.; Cui, H. Mitoepigenetics and Its Emerging Roles in Cancer. Front. Cell Dev. Biol. 2020, 8, 4.
  10. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674.
  11. Altman, B.J.; Stine, Z.E.; Dang, C.V. From Krebs to clinic: Glutamine metabolism to cancer therapy. Nat. Rev. Cancer 2016, 16, 619–634.
  12. Cluntun, A.A.; Lukey, M.J.; Cerione, R.A.; Locasale, J.W. Glutamine Metabolism in Cancer: Understanding the Heterogeneity. Trends Cancer 2017, 3, 169–180.
  13. Ashton, T.M.; McKenna, W.G.; Kunz-Schughart, L.A.; Higgins, G.S. Oxidative Phosphorylation as an Emerging Target in Cancer Therapy. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2018, 24, 2482–2490.
  14. Newman, A.C.; Maddocks, O.D.K. One-carbon metabolism in cancer. Br. J. Cancer 2017, 116, 1499–1504.
  15. Zhao, E.; Hou, J.; Cui, H. Serine-glycine-one-carbon metabolism: Vulnerabilities in MYCN-amplified neuroblastoma. Oncogenesis 2020, 9, 14.
  16. Cairns, R.A.; Harris, I.S.; Mak, T.W. Regulation of cancer cell metabolism. Nat. Rev. Cancer 2011, 11, 85–95.
  17. Pavlova, N.N.; Thompson, C.B. The Emerging Hallmarks of Cancer Metabolism. Cell Metab. 2016, 23, 27–47.
  18. La Vecchia, S.; Sebastián, C. Metabolic pathways regulating colorectal cancer initiation and progression. Semin. Cell Dev. Biol. 2020, 98, 63–70.
  19. Sun, L.; Zhang, H.; Gao, P. Metabolic reprogramming and epigenetic modifications on the path to cancer. Protein Cell 2021, 1–43.
  20. Aspeslagh, S.; Morel, D.; Soria, J.C.; Postel-Vinay, S. Epigenetic modifiers as new immunomodulatory therapies in solid tumours. Ann. Oncol. 2018, 29, 812–824.
  21. Dong, Z.; Cui, H. Epigenetic modulation of metabolism in glioblastoma. Semin. Cancer Biol. 2019, 57, 45–51.
  22. Zhu, S.; Dong, Z.; Ke, X.; Hou, J.; Zhao, E.; Zhang, K.; Wang, F.; Yang, L.; Xiang, Z.; Cui, H. The roles of sirtuins family in cell metabolism during tumor development. Semin. Cancer Biol. 2019, 57, 59–71.
  23. Dong, Z.; Cui, H. Chapter 11—Sirtuins and cellular metabolism in cancers. In Sirtuin Biology in Cancer and Metabolic Disease; Maiese, K., Ed.; Academic Press: London, UK, 2021; pp. 195–217.
  24. Margot, J.B.; Cardoso, M.C.; Leonhardt, H. Mammalian DNA methyltransferases show different subnuclear distributions. J. Cell. Biochem. 2001, 83, 373–379.
  25. Jung, G.; Hernández-Illán, E.; Moreira, L.; Balaguer, F.; Goel, A. Epigenetics of colorectal cancer: Biomarker and therapeutic potential. Nat. Rev. Gastroenterol. Hepatol. 2020, 17, 111–130.
  26. Peng, X.; Qing, Z. CpG island methylation of tumor suppressor genes in colorectal cancer. Chin. J. Gen. Surg. 2012, 21, 1279–1282.
  27. Lao, V.V.; Grady, W.M. Epigenetics and colorectal cancer. Nat. Rev. Gastroenterol. Hepatol. 2011, 8, 686–700.
  28. Toyota, M.; Ahuja, N.; Ohe-Toyota, M.; Herman, J.G.; Baylin, S.B.; Issa, J.P. CpG island methylator phenotype in colorectal cancer. Proc. Natl. Acad. Sci. USA 1999, 96, 8681–8686.
  29. Bahrami, A.; Hassanian, S.M.; Khazaei, M.; Gharib, M.; Rahmani, M.; Fiuji, H.; Jazayeri, M.H.; Moetamani-Ahmadi, M.; Ferns, G.A.; Avan, A. The 9p21 locus as a potential therapeutic target and prognostic marker in colorectal cancer. Pharmacogenomics 2018, 19, 463–474.
  30. Cavestro, G.M.; Mannucci, A.; Zuppardo, R.A.; Di Leo, M.; Stoffel, E.; Tonon, G. Early onset sporadic colorectal cancer: Worrisome trends and oncogenic features. Dig. Liver Dis. 2018, 50, 521–532.
  31. Ogino, S.; Nosho, K.; Kirkner, G.J.; Kawasaki, T.; Chan, A.T.; Schernhammer, E.S.; Giovannucci, E.L.; Fuchs, C.S. A cohort study of tumoral LINE-1 hypomethylation and prognosis in colon cancer. J. Natl. Cancer Inst. 2008, 100, 1734–1738.
  32. Lam, K.; Pan, K.; Linnekamp, J.F.; Medema, J.P.; Kandimalla, R. DNA methylation based biomarkers in colorectal cancer: A systematic review. Biochim. Biophys. Acta 2016, 1866, 106–120.
  33. Dong, Z.; Dai, L.; Zhang, Y.; Fang, C.; Shi, G.; Chen, Y.; Li, J.; Wang, Q.; Fu, J.; Yu, Y.; et al. Hypomethylation of GDNF family receptor alpha 1 promotes epithelial-mesenchymal transition and predicts metastasis of colorectal cancer. PLoS Genet. 2020, 16, e1009159.
  34. Jeschke, J.; Collignon, E.; Fuks, F. Portraits of TET-mediated DNA hydroxymethylation in cancer. Curr. Opin. Genet. Dev. 2016, 36, 16–26.
  35. Chapman, C.G.; Mariani, C.J.; Wu, F.; Meckel, K.; Butun, F.; Chuang, A.; Madzo, J.; Bissonnette, M.B.; Kwon, J.H.; Godley, L.A. TET-catalyzed 5-hydroxymethylcytosine regulates gene expression in differentiating colonocytes and colon cancer. Sci. Rep. 2015, 5, 17568.
  36. Zhong, A.; Tian, Y.; Zhang, H.; Lai, M. DNA hydroxymethylation of colorectal primary carcinoma and its association with survival. J. Surg. Oncol. 2018, 117, 1029–1037.
  37. Li, M.; Gao, F.; Xia, Y.; Tang, Y.; Zhao, W.; Jin, C.; Luo, H.; Wang, J.; Li, Q.; Wang, Y. Filtrating colorectal cancer associated genes by integrated analyses of global DNA methylation and hydroxymethylation in cancer and normal tissue. Sci. Rep. 2016, 6, 31826.
  38. Uribe-Lewis, S.; Stark, R.; Carroll, T.; Dunning, M.J.; Bachman, M.; Ito, Y.; Stojic, L.; Halim, S.; Vowler, S.L.; Lynch, A.G.; et al. 5-hydroxymethylcytosine marks promoters in colon that resist DNA hypermethylation in cancer. Genome Biol. 2015, 16, 69.
  39. Gerecke, C.; Schumacher, F.; Edlich, A.; Wetzel, A.; Yealland, G.; Neubert, L.K.; Scholtka, B.; Homann, T.; Kleuser, B. Vitamin C promotes decitabine or azacytidine induced DNA hydroxymethylation and subsequent reactivation of the epigenetically silenced tumour suppressor CDKN1A in colon cancer cells. Oncotarget 2018, 9, 32822–32840.
  40. Huang, Q.; Mo, M.; Zhong, Y.; Yang, Q.; Zhang, J.; Ye, X.; Zhang, L.; Cai, C. The Anticancer Role of Omega-3 Polyunsaturated Fatty Acids was Closely Associated with the Increase in Genomic DNA Hydroxymethylation. Anti-Cancer Agents Med. Chem. 2019, 19, 330–336.
  41. Sharma, S.; Kelly, T.K.; Jones, P.A. Epigenetics in cancer. Carcinogenesis 2010, 31, 27–36.
  42. Hebbes, T.R.; Thorne, A.W.; Crane-Robinson, C. A direct link between core histone acetylation and transcriptionally active chromatin. EMBO J. 1988, 7, 1395–1402.
  43. Karczmarski, J.; Rubel, T.; Paziewska, A.; Mikula, M.; Bujko, M.; Kober, P.; Dadlez, M.; Ostrowski, J. Histone H3 lysine 27 acetylation is altered in colon cancer. Clin. Proteom. 2014, 11, 24.
  44. Ashktorab, H.; Belgrave, K.; Hosseinkhah, F.; Brim, H.; Nouraie, M.; Takkikto, M.; Hewitt, S.; Lee, E.L.; Dashwood, R.H.; Smoot, D. Global histone H4 acetylation and HDAC2 expression in colon adenoma and carcinoma. Dig. Dis. Sci. 2009, 54, 2109–2117.
  45. Benard, A.; Goossens-Beumer, I.J.; van Hoesel, A.Q.; de Graaf, W.; Horati, H.; Putter, H.; Zeestraten, E.C.M.; van de Velde, C.J.H.; Kuppen, P.J.K. Histone trimethylation at H3K4, H3K9 and H4K20 correlates with patient survival and tumor recurrence in early-stage colon cancer. BMC Cancer 2014, 14, 531.
  46. Valencia, A.M.; Kadoch, C. Chromatin regulatory mechanisms and therapeutic opportunities in cancer. Nat. Cell Biol. 2019, 21, 152–161.
  47. Yin, C.; Ke, X.; Zhang, R.; Hou, J.; Dong, Z.; Wang, F.; Zhang, K.; Zhong, X.; Yang, L.; Cui, H. G9a promotes cell proliferation and suppresses autophagy in gastric cancer by directly activating mTOR. FASEB J. 2019, 33, 14036–14050.
  48. Dong, Z.; Zou, J.; Li, J.; Pang, Y.; Liu, Y.; Deng, C.; Chen, F.; Cui, H. MYST1/KAT8 contributes to tumor progression by activating EGFR signaling in glioblastoma cells. Cancer Med. 2019, 8, 7793–7808.
  49. Dong, Z.; Yang, J.; Li, L.; Tan, L.; Shi, P.; Zhang, J.; Zhong, X.; Ge, L.; Wu, Z.; Cui, H. FOXO3a-SIRT6 axis suppresses aerobic glycolysis in melanoma. Int. J. Oncol. 2020, 56, 728–742.
  50. Dong, Z.; Zhong, X.; Lei, Q.; Chen, F.; Cui, H. Transcriptional activation of SIRT6 via FKHRL1/FOXO3a inhibits the Warburg effect in glioblastoma cells. Cell. Signal. 2019, 60, 100–113.
  51. Yang, L.; Lei, Q.; Li, L.; Yang, J.; Dong, Z.; Cui, H. Silencing or inhibition of H3K79 methyltransferase DOT1L induces cell cycle arrest by epigenetically modulating c-Myc expression in colorectal cancer. Clin. Epigenet. 2019, 11, 199.
  52. Peng, K.; Kou, L.; Yu, L.; Bai, C.; Li, M.; Mo, P.; Li, W.; Yu, C. Histone Demethylase JMJD2D Interacts With β-Catenin to Induce Transcription and Activate Colorectal Cancer Cell Proliferation and Tumor Growth in Mice. Gastroenterology 2019, 156, 1112–1126.
  53. Sun, L.N.; Zhi, Z.; Chen, L.Y.; Zhou, Q.; Li, X.M.; Gan, W.J.; Chen, S.; Yang, M.; Liu, Y.; Shen, T.; et al. SIRT1 suppresses colorectal cancer metastasis by transcriptional repression of miR-15b-5p. Cancer Lett. 2017, 409, 104–115.
  54. Tampakis, A.; Tampaki, E.C.; Nebiker, C.A.; Kouraklis, G. Histone deacetylase inhibitors and colorectal cancer: What is new? Anticancer Agents Med. Chem. 2014, 14, 1220–1227.
  55. Balacescu, O.; Sur, D.; Cainap, C.; Visan, S.; Cruceriu, D.; Manzat-Saplacan, R.; Muresan, M.-S.; Balacescu, L.; Lisencu, C.; Irimie, A. The Impact of miRNA in Colorectal Cancer Progression and Its Liver Metastases. Int. J. Mol. Sci. 2018, 19, 3711.
  56. Deng, J.; Lei, W.; Fu, J.-C.; Zhang, L.; Li, J.-H.; Xiong, J.-P. Targeting miR-21 enhances the sensitivity of human colon cancer HT-29 cells to chemoradiotherapy in vitro. Biochem. Biophys. Res. Commun. 2014, 443, 789–795.
  57. Akçakaya, P.; Ekelund, S.; Kolosenko, I.; Caramuta, S.; Ozata, D.M.; Xie, H.; Lindforss, U.; Olivecrona, H.; Lui, W.-O. miR-185 and miR-133b deregulation is associated with overall survival and metastasis in colorectal cancer. Int. J. Oncol. 2011, 39, 311–318.
  58. Hu, G.; Chen, D.; Li, X.; Yang, K.; Wang, H.; Wu, W. miR-133b regulates the MET proto-oncogene and inhibits the growth of colorectal cancer cells in vitro and in vivo. Cancer Biol. Ther. 2010, 10, 190–197.
  59. Gomes, S.E.; Simões, A.E.S.; Pereira, D.M.; Castro, R.E.; Rodrigues, C.M.P.; Borralho, P.M. miR-143 or miR-145 overexpression increases cetuximab-mediated antibody-dependent cellular cytotoxicity in human colon cancer cells. Oncotarget 2016, 7, 9368–9387.
  60. Wang, L.; Cho, K.B.; Li, Y.; Tao, G.; Xie, Z.; Guo, B. Long Noncoding RNA (lncRNA)-Mediated Competing Endogenous RNA Networks Provide Novel Potential Biomarkers and Therapeutic Targets for Colorectal Cancer. Int. J. Mol. Sci. 2019, 20, 5758.
  61. Zhang, M.; Weng, W.; Zhang, Q.; Wu, Y.; Ni, S.; Tan, C.; Xu, M.; Sun, H.; Liu, C.; Wei, P.; et al. The lncRNA NEAT1 activates Wnt/β-catenin signaling and promotes colorectal cancer progression via interacting with DDX5. J. Hematol. Oncol. 2018, 11, 113.
  62. Xiang, J.F.; Yin, Q.F.; Chen, T.; Zhang, Y.; Zhang, X.O.; Wu, Z.; Zhang, S.; Wang, H.B.; Ge, J.; Lu, X.; et al. Human colorectal cancer-specific CCAT1-L lncRNA regulates long-range chromatin interactions at the MYC locus. Cell Res. 2014, 24, 513–531.
  63. Hussain, A.; Qazi, A.K.; Mupparapu, N.; Guru, S.K.; Kumar, A.; Sharma, P.R.; Singh, S.K.; Singh, P.; Dar, M.J.; Bharate, S.B.; et al. Modulation of glycolysis and lipogenesis by novel PI3K selective molecule represses tumor angiogenesis and decreases colorectal cancer growth. Cancer Lett. 2016, 374, 250–260.
  64. Saunier, E.; Antonio, S.; Regazzetti, A.; Auzeil, N.; Laprévote, O.; Shay, J.W.; Coumoul, X.; Barouki, R.; Benelli, C.; Huc, L.; et al. Resveratrol reverses the Warburg effect by targeting the pyruvate dehydrogenase complex in colon cancer cells. Sci. Rep. 2017, 7, 6945.
  65. Brown, R.E.; Short, S.P.; Williams, C.S. Colorectal Cancer and Metabolism. Curr. Colorectal Cancer Rep. 2018, 14, 226–241.
  66. Neitzel, C.; Demuth, P.; Wittmann, S.; Fahrer, J. Targeting Altered Energy Metabolism in Colorectal Cancer: Oncogenic Reprogramming, the Central Role of the TCA Cycle and Therapeutic Opportunities. Cancers 2020, 12, 1731.
  67. Fang, S.; Fang, X. Advances in glucose metabolism research in colorectal cancer. Biomed. Rep. 2016, 5, 289–295.
  68. 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.
  69. Ucar, G.; Ergun, Y.; Esen, S.A.; Acikgoz, Y.; Dirikoc, M.; Esen, İ.; Bal, Ö.; Uncu, D. Prognostic and predictive value of KRAS mutation number in metastatic colorectal cancer. Medicine 2020, 99, e22407.
  70. Yun, J.; Rago, C.; Cheong, I.; Pagliarini, R.; Angenendt, P.; Rajagopalan, H.; Schmidt, K.; Willson, J.K.V.; Markowitz, S.; Zhou, S.; et al. Glucose deprivation contributes to the development of KRAS pathway mutations in tumor cells. Science 2009, 325, 1555–1559.
  71. Cao, D.; Hou, M.; Guan, Y.-s.; Jiang, M.; Yang, Y.; Gou, H.-f. Expression of HIF-1alpha and VEGF in colorectal cancer: Association with clinical outcomes and prognostic implications. BMC Cancer 2009, 9, 432.
  72. Olivier, M.; Hollstein, M.; Hainaut, P. TP53 mutations in human cancers: Origins, consequences, and clinical use. Cold Spring Harb. Perspect. Biol. 2010, 2, a001008.
  73. Ma, W.; Sung, H.J.; Park, J.Y.; Matoba, S.; Hwang, P.M. A pivotal role for p53: Balancing aerobic respiration and glycolysis. J. Bioenerg. Biomembr. 2007, 39, 243–246.
  74. Zhang, J.; Pavlova, N.N.; Thompson, C.B. Cancer cell metabolism: The essential role of the nonessential amino acid, glutamine. EMBO J. 2017, 36, 1302–1315.
  75. Xiang, L.; Mou, J.; Shao, B.; Wei, Y.; Liang, H.; Takano, N.; Semenza, G.L.; Xie, G. Glutaminase 1 expression in colorectal cancer cells is induced by hypoxia and required for tumor growth, invasion, and metastatic colonization. Cell Death Dis. 2019, 10, 40.
  76. Gao, P.; Tchernyshyov, I.; Chang, T.-C.; Lee, Y.-S.; Kita, K.; Ochi, T.; Zeller, K.I.; De Marzo, A.M.; Van Eyk, J.E.; Mendell, J.T.; et al. c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature 2009, 458, 762–765.
  77. Xu, X.; Li, J.; Sun, X.; Guo, Y.; Chu, D.; Wei, L.; Li, X.; Yang, G.; Liu, X.; Yao, L.; et al. Tumor suppressor NDRG2 inhibits glycolysis and glutaminolysis in colorectal cancer cells by repressing c-Myc expression. Oncotarget 2015, 6, 26161–26176.
  78. Angius, A.; Uva, P.; Pira, G.; Muroni, M.R.; Sotgiu, G.; Saderi, L.; Uleri, E.; Caocci, M.; Ibba, G.; Cesaraccio, M.R.; et al. Integrated Analysis of miRNA and mRNA Endorses a Twenty miRNAs Signature for Colorectal Carcinoma. Int. J. Mol. Sci. 2019, 20, 4067.
  79. Shiyong, L.; Bo, Y.; Ping, A.; Zhenjia, L.; Shujun, Y. Expression of thymidylate synthase gene and recurrence of colorectal carcinoma: Their relation and clinical significance. Chin. J. Surg. 2000, 38, 60–62.
  80. Houqiang, L.; Yinghao, Y. TS gene polymorphisms in gastric cancer and colorectal cancer. World Chin. J. Dig. 2013, 21, 865–872.
  81. Menendez, J.A.; Lupu, R. Fatty acid synthase and the lipogenic phenotype in cancer pathogenesis. Nat. Rev. Cancer 2007, 7, 763–777.
  82. Khwairakpam, A.D.; Shyamananda, M.S.; Sailo, B.L.; Rathnakaram, S.R.; Padmavathi, G.; Kotoky, J.; Kunnumakkara, A.B. ATP citrate lyase (ACLY): A promising target for cancer prevention and treatment. Curr. Drug Targets 2015, 16, 156–163.
  83. Zhan, Y.; Ginanni, N.; Tota, M.R.; Wu, M.; Bays, N.W.; Richon, V.M.; Kohl, N.E.; Bachman, E.S.; Strack, P.R.; Krauss, S. Control of cell growth and survival by enzymes of the fatty acid synthesis pathway in HCT-116 colon cancer cells. Clin. Cancer Res. 2008, 14, 5735–5742.
  84. Pira, G.; Uva, P.; Scanu, A.M.; Rocca, P.C.; Murgia, L.; Uleri, E.; Piu, C.; Porcu, A.; Carru, C.; Manca, A.; et al. Landscape of transcriptome variations uncovering known and novel driver events in colorectal carcinoma. Sci. Rep. 2020, 10, 432.
  85. Narayanankutty, A. PI3K/Akt/mTOR Pathway as a Therapeutic Target for Colorectal Cancer: A Review of Preclinical and Clinical Evidence. Curr. Drug Targets 2019, 20, 1217–1226.
  86. Hua, H.; Kong, Q.; Zhang, H.; Wang, J.; Luo, T.; Jiang, Y. Targeting mTOR for cancer therapy. J. Hematol. Oncol. 2019, 12, 71.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 367
Revisions: 2 times (View History)
Update Date: 12 Nov 2021
1000/1000