Next Article in Journal
Exploring the Dynamic Crosstalk between the Immune System and Genetics in Gastrointestinal Stromal Tumors
Previous Article in Journal
Combination of Small Extracellular Vesicle-Derived Annexin A2 Protein and mRNA as a Potential Predictive Biomarker for Chemotherapy Responsiveness in Aggressive Triple-Negative Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epimutations and Their Effect on Chromatin Organization: Exciting Avenues for Cancer Treatment

Department of Physiological Sciences, College of Veterinary Medicine, Oklahoma State University, Stillwater, OK 74078, USA
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(1), 215; https://doi.org/10.3390/cancers15010215
Submission received: 30 September 2022 / Revised: 14 December 2022 / Accepted: 28 December 2022 / Published: 29 December 2022
(This article belongs to the Section Molecular Cancer Biology)

Abstract

:

Simple Summary

Epigenetic modifications, such as DNA methylation and histone modification, have been found to alter in various cancer types. These modifications lead to uncontrolled cellular proliferation, evasion from apoptosis, and metastasis. Deregulation in epigenetic pathways often results in the suppression of tumor-suppression genes or activation of oncogenes in cancers. Inhibitors targeting deregulated enzymes can restore balance by reactivating altered pathways. Several inhibitors that target DNA methylation and histone modifications are currently being used in clinics and have shown promising results in cancer therapeutics.

Abstract

The three-dimensional architecture of genomes is complex. It is organized as fibers, loops, and domains that form high-order structures. By using different chromosome conformation techniques, the complex relationship between transcription and genome organization in the three-dimensional organization of genomes has been deciphered. Epigenetic changes, such as DNA methylation and histone modification, are the hallmark of cancers. Tumor initiation, progression, and metastasis are linked to these epigenetic modifications. Epigenetic inhibitors can reverse these altered modifications. A number of epigenetic inhibitors have been approved by FDA that target DNA methylation and histone modification. This review discusses the techniques involved in studying the three-dimensional organization of genomes, DNA methylation and histone modification, epigenetic deregulation in cancer, and epigenetic therapies targeting the tumor.

1. Introduction

In the Eukaryotic cell, genetic material is organized into a condensed structure inside the nucleus called chromatin. Chromatin is composed of ~147 base pairs (bp) of DNA wrapped around nucleosomes, which consist of octamers of core histone proteins H2A, H2B, H3, and H4 [1]. In the nucleus, chromatin forms higher-order structures such as fibers, loops, domains, etc. These fibers are typically 5–24 nm in diameter and folded irregularly into loops and topologically associated domains (TAD) [2,3]. These loops range in size from kilobases (Kb) to megabases (Mb) [4]. The formation of chromatin loops is associated with the activation or repression of genes [5,6]. The looping of chromatin brings upstream regulatory elements, such as enhancers and promoters, closer to neighboring genes which in turn regulate their expression. TADs are key units of the three-dimensional (3D) genome organization. TADs are formed by the physical compartmentalization of chromatin extrusion and insulation [7], which are regulated by multiple factors, including DNA methylation, histone modification, and DNA binding proteins such as CCCTC-binding factor (CTCF) and the cohesin complex. Epigenetics not only regulates gene transcription through promoters and enhancers but also genome stability. The deregulation of these processes can alter gene expression, which is often observed in cancers. These epigenetic marks are dynamic and reversible, and recent advances in the field have indicated that the enzymes involved in regulating these epigenetic marks are druggable targets.
In the following sections, we will discuss techniques involved in investigating high-order chromatin organization, factors regulating chromatin, and how mutations in these chromatin-remodeling complexes result in diseases such as cancer.

2. Techniques to Study Chromatin Interactions

In higher eukaryotes, chromosomes do not occupy random locations during the cell cycle. Still, they are arranged in specific patterns, where each chromosome occupies a defined and mutually exclusive region called chromosome territory (CT) [8]. Dekker et al. [9] developed an approach based on molecular biology to study chromatin organization in a 3D space called chromosome conformation capture (3C). 3C technique allows us to study the 3D organization of genomes and has revealed an intricate relationship between transcription and genome organization. As a result, a number of derivative methods have been developed [10], including chromosome conformation capture on a chip (4C), chromosome conformation capture carbon copy (5C), Hi-C, Micro-C, Capture-C, and the chromatin interaction analysis paired-end tag sequencing (ChIA-PET) allowing us to infer 3D chromosome organization.

2.1. Chromosome Conformation Capture (3C)

To determine genomic architecture, 3C quantifies the frequency of contacts between two distal selected genomic elements, such as enhancers and promoters, in a population of cells. In 3C, cells are formaldehyde cross-linked, chromatin is isolated and treated with a restriction enzyme and followed by the ligation of DNA fragments performed at low DNA concentrations, which aids intramolecular ligation of cross-linked chromatin segments (Figure 1a). At the point of cross-linking, a one-dimensional template representing a cellular 3D environment is captured. This template is used for semiquantitative or quantitative polymerase chain reaction (PCR) to interrogate short- and long-range chromatin interactions at the region of interest (Figure 1b) [11]. Despite 3C’s ability to visualize the genome at high resolution, there are still some limitations, including the need for genomic information that can be used to amplify regions of interest. In addition, 3C can only detect contact within a few hundred Kb.

2.2. Chromosome Conformation Capture-on-Chip (4C)

Using 4C, a region of interest can be identified at the genome-wide level. To analyze the chromatin organization, 3C methodology is used in the initial steps of the protocol (Figure 1a), followed by a ligation step after obtaining the 3C template (Figure 1c), resulting in small DNA circles. A reverse PCR strategy is then employed to amplify the contacting sequences using primers designed on the region of interest, also known as viewpoint. Using next-generation sequencing (NGS), one of the most common sequencing methods, viewpoint-contacting regions are identified. 4C can identify all genomic regions involved in the gene- or regulatory interaction [12].

2.3. Chromosome Conformation Capture Carbon Copy (5C)

Using 3C and 4C methods, it is not possible to infer genome organization. However, 5C provides information about multiple fragments establishing contacts in a large genomic region and can be investigated in parallel [13,14]. Based on the conventional 3C protocol, it uses highly multiplexed ligation-mediated amplification with pools of locus-specific primers to interrogate the 3C library, followed by sequencing or microarrays to quantify ligated pairs (Figure 1d). 5C enables the analysis of all interactions between two sets of genomic loci-one set recognized by the forward primers, the other by reverse primers. With thousands of 5C primers, 5C can be multiplexed very efficiently.

2.4. Hi-C

A series of “all versus all” methods emerged with the development of high-throughput sequencing technology. Hi-C has been extensively utilized to detect altered conformational changes in human diseases such as cancer. Hi-C method generates genome-wide contact maps without requiring specific primers [15,16]. Hi-C follows the same initial step as 3C to generate a template (Figure 1a). The sticky ends are filled in with biotin-labeled nucleotides and blunt-end ligated after restriction enzyme digestion. For high-throughput sequencing, only biotinylated junctions are selected by purifying them with streptavidin beads (Figure 1e).

2.5. Micro-C

Micro-C is a 3C-based method that is an enhanced and new variation of the Hi-C [17]. Micro-C was developed to elucidate the full spectrum of chromatin folding that can be observed from nucleosomes to chromosomes. Furthermore, Micro-C can provide more information on the boundaries of chromatin domains and identify new chromatin interactions and loops, providing higher resolution and higher signal-to-noise ratio. Micrococcal nuclease (MNase) is used in this protocol to digest the genome (Figure 1f), which results in the fragmentation of the chromatin into mononucleosomes. The protocol can overcome the resolution limitation of restriction enzyme-based approaches, thereby enabling a more accurate estimation of nucleosomes’ proximity to one another.

2.6. Chromatin Interaction Analysis Paired-End Tag Sequencing (ChIA-PET)

ChIA-PET is an innovative method for analyzing long-range chromatin interaction [18]. ChIA-PET is a combination of chromatin immunoprecipitation (ChIP) and 3C-type analysis. This method detects long-range chromatin interactions associated with a specific protein at a genome-wide level. ChIA-PET involves cross-linking DNA and proteins, fragmenting them by sonication, and capturing them with protein-specific antibodies. Biotin-labeled oligonucleotides with MmeI restriction sites are used to link the captured chromatin to the biotin-labeled oligonucleotides. After ligating the adjacent linkers, DNA fragments with paired-end tags (PETs) are obtained by digesting the linkers with MmeI restriction enzyme, followed by high-throughput sequencing using PETs (Figure 1g).

2.7. Capture-C

Capture-C is an improved version of 3C-sequencing and Hi-C and is a “many vs. all” approach. Capture-C is based on a standard 3C experiment, but the fragments ligated in the 3C are sonicated at the end of the experiment (Figure 1h). Capture-C uses biotinylated probes to capture target regions before they are sequenced downstream with high-throughput sequencing [19]. In Capture-C, the selected regions of interest undergo enrichment before sequencing. As a result, there is an increase in throughput and an improvement in genome resolution. This approach was later applied to Hi-C libraries; as a result, Capture Hi-C (CHi-C) was developed [20]. An improvement to Capture-C is called next-generation (NG) capture-C, which uses biotinylated single-stranded DNA (ssDNA) oligonucleotides in sequential “double capture” for indexed multiplexed 3C libraries [21].
Inside the nucleus, chromosomes are folded into a 3D structure. Genomic architecture has been studied using a variety of techniques. Among them, 3C-based techniques have become one of the most effective ways to understand how spatial genome organization is established and maintained. 3C technologies have helped our understanding of how changes in the genome result in a disease such as cancer [10].

3. Epigenetic Factors Involved in Regulating Chromatin Organization

This section will discuss proteins involved in modifying chromatin and how they regulate transcriptional landscape.

3.1. DNA Methylation

DNA methylation involves a covalent transfer of methyl group to cytosine in position 5, resulting in 5-methylcytosine (5 mC); this process is catalyzed by DNA methyltransferases (DNMTs) [22]. It is an essential epigenetic modification involved in regulating gene expression, cellular proliferation, differentiation, and stem cell maintenance [23]. The methylation pattern of DNA is established and maintained by DNMTs [22]. There are three major types of DNMTs; DNMT1, DNMT3a, and DNMT3b. DNMT3a and DNMT3b are involved in de novo DNA methylation, and DNMT1 maintains the methylation of DNA during DNA replication [24,25]. Most DNA methylation occurs on cytosines preceding guanines, or CpG sites, creating a transcriptionally repressive chromatin state. As a result of DNA methylation, a few transcription factors are unable to bind to DNA, and proteins that recognize methylated DNA through their methyl-CpG binding domain (MBD) are recruited to inhibit gene expression. In normal cells, CpG islands in gene promoters are usually unmethylated and are transcriptionally permissive, which keeps the chromatin open and enhances transcription. CpG islands-containing tumor suppressor genes at the promoters are methylated in tumor cells, turning euchromatin into heterochromatin, while in normal tissues, these CpG islands are typically unmethylated [26].

3.2. Histone Acetylation

Histone acetylation involves adding an acetyl group to the lysine residues of histones. All four core histone proteins can be acetylated in the nucleosome octamer at various lysine residues [27]. Acetylated histone tail facilitates nucleosome assembly and creates a transcriptionally active chromatin environment, allowing DNA sequences to be exposed and transcribed.
Histone acetylation levels are maintained by two enzymes, histone acetyltransferases (HATs) and histone deacetylases (HDACs) [28]. HATs utilize acetyl CoA as a cofactor to transfer an acetyl group to the ε-amino group of the lysine side chain of histone and non-histone proteins. HATs are classified into three major groups: Gcn5-related N-acetyltransferases (GNATs), MYSTs (MOZ, Ybf2, Sas2, TIP60), and orphans (CBP/EP300) [29]. Depending upon the acetylation of histone in the cytoplasm or the nucleus, HATs are categorized into two classes, type-A, and type-B. Nucleosomes and proteins involved in access to DNA are regulated by type-A HATs, whereas type-B HATs acetylate free histone, predominantly in the cytoplasm. HATs act as tumor suppressors and oncogenes, suggesting acetylation is critical in maintaining the balance. Deregulation of this balance often results in diseases such as cancers [30,31].
HDACs remove acetylation from histone tails and non-histone proteins. Human HDACs can be divided into two families based on their catalytic mechanisms. Zn2+ dependent enzymes (Class I, II, and IV) and NAD+ cofactor-dependent enzymes (Class III) [32]. There are 18 human HDACs which are further categorized into four classes. Class I (HDAC1, HDAC2, HDAC3 and HDAC8), Class IIa (HDAC4, HDAC5, HDAC7 and HDAC9), Class IIb (HDAC6 and HDAC10), Class III (Sirtuins 1–7), and Class IV (HDAC11) [33].

3.3. Histone Methylation

Histone methylation is one of the crucial post-transcriptional modifications. The methylation of histones does not affect histone-DNA binding directly but instead interferes with chromatin-binding factors’ interaction with DNA. To activate or repress gene expression, histone lysine residues can be mono-, di-, or tri-methylated (me1, me2, or me3) [34]. Methylation of histones has a diverse outcome; for example, methylation on H3K4, H3K36, and H3K79 are activating marks and are in an actively transcribed region, whereas H3K9, H3K27, and H4K20 are usually associated with silenced gene expression or condensed chromatin [35].
The methylation of histones is controlled by two families of enzymes: histone methyltransferases (HMTs) and histone demethylases (HDMs). HMTs transfer a methyl group to the lysine and arginine residues of the target protein, whereas HDMs can remove the methyl group [36]. This addition can be one, two, or three methyl groups of S-adenosyl-L-methionine transferred to lysine or arginine residues [37,38]. There are two families of histone demethylases: amino oxidase homolog lysine demethylase 1 (KDM1) and Jumonji C (JmjC) domain-containing histone demethylases [39]. The KDM1 family consists of two members, KDM1A, also known as lysine-specific demethylase 1 (LSD1), and KDM1B. KDM1A removes mono- and di-methylated lysine 4 or lysine 9 of histone H3 [40]. KDM1B is the other amine oxidase homolog that targets H3K4me1 and H3K4me2 [41]. A second group consists of demethylases containing the JmjC domain. Approximately 20 of the identified JmjC domain proteins demethylate lysines [42]. Unlike KDM1, these enzymes remove trimethylations [43].

3.4. Histone Phosphorylation

Phosphorylation of cellular proteins is one of the common post-translational modifications (PTM). Interestingly, histone proteins are also phosphorylated on serine, threonine, and tyrosine residues, resulting in changes in chromatin organization [37]. The phosphorylation of histone tails occurs predominantly at the N-terminal amino acid of histone tails. Phosphorylation of histone tails reduces the positive charge on the histones and weakens histone-DNA interaction, which leads to DNA access by protein complexes [44,45]. Several biological processes depend on core histone phosphorylation, including transcription, cell cycle regulation, DNA repair, and apoptosis [46]. Amongst the intracellular signaling pathways involved in chromatin remodeling, phosphorylation plays an important role in connecting the signaling pathways and chromatin [47]. For example, phosphorylation of histone H3 on serine 10 (H3S10) by Aurora-B dissociates heterochromatin protein 1 (HP1) from chromatin and prevents heterochromatin formation [48]. Some pathological events are associated with abnormal kinase activity, cancer being one of the most prominent [49,50,51].

3.5. Ubiquitination

Ubiquitin is an evolutionarily conserved protein among eukaryotes. Ubiquitin is a 76 amino acid residue polypeptide. It has a molecular mass of 8.5 kDa, and the ubiquitin exists as a free molecule or covalently conjugated with other proteins [52]. In cells, ubiquitin mostly functions as part of a multistep proteolytic pathway. Ubiquitylation of protein involves three main enzymes, ubiquitin-activating enzymes (E1), binding enzymes (E2), and ligases (E3) [53]. In cells, ubiquitylation is usually associated with the degradation of proteins. However, the addition of ubiquitin to histones acts as a regulatory mechanism [54,55]. In fact, in the nucleus, histones are the most abundant ubiquitinated proteins [56], and their ubiquitination is crucial for transcription, maintaining chromatin structure, and repairing DNA. Ubiquitination also plays a significant role in regulating tumor-suppressing and tumor-promoting pathways [57]. For example, TP53, a tumor suppressor gene that prevents instability of the genome, is targeted for degradation by ubiquitylation. By interacting with the N-terminus transactivation domain (TAC) of p53, MDM2 ubiquitinates p53 and triggers degradation by the ubiquitin-proteasomes system [58]. Another tumor suppressor gene, BRCA1, which is involved in DNA damage repair and the control of cell cycle checkpoints, is known to act as an E3 ligase, and mutations in E3 ubiquitin ligase activity result in tumorigenesis [59].

3.6. SUMOylation

Small ubiquitin-like modified proteins (SUMO) bear many similarities to ubiquitin proteins. SUMO is an evolutionarily conserved protein expressed in all eukaryotes [60]. Like ubiquitin, SUMO proteins are approximately 10 kDa in size. SUMOylation is an important post-translational protein modification that is conjugated to lysine residues of the target proteins. Four SUMO isoforms have been identified, which are SUMO1, SUMO2/3 and SUMO4 [61]. The SUMO proteins are initially translated as C-terminally stretched precursors, which are then cleaved by SUMO-specific proteases into proteins ending in a pair of glycine residues. The alpha-carboxylates at the C-terminus are used to attach other proteins covalently, and this process is known as sumoylation [62]. Target proteins of SUMO are often transcriptional coactivators or corepressors [63]. SUMO protein is primarily found in the nucleus, and diverse cellular processes are regulated by sumoylation, including transcription, DNA replication, cell-cycle progression, mitochondrial dynamics, DNA repair, and apoptosis [64,65].

3.7. Citrullination

Citrullination is another PTM catalyzed by peptidylarginine deiminase (PAD). Unlike epigenetic modifications, citrullination involves a Ca2+-driven enzymatic conversion of arginine residues to citrulline [66]. There are five isoenzymes, PAD1, PAD2, PAD3, PAD4, and PAD6 identified in humans. PADs have been shown to citrullinate histone H1, H2A, H3, and H4 [67,68,69]. It has been reported that calcium-activated PAD4 reduced the methylation of recombinant histones H3 and H4 and can affect chromatin structure by promoting decondensation in vivo [68]. Interestingly, citrullinated histones have been shown to interact with other modified histones, such as those that are methylated. In its dual enzyme activity, PAD4 converts arginine and methylated arginine into citrulline through deamination and demethylation, respectively [68,70], suggesting a crosstalk between methylation and citrullination. Citrullination of histones plays a significant role in transcription, pluripotency [69], development of embryos [71] and neutrophil extracellular traps (NET) formation [72].

3.8. ADP Ribosylation

ADP-ribosylation is a covalent post-translational modification of proteins catalyzed by ADP-ribosyltransferases. ADP-ribosylation is a process in which mono or poly-ADP-ribose molecules are transferred from the cofactor nicotinamide adenine dinucleotide (NAD) to the target protein by a process known as mono-ADP-ribosylation (MARylation) or poly-ADP-ribosylation (PARylation). In addition to the core histones, H1 linker histones are ADP-ribosylated. ADP-ribosylation occurs in glutamic acid, aspartic acid, lysine, arginine, and serine [73,74,75]. A majority of histone ribosylation events in mammals are catalyzed by PARP1, which is supported by PARP2/3/7/10 [76]. It plays an essential role in cell cycle regulation, DNA damage response [77], replication, and transcription [78].

3.9. Histone Glycosylation

Proteins and lipids undergo glycosylation, the process in which monosaccharides are attached to proteins and lipids to modify their properties. Histone glycosylation is a post-translational modification. In the human genome, the modification of the hydroxyl group of a serine or threonine residue by monosaccharides on the core histones proteins H2A, H2B, H3, and H4 results in conjugation to O-linked N-acetylglucosamine (O-GlcNAc). It has been reported that various core histone proteins have O-GlcNAc modifications with varying functions, for example, H2AT101, H2BS36, H3S10, and H4S47, as well as H3T32 [79,80]. Glycosylation has been shown to regulate the cell cycle, transcriptional activation, and chromatin dynamics [81,82,83,84]. A wide variety of cancer types have been found to have aberrant O-GlcNAcylation, and studies suggest that O-GlcNAcylation may play a regulatory role in cancer development [85].

3.10. Proline Isomerization

In amino acids, functional groups on carbon chains are positioned according to their isomeric form: cis and trans. Protein folding is affected by the changes in proline residues between the cis and trans forms. Peptidylprolyl isomerases (PPIases), which operate by switching between conformations, are enzymes that facilitate the isomerization process. Initially, Tony Kouzarides’ lab reported proline isomerization on the tail of histone H3 [86]. Histone H3P38’s trans-proline conformation is essential for methylating lysine 36 of histone H3 (H3K36). On the other hand, cis-isomerization is related to lower H3K36 methylation and decreased transcription elongation due to cis-isomerization. The acetylation of H3K14 facilitates H3P16 trans-isomerization. As a result, H3K4me3 levels are reduced in vivo, indicating that H3P16 trans-isomerization may be involved in the repression of transcription [87].

4. Epimutations, TADs and Cancer

4.1. Writes, Readers, and Erasers

Numerous cancers have been linked to epigenetic modifications, such as DNA methylation and histone modifications [88,89]. These enzymes can further be classified as writers (DNA methyltransferases [DNMTs], histone acetyltransferases [HATs] and histone methyltransferase [HMTs], etc.), erasers (histone demethylases [HDMs], and histone deacetylases [HDACs]), readers (bromodomain and chromodomain proteins that are known to read acetylated and methylated residues, respectively) (Table 1 and Figure 2), and insulators (CTCF and the cohesin complex).

4.2. Loops, TADs, and Insulators

The DNA is wrapped around histones to form chromatin which further undergoes high-order organization resulting in loops and TADs (Figure 3). This high-order organization is facilitated by two key protein complexes, CTCF and cohesin. CTCF is a highly conserved, ubiquitously expressed, and essential protein in eukaryotes [93]. CTCF is an 82-kDa protein that comprises three separate regions, an 11-zinc fingers region with a central zinc-finger domain that binds to the DNA flanked by large C- and N-terminal domains [94]. CTCF acts as both a transcriptional activator and repressor by interacting with cofactors involved in transcriptional processes [95]. CTCF functions as an insulator when positioned between the promoter and enhancer through its ability to form loops and domains [96]. Cohesin is a ring-shaped protein complex composed of four core subunits encoded by the genes Smc1, Smc3, Rad21, and STAG1/2 [97]. Cohesin and CTCF colocalize in the mammalian genomes to form high-order chromatin structures [98]. One of the functions of cohesin is to fold chromatin, create CTCF loops, and bring cis-acting elements, such as enhancers, to gene promoters in proximity [99]. By doing so, the cohesin protein complex plays an important role in cell cycle regulation, DNA repair, chromatin organization, and gene expression regulation [100].
TADs are separated by insulators that maintain higher-order chromatin structures. Changes in their expression level or mutation in the binding site result in an abnormal scenario. Indeed, mutations in CTCF and cohesin have been identified in breast, prostate, and uterine cancers [101]. In prostate cancer, loss of CTCF results in hypermethylation at CTCF binding sites. The cohesin subunits (Smc1, Smc3, Rad21, and STAG1/2) are also frequently mutated in cancer. For example, RAD21 somatic mutation and amplification are found in both solid and hematopoietic tumors. Somatic mutation is mainly found in hematopoietic malignancy, while overexpression is associated with solid tumors [102]. Other components of the cohesin complex are also known to be mutated in cancers. For example, SMC1A is frequently mutated in colorectal cancer, and its variants have been detected in leukemia and bladder cancer [103]. Studies have shown that STAG2 acts as a tumor suppressor gene. STAG2 mutations are mostly nonsense, frameshift, and splice mutants found in melanoma, acute myeloid leukemia (AML), Ewing’s sarcoma, and myelodysplastic syndrome (MDS) [104].

4.3. Epigenetic Inhibitors

Epigenetic mechanisms account for one-third to one-half of all genetic alterations in cancer [105]. Deregulation of their enzyme activity can be targeted by small molecules such as DNMT inhibitors (DNMTi), HDAC inhibitors (HDACi), DOT1L inhibitors, EZH2 inhibitors, and BET inhibitors [106,107] (Table 2). In this section, we will discuss a few of the inhibitors showing promising preclinical and clinical data for cancer treatment.

4.3.1. DNA Modification Inhibitors

As discussed earlier, malignant transformation is frequently associated with hypermethylation of CpG-rich regions leading to the silencing of tumor suppressor genes, and this can be prevented by exposing cancer cells to DNMTi. DNMTi are potent therapeutic agents for cancer treatment as they reverse the hypermethylation of DNA in tumor cells. DNMTi have been classified into two types; nucleoside analogs and non-nucleoside analogs. Nucleoside analogs are made from modified cytosine rings with nitrogen at position 5 (instead of carbon), and these modified analogs are incorporated into newly synthesized DNA. In contrast, non-nucleoside DNMTi are small molecules directly targeting the catalytic sites instead of being incorporated into DNA [163]. Azacytidine (5-azacytidine) and decitabine (5-aza-2′-deoxycytidine) are the two well-studied nucleoside analogs and potent DNA hypomethylating agents. Mechanistically, azacytidine and decitabine form an irreversible complex with DNMTs, causing degradation of the complex [164]. The FDA has approved these compounds for treating AML and MDS [107] as they promote differentiation and reexpression of inactivated genes. Decitabine and azacytidine are also being studied as therapeutic options in multiple solid cancers [165,166]. DNA methylation is inhibited in a dose-dependent manner by decitabine. Interestingly, when used at low doses, decitabine can reactivate silenced genes, but at high doses, it becomes cytotoxic, and at all doses, myelosuppression is a prominent side effect [167]. Second-generation DNMTi, such as SGI110, were demonstrated to be more stable and less toxic than first-generation DNMTi [168,169].
Demethylation of DNA is mediated by the ten-eleven translocation (TET) protein family (TET1/2/3), which converts 5-methylcytosine (5mC) into 5-hydroxymethylcytosine (5hmC) during DNA demethylation [170]. Isocitrate dehydrogenase (IDH) is responsible for generating the metabolite (2-oxoglutarate), which is necessary for the activity of the TET enzyme. AG-120 (ivosidenib) and AG-221 (enasidenib) are IDH inhibitors that ablate IDH1/2-mediated conversion of α-ketoglutarate to 2-hydroxygluterate and accumulation of 2-hydroxygluterate inhibits TET activity [171]. It has been demonstrated that targeting DNMTs not only activates tumor suppressor genes but also restores optimal methylation levels, which then modulates the binding of proteins such as CTCF and cohesin. For instance, in breast and prostate cancer, loss of CTCF copy number leads to hypermethylation of DNA at surrounding CTCF binding sites [172]. Inhibition by azacytidine reverses the change in gene expression after the loss of CTCF sites [173].

4.3.2. Histone Modification Inhibitors

A number of epi-drug that inhibit histone modification have been developed and used for cancer treatment. For example, HDACi such as vorinostat (SAHA), belinostat (PXD-101), panobinostat (LBH589), and romidepsin (FK228) have shown promising results in treating cancer [174,175]. For example, exposure to vorinostat induced cell cycle arrest in refractory/relapsed B-cell lymphoma and is due to an increase of p21 and the acetylation of histone H3 [176,177]. Similarly, another HDACi, belinostat, has been approved for use in relapsed or refractory peripheral T-cell lymphoma (PTCL), panobinostat has been used for multiple myeloma (MM) [138,178], and romidepsin has been approved for the treatment of cutaneous T-cell lymphoma (CTCL) [179]. Like HDACi, HAT inhibitors (HATi) have also shown promising results. For example, the small molecule C646 selectively inhibits p300/CBP activity [180,181] and induces cell cycle arrest and apoptosis. In addition to HATi and HDACi, HMT inhibitors (HMTi) have also been developed as anticancer drugs. Pharmacological inhibition of DOT1L by EPZ-5676 reduces H3K79me methylation and decreases proliferation. Loss of cohesin resulted in increased self-renewal and increased HoxA9 expression, which is reversed by DOT1L inhibition (pinometostat) [182]. Inhibition of DOT1L in cohesin-depleted cells also shows a decrease in H3K79me2 and a concomitant increase in H3K27me3 [182]. Deregulation in another HMT, LSD1, promotes acute lymphoblastic leukemia (ALL) and AML [183,184]. Several LSD1 inhibitors (TCP, ORY-1001, GSK-2879552, and ORY-2001) are currently being evaluated for cancer treatment and have been shown to inhibit proliferation, differentiation, and invasion in both in vitro and in vivo studies [185].

5. Future Perspective

The human genome is organized in a complex 3D structure. As epigenetic mechanisms play a significant role in regulating tumorigenesis, understanding how DNA and histone modifications are involved in maintaining genome organization will be an exciting avenue in the future. Furthermore, since epigenetic pathways are reversible, targeting the enzymes deregulated in cancer can restore this balance, offering promising avenues for anticancer treatment. Clustered regularly interspaced short palindromic repeats (CRISPR)—CRISPR-associated endonuclease (Cas9) has emerged as a powerful genome-editing technique that allows specific changes to be made to the genome with high efficiency and specificity. CRISPR-Cas9 genome editing is being explored extensively for cancer therapy and has been used to target epigenetic modification. Several epigenetic therapies reverse gene expression, such as epi-drugs and combinatorial treatment. However, a significant obstacle to their clinical application is that epi-drugs might activate off-target genes in normal cells that cause mutagenesis. Since the CRISPR-Cas9 system enables specificity, modified versions such as CRISPR activation (CRISPRa) and CRISPR inactivation (CRSPRi) might be effective tools for cancer treatment.

Author Contributions

Conceptualization, S.J. and A.M.; writing—original draft preparation, A.M.; writing—review and editing, S.J.; supervision, S.J.; funding acquisition, S.J. All authors have read and agreed to the published version of the manuscript.

Funding

Jha lab is funded by a start-up from Oklahoma State University College of Veterinary Medicine, Department of Physiological Sciences to S.J.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

3Cchromosome conformation capture
3Dthree-dimensional
4Cchromosome conformation capture on a chip
53BP1tumor protein p53 binding protein 1
5Cchromosome conformation capture carbon copy
5mC5-methylcytosine
AMLacute myeloid leukemia
APLFaprataxin and PNKP like factor
AURKBAurora kinase B
BARD1BRCA1 associated RING domain 1
BAZ1Bbromodomain adjacent to zinc finger domain 1B
BETbromodomain and extra-terminal
BIRBaculovirus IAP repeat
BRCA1breast cancer gene 1
BRCC36BRCA1/BRCA2-containing complex subunit 3
BRCTBRCA1 C terminus
BRD2/3/4/Tbromodomain containing 2/3/4/T
BRPF1BRD and PHD finger-containing protein 1
CREBBP/CBPCREB-binding protein
CHFRcheckpoint with forkhead and RING finger domains
ChIA-PETchromatin interaction analysis paired-end tag sequencing
ChIPChromatin immunoprecipitation
CHi-CCapture Hi-C
CLLchronic lymphocytic leukemia
CRISPRclustered regularly interspaced short palindromic repeats
Cas9CRISPR-associated endonuclease
CRISPRaCRISPR activation
CRISPRiCRISPR inactivation
CTchromosome territory
CTCFCCCTC-binding factor
CTCLcutaneous T-cell lymphoma
DLBCLdiffuse large B-cell lymphoma
DNMT1/3a/3bDNA methyltransferase 1/3a/3b
DNMTiDNA methylatransferase inhibitor
DNMTsDNA methyltransferases
DOT1LDOT1 like histone lysine methyltransferase
EP300E1A binding protein p300
EYA1/3eyes absent homolog 1/3
EZH1/2enhancer of Zeste homolog 1/2
G9a/EHMT2euchromatic histone lysine methyltransferase 2
GNATsGcn5-related N-acetyltransferases
GLP/EHMT1euchromatic histone lysine methyltransferase 1
HAThistone acetyltransferase
HATihistone acetylatransferase inhibitor
HATshistone acetyltransferases
HDAChistone deacetylase
HDACihistone deacetylase inhibitor
HDACshistone deacetylases
HDMhistone demethylase
HDMshistone demethylases
HMThistone methyltransferase
HMTihistone methyltransferase inhibitor
HMTshistone methyltransferases
HoxA9Homeobox A9
HP1heterochromatin protein 1
IDHisocitrate dehydrogenase
JmjCJumonji C
KAT5lysine acetyltransferase 5
KDM1/1A/1B/2A/2B/3A/3B/4/5Blysine demethylase 1/1A/1B/2A/2B/3A/3B/4/5B
KilobasesKb
KilodaltonskDa
KMT4/7lysine N-methyltransferase4/7
LSD1lysine-specific demethylase 1
MBDmethyl-CpG binding domain
MBTmalignant brain tumor
MDSmyelodysplastic syndrome
MegabasesMb
MMMultiple myeloma
MORFmonocytic leukemia zinc finger protein-related factor
MRG15MORF4-related gene on chromosome 15
MYSTMOZ, Ybf2,Sas2, TIP60
NADnicotinamide adenine dinucleotide
NBS1Nijmegen breakage syndrome 1
NETneutrophil extracellular traps
NGnext-generation
NGSnext-generation sequencing
NSCLCnon-small cell lung cancer
NSD1-3nuclear receptor binding SET domain proteins 1-3
O-GlcNAcO-linked N-acetylglucosamine
OTUB1/2OTU domain-containing ubiquitin aldehyde-binding protein 1/2
p21Cyclin dependent kinase inhibitor 1A
p300E1A binding protein p300
PADpeptidyl arginine deiminase
PARGpoly(ADP-ribose) glycohydrolase
PARP1poly(ADP-ribose) polymerase 1
PBZPAR-binding zinc finger
PCRpolymerase chain reaction
PETspaired-end tags
PHFplant homeodomain (PHD) finger
PHF1/19PHD finger protein 1/19
PP1protein phosphatase 1
PP2protein phosphatase 2
PPARgammaperoxisome proliferator activated receptor gamma
PPIasespeptidylprolyl isomerases
PPP1/2CAprotein phosphatase 1/2 catalytic subunit alpha
PRMT1/2/4/5/6/7protein arginine methyltransferase 1/2/4/5/6/7
PTCLperipheral T-cell lymphoma
PWWPPro-Trp-Trp-Pro
Rad21RAD21 cohesin complex component
RNF146RING finger protein 146
SCCsquamous cell carcinoma
ssDNAsingle-stranded DNA
SIRTSirtuin 1
SMC1/3structural maintenance of chromosomes 1/3
STAG1/3stromal antigen 1/3
SUMOsmall ubiquitin-like modified proteins
SUV39H1/2suppressor of variegation 3-9 homolog 1/2
TACtransactivation domain
TADtopologically associated domain
TAF3TATA-Box Binding Protein Associated Factor 3
TBDTo be determined
TDRD3Tudor domain containing proteins 3
TET1/2ten-eleven translocation 1/2
TP53tumor protein 53
USP3/16/26/44ubiquitin specific peptidase 3/16/26/44
WSTFWilliams Syndrome transcription factor
XRCC1X-ray repair cross complementing 1

References

  1. Felsenfeld, G.; Groudine, M. Controlling the double helix. Nature 2003, 421, 448–453. [Google Scholar] [CrossRef] [Green Version]
  2. Maeshima, K.; Ide, S.; Babokhov, M. Dynamic chromatin organization without the 30-nm fiber. Curr. Opin. Cell Biol. 2019, 58, 95–104. [Google Scholar] [CrossRef]
  3. Ou, H.D.; Phan, S.; Deerinck, T.J.; Thor, A.; Ellisman, M.H.; O’Shea, C.C. ChromEMT: Visualizing 3D chromatin structure and compaction in interphase and mitotic cells. Science 2017, 357, eaag0025. [Google Scholar] [CrossRef] [Green Version]
  4. Dekker, J.; Misteli, T. Long-Range Chromatin Interactions. Cold Spring Harb. Perspect. Biol. 2015, 7, a019356. [Google Scholar] [CrossRef] [Green Version]
  5. Carter, D.; Chakalova, L.; Osborne, C.S.; Dai, Y.-f.; Fraser, P. Long-range chromatin regulatory interactions in vivo. Nat. Genet. 2002, 32, 623–626. [Google Scholar] [CrossRef]
  6. Tolhuis, B.; Palstra, R.J.; Splinter, E.; Grosveld, F.; de Laat, W. Looping and interaction between hypersensitive sites in the active beta-globin locus. Mol. Cell 2002, 10, 1453–1465. [Google Scholar] [CrossRef]
  7. Szabo, Q.; Bantignies, F.; Cavalli, G. Principles of genome folding into topologically associating domains. Sci. Adv. 2019, 5, eaaw1668. [Google Scholar] [CrossRef] [Green Version]
  8. Cremer, T.; Cremer, M. Chromosome territories. Cold Spring Harb. Perspect. Biol. 2010, 2, a003889. [Google Scholar] [CrossRef] [Green Version]
  9. Dekker, J.; Rippe, K.; Dekker, M.; Kleckner, N. Capturing chromosome conformation. Science 2002, 295, 1306–1311. [Google Scholar] [CrossRef] [Green Version]
  10. Denker, A.; de Laat, W. The second decade of 3C technologies: Detailed insights into nuclear organization. Genes Dev. 2016, 30, 1357–1382. [Google Scholar] [CrossRef]
  11. Hagège, H.; Klous, P.; Braem, C.; Splinter, E.; Dekker, J.; Cathala, G.; de Laat, W.; Forné, T. Quantitative analysis of chromosome conformation capture assays (3C-qPCR). Nat. Protoc. 2007, 2, 1722–1733. [Google Scholar] [CrossRef] [Green Version]
  12. Simonis, M.; Klous, P.; Splinter, E.; Moshkin, Y.; Willemsen, R.; de Wit, E.; van Steensel, B.; de Laat, W. Nuclear organization of active and inactive chromatin domains uncovered by chromosome conformation capture-on-chip (4C). Nat. Genet. 2006, 38, 1348–1354. [Google Scholar] [CrossRef]
  13. Dostie, J.; Richmond, T.A.; Arnaout, R.A.; Selzer, R.R.; Lee, W.L.; Honan, T.A.; Rubio, E.D.; Krumm, A.; Lamb, J.; Nusbaum, C.; et al. Chromosome Conformation Capture Carbon Copy (5C): A massively parallel solution for mapping interactions between genomic elements. Genome Res. 2006, 16, 1299–1309. [Google Scholar] [CrossRef] [Green Version]
  14. Dostie, J.; Dekker, J. Mapping networks of physical interactions between genomic elements using 5C technology. Nat. Protoc. 2007, 2, 988–1002. [Google Scholar] [CrossRef]
  15. Lieberman-Aiden, E.; van Berkum, N.L.; Williams, L.; Imakaev, M.; Ragoczy, T.; Telling, A.; Amit, I.; Lajoie, B.R.; Sabo, P.J.; Dorschner, M.O.; et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 2009, 326, 289–293. [Google Scholar] [CrossRef] [Green Version]
  16. van Berkum, N.L.; Lieberman-Aiden, E.; Williams, L.; Imakaev, M.; Gnirke, A.; Mirny, L.A.; Dekker, J.; Lander, E.S. Hi-C: A method to study the three-dimensional architecture of genomes. J. Vis. Exp. 2010, 39, e1869. [Google Scholar] [CrossRef] [Green Version]
  17. Hsieh, T.H.; Weiner, A.; Lajoie, B.; Dekker, J.; Friedman, N.; Rando, O.J. Mapping Nucleosome Resolution Chromosome Folding in Yeast by Micro-C. Cell 2015, 162, 108–119. [Google Scholar] [CrossRef] [Green Version]
  18. Li, X.; Luo, O.J.; Wang, P.; Zheng, M.; Wang, D.; Piecuch, E.; Zhu, J.J.; Tian, S.Z.; Tang, Z.; Li, G.; et al. Long-read ChIA-PET for base-pair-resolution mapping of haplotype-specific chromatin interactions. Nat. Protoc. 2017, 12, 899–915. [Google Scholar] [CrossRef] [Green Version]
  19. Hughes, J.R.; Roberts, N.; McGowan, S.; Hay, D.; Giannoulatou, E.; Lynch, M.; De Gobbi, M.; Taylor, S.; Gibbons, R.; Higgs, D.R. Analysis of hundreds of cis-regulatory landscapes at high resolution in a single, high-throughput experiment. Nat. Genet. 2014, 46, 205–212. [Google Scholar] [CrossRef] [PubMed]
  20. Mifsud, B.; Tavares-Cadete, F.; Young, A.N.; Sugar, R.; Schoenfelder, S.; Ferreira, L.; Wingett, S.W.; Andrews, S.; Grey, W.; Ewels, P.A.; et al. Mapping long-range promoter contacts in human cells with high-resolution capture Hi-C. Nat. Genet. 2015, 47, 598–606. [Google Scholar] [CrossRef]
  21. Davies, J.O.J.; Telenius, J.M.; McGowan, S.J.; Roberts, N.A.; Taylor, S.; Higgs, D.R.; Hughes, J.R. Multiplexed analysis of chromosome conformation at vastly improved sensitivity. Nat. Methods 2016, 13, 74–80. [Google Scholar] [CrossRef] [PubMed]
  22. Moore, L.D.; Le, T.; Fan, G. DNA Methylation and Its Basic Function. Neuropsychopharmacology 2013, 38, 23–38. [Google Scholar] [CrossRef] [Green Version]
  23. Gereige, L.-M.; Mikkola, H.K.A. DNA methylation is a guardian of stem cell self-renewal and multipotency. Nat. Genet. 2009, 41, 1164–1166. [Google Scholar] [CrossRef]
  24. Okano, M.; Bell, D.W.; Haber, D.A.; Li, E. DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 1999, 99, 247–257. [Google Scholar] [CrossRef] [Green Version]
  25. Pradhan, S.; Bacolla, A.; Wells, R.D.; Roberts, R.J. Recombinant human DNA (cytosine-5) methyltransferase. I. Expression, purification, and comparison of de novo and maintenance methylation. J. Biol. Chem. 1999, 274, 33002–33010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Baylin, S.B.; Jones, P.A. A decade of exploring the cancer epigenome-biological and translational implications. Nat. Rev. Cancer 2011, 11, 726–734. [Google Scholar] [CrossRef]
  27. Zentner, G.E.; Henikoff, S. Regulation of nucleosome dynamics by histone modifications. Nat. Struct. Mol. Biol. 2013, 20, 259–266. [Google Scholar] [CrossRef] [PubMed]
  28. Grunstein, M. Histone acetylation in chromatin structure and transcription. Nature 1997, 389, 349–352. [Google Scholar] [CrossRef]
  29. Marmorstein, R.; Zhou, M.M. Writers and readers of histone acetylation: Structure, mechanism, and inhibition. Cold Spring Harb. Perspect. Biol. 2014, 6, a018762. [Google Scholar] [CrossRef] [Green Version]
  30. Di Cerbo, V.; Schneider, R. Cancers with wrong HATs: The impact of acetylation. Brief. Funct. Genom. 2013, 12, 231–243. [Google Scholar] [CrossRef]
  31. Cang, S.; Feng, J.; Konno, S.; Han, L.; Liu, K.; Sharma, S.C.; Choudhury, M.; Chiao, J.W. Deficient histone acetylation and excessive deacetylase activity as epigenomic marks of prostate cancer cells. Int. J. Oncol. 2009, 35, 1417–1422. [Google Scholar] [CrossRef] [Green Version]
  32. Seto, E.; Yoshida, M. Erasers of histone acetylation: The histone deacetylase enzymes. Cold Spring Harb. Perspect. Biol. 2014, 6, a018713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Porter, N.J.; Christianson, D.W. Structure, mechanism, and inhibition of the zinc-dependent histone deacetylases. Curr. Opin. Struct. Biol. 2019, 59, 9–18. [Google Scholar] [CrossRef] [PubMed]
  34. Hyun, K.; Jeon, J.; Park, K.; Kim, J. Writing, erasing and reading histone lysine methylations. Exp. Mol. Med. 2017, 49, e324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Greer, E.L.; Shi, Y. Histone methylation: A dynamic mark in health, disease and inheritance. Nat. Rev. Genet. 2012, 13, 343–357. [Google Scholar] [CrossRef] [Green Version]
  36. Black, J.C.; Van Rechem, C.; Whetstine, J.R. Histone lysine methylation dynamics: Establishment, regulation, and biological impact. Mol. Cell 2012, 48, 491–507. [Google Scholar] [CrossRef] [Green Version]
  37. Bannister, A.J.; Kouzarides, T. Regulation of chromatin by histone modifications. Cell Res. 2011, 21, 381–395. [Google Scholar] [CrossRef] [PubMed]
  38. Wolf, S.S. The protein arginine methyltransferase family: An update about function, new perspectives and the physiological role in humans. Cell. Mol. Life Sci. 2009, 66, 2109–2121. [Google Scholar] [CrossRef] [PubMed]
  39. Shi, Y.; Whetstine, J.R. Dynamic regulation of histone lysine methylation by demethylases. Mol. Cell 2007, 25, 1–14. [Google Scholar] [CrossRef]
  40. Shi, Y.; Lan, F.; Matson, C.; Mulligan, P.; Whetstine, J.R.; Cole, P.A.; Casero, R.A.; Shi, Y. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 2004, 119, 941–953. [Google Scholar] [CrossRef]
  41. Ciccone, D.N.; Su, H.; Hevi, S.; Gay, F.; Lei, H.; Bajko, J.; Xu, G.; Li, E.; Chen, T. KDM1B is a histone H3K4 demethylase required to establish maternal genomic imprints. Nature 2009, 461, 415–418. [Google Scholar] [CrossRef]
  42. Højfeldt, J.W.; Agger, K.; Helin, K. Histone lysine demethylases as targets for anticancer therapy. Nat. Rev. Drug Discov. 2013, 12, 917–930. [Google Scholar] [CrossRef] [PubMed]
  43. Cloos, P.A.; Christensen, J.; Agger, K.; Helin, K. Erasing the methyl mark: Histone demethylases at the center of cellular differentiation and disease. Genes Dev. 2008, 22, 1115–1140. [Google Scholar] [CrossRef] [Green Version]
  44. Banerjee, T.; Chakravarti, D. A peek into the complex realm of histone phosphorylation. Mol. Cell. Biol. 2011, 31, 4858–4873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. North, J.A.; Javaid, S.; Ferdinand, M.B.; Chatterjee, N.; Picking, J.W.; Shoffner, M.; Nakkula, R.J.; Bartholomew, B.; Ottesen, J.J.; Fishel, R.; et al. Phosphorylation of histone H3(T118) alters nucleosome dynamics and remodeling. Nucleic Acids Res. 2011, 39, 6465–6474. [Google Scholar] [CrossRef] [PubMed]
  46. Loury, R.; Sassone-Corsi, P. Histone phosphorylation: How to proceed. Methods 2003, 31, 40–48. [Google Scholar] [CrossRef]
  47. Cohen, P. The structure and regulation of protein phosphatases. Annu. Rev. Biochem. 1989, 58, 453–508. [Google Scholar] [CrossRef]
  48. Hirota, T.; Lipp, J.J.; Toh, B.-H.; Peters, J.-M. Histone H3 serine 10 phosphorylation by Aurora B causes HP1 dissociation from heterochromatin. Nature 2005, 438, 1176–1180. [Google Scholar] [CrossRef]
  49. Cohen, P. The origins of protein phosphorylation. Nat. Cell. Biol. 2002, 4, E127–E130. [Google Scholar] [CrossRef]
  50. Ubersax, J.A.; Ferrell, J.E., Jr. Mechanisms of specificity in protein phosphorylation. Nat. Rev. Mol. Cell Biol. 2007, 8, 530–541. [Google Scholar] [CrossRef]
  51. Cohen, P. Protein kinases--the major drug targets of the twenty-first century? Nat. Rev. Drug Discov. 2002, 1, 309–315. [Google Scholar] [CrossRef] [PubMed]
  52. Callis, J. The ubiquitination machinery of the ubiquitin system. Arab. Book/Am. Soc. Plant Biol. 2014, 12, e0174. [Google Scholar] [CrossRef] [Green Version]
  53. Kimura, Y.; Tanaka, K. Regulatory mechanisms involved in the control of ubiquitin homeostasis. J. Biochem. 2010, 147, 793–798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Jason, L.J.; Moore, S.C.; Lewis, J.D.; Lindsey, G.; Ausió, J. Histone ubiquitination: A tagging tail unfolds? Bioessays 2002, 24, 166–174. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, H.; Zhai, L.; Xu, J.; Joo, H.Y.; Jackson, S.; Erdjument-Bromage, H.; Tempst, P.; Xiong, Y.; Zhang, Y. Histone H3 and H4 ubiquitylation by the CUL4-DDB-ROC1 ubiquitin ligase facilitates cellular response to DNA damage. Mol. Cell 2006, 22, 383–394. [Google Scholar] [CrossRef]
  56. Vaughan, R.M.; Kupai, A.; Rothbart, S.B. Chromatin Regulation through Ubiquitin and Ubiquitin-like Histone Modifications. Trends Biochem. Sci. 2021, 46, 258–269. [Google Scholar] [CrossRef]
  57. Hoeller, D.; Dikic, I. Targeting the ubiquitin system in cancer therapy. Nature 2009, 458, 438–444. [Google Scholar] [CrossRef]
  58. Pan, Y.; Chen, J. MDM2 promotes ubiquitination and degradation of MDMX. Mol. Cell. Biol. 2003, 23, 5113–5121. [Google Scholar] [CrossRef] [Green Version]
  59. Witus, S.R.; Stewart, M.D.; Klevit, R.E. The BRCA1/BARD1 ubiquitin ligase and its substrates. Biochem. J. 2021, 478, 3467–3483. [Google Scholar] [CrossRef]
  60. Johnson, E.S. Protein modification by SUMO. Annu. Rev. Biochem. 2004, 73, 355–382. [Google Scholar] [CrossRef]
  61. Han, Z.J.; Feng, Y.H.; Gu, B.H.; Li, Y.M.; Chen, H. The post-translational modification, SUMOylation, and cancer (Review). Int. J. Oncol. 2018, 52, 1081–1094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Johnson, P.R.; Hochstrasser, M. SUMO-1: Ubiquitin gains weight. Trends Cell Biol. 1997, 7, 408–413. [Google Scholar] [CrossRef] [PubMed]
  63. Rosonina, E.; Akhter, A.; Dou, Y.; Babu, J.; Sri Theivakadadcham, V.S. Regulation of transcription factors by sumoylation. Transcription 2017, 8, 220–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Flotho, A.; Melchior, F. Sumoylation: A regulatory protein modification in health and disease. Annu. Rev. Biochem. 2013, 82, 357–385. [Google Scholar] [CrossRef]
  65. Ryu, H.Y.; Ahn, S.H.; Hochstrasser, M. SUMO and cellular adaptive mechanisms. Exp. Mol. Med. 2020, 52, 931–939. [Google Scholar] [CrossRef]
  66. Rogers, G.E.; Simmonds, D.H. Content of citrulline and other amino-acids in a protein of hair follicles. Nature 1958, 182, 186–187. [Google Scholar] [CrossRef]
  67. Saiki, M.; Watase, M.; Matsubayashi, H.; Hidaka, Y. Recognition of the N-terminal histone H2A and H3 peptides by peptidylarginine deiminase IV. Protein Pept. Lett. 2009, 16, 1012–1016. [Google Scholar] [CrossRef]
  68. Wang, Y.; Wysocka, J.; Sayegh, J.; Lee, Y.H.; Perlin, J.R.; Leonelli, L.; Sonbuchner, L.S.; McDonald, C.H.; Cook, R.G.; Dou, Y.; et al. Human PAD4 regulates histone arginine methylation levels via demethylimination. Science 2004, 306, 279–283. [Google Scholar] [CrossRef] [PubMed]
  69. Christophorou, M.A.; Castelo-Branco, G.; Halley-Stott, R.P.; Oliveira, C.S.; Loos, R.; Radzisheuskaya, A.; Mowen, K.A.; Bertone, P.; Silva, J.C.R.; Zernicka-Goetz, M.; et al. Citrullination regulates pluripotency and histone H1 binding to chromatin. Nature 2014, 507, 104–108. [Google Scholar] [CrossRef] [Green Version]
  70. Cuthbert, G.L.; Daujat, S.; Snowden, A.W.; Erdjument-Bromage, H.; Hagiwara, T.; Yamada, M.; Schneider, R.; Gregory, P.D.; Tempst, P.; Bannister, A.J.; et al. Histone deimination antagonizes arginine methylation. Cell 2004, 118, 545–553. [Google Scholar] [CrossRef]
  71. Zhang, X.; Liu, X.; Zhang, M.; Li, T.; Muth, A.; Thompson, P.R.; Coonrod, S.A.; Zhang, X. Peptidylarginine deiminase 1-catalyzed histone citrullination is essential for early embryo development. Sci. Rep. 2016, 6, 38727. [Google Scholar] [CrossRef] [Green Version]
  72. Maronek, M.; Gardlik, R. The Citrullination-Neutrophil Extracellular Trap Axis in Chronic Diseases. J. Innate Immun. 2022, 14, 393–417. [Google Scholar] [CrossRef]
  73. Leidecker, O.; Bonfiglio, J.J.; Colby, T.; Zhang, Q.; Atanassov, I.; Zaja, R.; Palazzo, L.; Stockum, A.; Ahel, I.; Matic, I. Serine is a new target residue for endogenous ADP-ribosylation on histones. Nat. Chem. Biol. 2016, 12, 998–1000. [Google Scholar] [CrossRef] [Green Version]
  74. Laing, S.; Unger, M.; Koch-Nolte, F.; Haag, F. ADP-ribosylation of arginine. Amino Acids 2011, 41, 257–269. [Google Scholar] [CrossRef] [Green Version]
  75. Zhang, Y.; Wang, J.; Ding, M.; Yu, Y. Site-specific characterization of the Asp- and Glu-ADP-ribosylated proteome. Nat. Methods 2013, 10, 981–984. [Google Scholar] [CrossRef]
  76. Liu, C.; Vyas, A.; Kassab, M.A.; Singh, A.K.; Yu, X. The role of poly ADP-ribosylation in the first wave of DNA damage response. Nucleic Acids Res. 2017, 45, 8129–8141. [Google Scholar] [CrossRef]
  77. Brustel, J.; Muramoto, T.; Fumimoto, K.; Ellins, J.; Pears, C.J.; Lakin, N.D. Linking DNA repair and cell cycle progression through serine ADP-ribosylation of histones. Nat. Commun. 2022, 13, 185. [Google Scholar] [CrossRef]
  78. Messner, S.; Hottiger, M.O. Histone ADP-ribosylation in DNA repair, replication and transcription. Trends Cell Biol. 2011, 21, 534–542. [Google Scholar] [CrossRef]
  79. Zhang, S.; Roche, K.; Nasheuer, H.-P.; Lowndes, N.F. Modification of Histones by Sugar β-N-Acetylglucosamine (GlcNAc) Occurs on Multiple Residues, Including Histone H3 Serine 10, and Is Cell Cycle-regulated. J. Biol. Chem. 2011, 286, 37483–37495. [Google Scholar] [CrossRef] [Green Version]
  80. Fong, J.J.; Nguyen, B.L.; Bridger, R.; Medrano, E.E.; Wells, L.; Pan, S.; Sifers, R.N. β-N-Acetylglucosamine (O-GlcNAc) Is a Novel Regulator of Mitosis-specific Phosphorylations on Histone H3. J. Biol. Chem. 2012, 287, 12195–12203. [Google Scholar] [CrossRef]
  81. Sakabe, K.; Hart, G.W. O-GlcNAc Transferase Regulates Mitotic Chromatin Dynamics. J. Biol. Chem. 2010, 285, 34460–34468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Ma, Z.; Vocadlo, D.J.; Vosseller, K. Hyper-O-GlcNAcylation Is Anti-apoptotic and Maintains Constitutive NF-κB Activity in Pancreatic Cancer Cells. J. Biol. Chem. 2013, 288, 15121–15130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Chen, Q.; Chen, Y.; Bian, C.; Fujiki, R.; Yu, X. TET2 promotes histone O-GlcNAcylation during gene transcription. Nature 2013, 493, 561–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhang, Q.; Liu, X.; Gao, W.; Li, P.; Hou, J.; Li, J.; Wong, J. Differential Regulation of the Ten-Eleven Translocation (TET) Family of Dioxygenases by O-Linked β-N-Acetylglucosamine Transferase (OGT). J. Biol. Chem. 2014, 289, 5986–5996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Singh, J.P.; Zhang, K.; Wu, J.; Yang, X. O-GlcNAc signaling in cancer metabolism and epigenetics. Cancer Lett. 2015, 356, 244–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Nelson, C.J.; Santos-Rosa, H.; Kouzarides, T. Proline isomerization of histone H3 regulates lysine methylation and gene expression. Cell 2006, 126, 905–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Howe, F.S.; Boubriak, I.; Sale, M.J.; Nair, A.; Clynes, D.; Grijzenhout, A.; Murray, S.C.; Woloszczuk, R.; Mellor, J. Lysine Acetylation Controls Local Protein Conformation by Influencing Proline Isomerization. Mol. Cell 2014, 55, 733–744. [Google Scholar] [CrossRef] [Green Version]
  88. Wainwright, E.N.; Scaffidi, P. Epigenetics and Cancer Stem Cells: Unleashing, Hijacking, and Restricting Cellular Plasticity. Trends Cancer 2017, 3, 372–386. [Google Scholar] [CrossRef] [Green Version]
  89. Easwaran, H.; Tsai, H.C.; Baylin, S.B. Cancer epigenetics: Tumor heterogeneity, plasticity of stem-like states, and drug resistance. Mol. Cell 2014, 54, 716–727. [Google Scholar] [CrossRef] [Green Version]
  90. Bates, S.E. Epigenetic Therapies for Cancer. N. Engl. J. Med. 2020, 383, 650–663. [Google Scholar] [CrossRef] [PubMed]
  91. Audia, J.E.; Campbell, R.M. Histone Modifications and Cancer. Cold Spring Harb. Perspect. Biol. 2016, 8, a019521. [Google Scholar] [CrossRef] [PubMed]
  92. Mio, C.; Bulotta, S.; Russo, D.; Damante, G. Reading Cancer: Chromatin Readers as Druggable Targets for Cancer Treatment. Cancers 2019, 11, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Kim, S.; Yu, N.K.; Kaang, B.K. CTCF as a multifunctional protein in genome regulation and gene expression. Exp. Mol. Med. 2015, 47, e166. [Google Scholar] [CrossRef] [Green Version]
  94. Nikolaev, L.G.; Akopov, S.B.; Didych, D.A.; Sverdlov, E.D. Vertebrate Protein CTCF and its Multiple Roles in a Large-Scale Regulation of Genome Activity. Curr. Genom. 2009, 10, 294–302. [Google Scholar] [CrossRef] [PubMed]
  95. Alharbi, A.B.; Schmitz, U.; Bailey, C.G.; Rasko, J.E.J. CTCF as a regulator of alternative splicing: New tricks for an old player. Nucleic Acids Res. 2021, 49, 7825–7838. [Google Scholar] [CrossRef]
  96. Jia, Z.; Li, J.; Ge, X.; Wu, Y.; Guo, Y.; Wu, Q. Tandem CTCF sites function as insulators to balance spatial chromatin contacts and topological enhancer-promoter selection. Genome Biol. 2020, 21, 75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Di Nardo, M.; Pallotta, M.M.; Musio, A. The multifaceted roles of cohesin in cancer. J. Exp. Clin. Cancer Res. 2022, 41, 96. [Google Scholar] [CrossRef]
  98. Baranello, L.; Kouzine, F.; Levens, D. CTCF and cohesin cooperate to organize the 3D structure of the mammalian genome. Proc. Natl. Acad. Sci. USA 2014, 111, 889–890. [Google Scholar] [CrossRef] [Green Version]
  99. Rao, S.S.P.; Huang, S.C.; Glenn St Hilaire, B.; Engreitz, J.M.; Perez, E.M.; Kieffer-Kwon, K.R.; Sanborn, A.L.; Johnstone, S.E.; Bascom, G.D.; Bochkov, I.D.; et al. Cohesin Loss Eliminates All Loop Domains. Cell 2017, 171, 305–320. [Google Scholar] [CrossRef] [Green Version]
  100. Peters, J.M.; Tedeschi, A.; Schmitz, J. The cohesin complex and its roles in chromosome biology. Genes Dev. 2008, 22, 3089–3114. [Google Scholar] [CrossRef]
  101. Oh, S.; Oh, C.; Yoo, K.H. Functional roles of CTCF in breast cancer. BMB Rep. 2017, 50, 445–453. [Google Scholar] [CrossRef] [Green Version]
  102. Cheng, H.; Zhang, N.; Pati, D. Cohesin subunit RAD21: From biology to disease. Gene 2020, 758, 144966. [Google Scholar] [CrossRef] [PubMed]
  103. Sarogni, P.; Palumbo, O.; Servadio, A.; Astigiano, S.; D’Alessio, B.; Gatti, V.; Cukrov, D.; Baldari, S.; Pallotta, M.M.; Aretini, P.; et al. Overexpression of the cohesin-core subunit SMC1A contributes to colorectal cancer development. J. Exp. Clin. Cancer Res. 2019, 38, 108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Hill, V.K.; Kim, J.S.; Waldman, T. Cohesin mutations in human cancer. Biochim. Biophys. Acta 2016, 1866, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Weinhold, B. Epigenetics: The science of change. Environ. Health Perspect. 2006, 114, A160–A167. [Google Scholar] [CrossRef] [Green Version]
  106. Jin, Y.; Liu, T.; Luo, H.; Liu, Y.; Liu, D. Targeting Epigenetic Regulatory Enzymes for Cancer Therapeutics: Novel Small-Molecule Epidrug Development. Front. Oncol. 2022, 12, 848221. [Google Scholar] [CrossRef]
  107. Liu, N.; Zhao, R.; Ma, Y.; Wang, D.; Yan, C.; Zhou, D.; Yin, F.; Li, Z. The Development of Epigenetics and Related Inhibitors for Targeted Drug Design in Cancer Therapy. Curr. Top. Med. Chem. 2018, 18, 2380–2394. [Google Scholar] [CrossRef]
  108. Bennett, R.L.; Licht, J.D. Targeting Epigenetics in Cancer. Annu. Rev. Pharmacol. Toxicol. 2018, 58, 187–207. [Google Scholar] [CrossRef]
  109. Kantarjian, H.; Issa, J.P.; Rosenfeld, C.S.; Bennett, J.M.; Albitar, M.; DiPersio, J.; Klimek, V.; Slack, J.; de Castro, C.; Ravandi, F.; et al. Decitabine improves patient outcomes in myelodysplastic syndromes: Results of a phase III randomized study. Cancer 2006, 106, 1794–1803. [Google Scholar] [CrossRef]
  110. Silverman, L.R.; Demakos, E.P.; Peterson, B.L.; Kornblith, A.B.; Holland, J.C.; Odchimar-Reissig, R.; Stone, R.M.; Nelson, D.; Powell, B.L.; DeCastro, C.M.; et al. Randomized controlled trial of azacitidine in patients with the myelodysplastic syndrome: A study of the cancer and leukemia group B. J. Clin. Oncol. 2002, 20, 2429–2440. [Google Scholar] [CrossRef]
  111. Marquez, V.E.; Barchi, J.J., Jr.; Kelley, J.A.; Rao, K.V.; Agbaria, R.; Ben-Kasus, T.; Cheng, J.C.; Yoo, C.B.; Jones, P.A. Zebularine: A unique molecule for an epigenetically based strategy in cancer chemotherapy. The magic of its chemistry and biology. Nucleosides Nucleotides Nucleic Acids 2005, 24, 305–318. [Google Scholar] [CrossRef] [PubMed]
  112. Rau, R.E.; Rodriguez, B.A.; Luo, M.; Jeong, M.; Rosen, A.; Rogers, J.H.; Campbell, C.T.; Daigle, S.R.; Deng, L.; Song, Y.; et al. DOT1L as a therapeutic target for the treatment of DNMT3A-mutant acute myeloid leukemia. Blood 2016, 128, 971–981. [Google Scholar] [CrossRef] [Green Version]
  113. Daigle, S.R.; Olhava, E.J.; Therkelsen, C.A.; Basavapathruni, A.; Jin, L.; Boriack-Sjodin, P.A.; Allain, C.J.; Klaus, C.R.; Raimondi, A.; Scott, M.P.; et al. Potent inhibition of DOT1L as treatment of MLL-fusion leukemia. Blood 2013, 122, 1017–1025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Yu, W.; Chory, E.J.; Wernimont, A.K.; Tempel, W.; Scopton, A.; Federation, A.; Marineau, J.J.; Qi, J.; Barsyte-Lovejoy, D.; Yi, J.; et al. Catalytic site remodelling of the DOT1L methyltransferase by selective inhibitors. Nat. Commun. 2012, 3, 1288. [Google Scholar] [CrossRef] [Green Version]
  115. Konze, K.D.; Ma, A.; Li, F.; Barsyte-Lovejoy, D.; Parton, T.; Macnevin, C.J.; Liu, F.; Gao, C.; Huang, X.P.; Kuznetsova, E.; et al. An orally bioavailable chemical probe of the Lysine Methyltransferases EZH2 and EZH1. ACS Chem. Biol. 2013, 8, 1324–1334. [Google Scholar] [CrossRef] [PubMed]
  116. Tan, J.; Yang, X.; Zhuang, L.; Jiang, X.; Chen, W.; Lee, P.L.; Karuturi, R.K.; Tan, P.B.; Liu, E.T.; Yu, Q. Pharmacologic disruption of Polycomb-repressive complex 2-mediated gene repression selectively induces apoptosis in cancer cells. Genes Dev. 2007, 21, 1050–1063. [Google Scholar] [CrossRef] [Green Version]
  117. McCabe, M.T.; Ott, H.M.; Ganji, G.; Korenchuk, S.; Thompson, C.; Van Aller, G.S.; Liu, Y.; Graves, A.P.; Iii, A.D.P.; Diaz, E.; et al. EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations. Nature 2012, 492, 108–112. [Google Scholar] [CrossRef]
  118. Sarkozy, C.; Morschhauser, F.; Dubois, S.; Molina, T.; Michot, J.M.; Cullières-Dartigues, P.; Suttle, B.; Karlin, L.; Le Gouill, S.; Picquenot, J.M.; et al. A LYSA Phase Ib Study of Tazemetostat (EPZ-6438) plus R-CHOP in Patients with Newly Diagnosed Diffuse Large B-Cell Lymphoma (DLBCL) with Poor Prognosis Features. Clin. Cancer Res. 2020, 26, 3145–3153. [Google Scholar] [CrossRef] [Green Version]
  119. Qi, W.; Chan, H.; Teng, L.; Li, L.; Chuai, S.; Zhang, R.; Zeng, J.; Li, M.; Fan, H.; Lin, Y.; et al. Selective inhibition of Ezh2 by a small molecule inhibitor blocks tumor cells proliferation. Proc. Natl. Acad. Sci. USA 2012, 109, 21360–21365. [Google Scholar] [CrossRef] [Green Version]
  120. Knutson, S.K.; Wigle, T.J.; Warholic, N.M.; Sneeringer, C.J.; Allain, C.J.; Klaus, C.R.; Sacks, J.D.; Raimondi, A.; Majer, C.R.; Song, J.; et al. A selective inhibitor of EZH2 blocks H3K27 methylation and kills mutant lymphoma cells. Nat. Chem. Biol. 2012, 8, 890–896. [Google Scholar] [CrossRef]
  121. Chang, Y.; Zhang, X.; Horton, J.R.; Upadhyay, A.K.; Spannhoff, A.; Liu, J.; Snyder, J.P.; Bedford, M.T.; Cheng, X. Structural basis for G9a-like protein lysine methyltransferase inhibition by BIX-01294. Nat. Struct. Mol. Biol. 2009, 16, 312–317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Yuan, Y.; Wang, Q.; Paulk, J.; Kubicek, S.; Kemp, M.M.; Adams, D.J.; Shamji, A.F.; Wagner, B.K.; Schreiber, S.L. A small-molecule probe of the histone methyltransferase G9a induces cellular senescence in pancreatic adenocarcinoma. ACS Chem. Biol. 2012, 7, 1152–1157. [Google Scholar] [CrossRef]
  123. Vedadi, M.; Barsyte-Lovejoy, D.; Liu, F.; Rival-Gervier, S.; Allali-Hassani, A.; Labrie, V.; Wigle, T.J.; Dimaggio, P.A.; Wasney, G.A.; Siarheyeva, A.; et al. A chemical probe selectively inhibits G9a and GLP methyltransferase activity in cells. Nat. Chem. Biol. 2011, 7, 566–574. [Google Scholar] [CrossRef] [PubMed]
  124. Greiner, D.; Bonaldi, T.; Eskeland, R.; Roemer, E.; Imhof, A. Identification of a specific inhibitor of the histone methyltransferase SU(VAR)3-9. Nat. Chem. Biol. 2005, 1, 143–145. [Google Scholar] [CrossRef] [PubMed]
  125. Cheung, N.; Fung, T.K.; Zeisig, B.B.; Holmes, K.; Rane, J.K.; Mowen, K.A.; Finn, M.G.; Lenhard, B.; Chan, L.C.; So, C.W. Targeting Aberrant Epigenetic Networks Mediated by PRMT1 and KDM4C in Acute Myeloid Leukemia. Cancer Cell 2016, 29, 32–48. [Google Scholar] [CrossRef] [Green Version]
  126. Eram, M.S.; Shen, Y.; Szewczyk, M.; Wu, H.; Senisterra, G.; Li, F.; Butler, K.V.; Kaniskan, H.; Speed, B.A.; Dela Seña, C.; et al. A Potent, Selective, and Cell-Active Inhibitor of Human Type I Protein Arginine Methyltransferases. ACS Chem. Biol. 2016, 11, 772–781. [Google Scholar] [CrossRef] [Green Version]
  127. Santer, F.R.; Höschele, P.P.; Oh, S.J.; Erb, H.H.; Bouchal, J.; Cavarretta, I.T.; Parson, W.; Meyers, D.J.; Cole, P.A.; Culig, Z. Inhibition of the acetyltransferases p300 and CBP reveals a targetable function for p300 in the survival and invasion pathways of prostate cancer cell lines. Mol. Cancer Ther. 2011, 10, 1644–1655. [Google Scholar] [CrossRef] [Green Version]
  128. Giordano, A.; Tommonaro, G. Curcumin and Cancer. Nutrients 2019, 11, 2376. [Google Scholar] [CrossRef] [Green Version]
  129. Ullah, M.F. Sulforaphane (SFN): An Isothiocyanate in a Cancer Chemoprevention Paradigm. Medicines 2015, 2, 141–156. [Google Scholar] [CrossRef] [Green Version]
  130. von Tresckow, B.; Sayehli, C.; Aulitzky, W.E.; Goebeler, M.E.; Schwab, M.; Braz, E.; Krauss, B.; Krauss, R.; Hermann, F.; Bartz, R.; et al. Phase I study of domatinostat (4SC-202), a class I histone deacetylase inhibitor in patients with advanced hematological malignancies. Eur. J. Haematol. 2019, 102, 163–173. [Google Scholar] [CrossRef]
  131. Brunetto, A.T.; Ang, J.E.; Lal, R.; Olmos, D.; Molife, L.R.; Kristeleit, R.; Parker, A.; Casamayor, I.; Olaleye, M.; Mais, A.; et al. First-in-human, pharmacokinetic and pharmacodynamic phase I study of Resminostat, an oral histone deacetylase inhibitor, in patients with advanced solid tumors. Clin. Cancer Res. 2013, 19, 5494–5504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Assouline, S.E.; Nielsen, T.H.; Yu, S.; Alcaide, M.; Chong, L.; MacDonald, D.; Tosikyan, A.; Kukreti, V.; Kezouh, A.; Petrogiannis-Haliotis, T.; et al. Phase 2 study of panobinostat with or without rituximab in relapsed diffuse large B-cell lymphoma. Blood 2016, 128, 185–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Olsen, E.A.; Kim, Y.H.; Kuzel, T.M.; Pacheco, T.R.; Foss, F.M.; Parker, S.; Frankel, S.R.; Chen, C.; Ricker, J.L.; Arduino, J.M.; et al. Phase IIb multicenter trial of vorinostat in patients with persistent, progressive, or treatment refractory cutaneous T-cell lymphoma. J. Clin. Oncol. 2007, 25, 3109–3115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Piekarz, R.L.; Frye, R.; Prince, H.M.; Kirschbaum, M.H.; Zain, J.; Allen, S.L.; Jaffe, E.S.; Ling, A.; Turner, M.; Peer, C.J.; et al. Phase 2 trial of romidepsin in patients with peripheral T-cell lymphoma. Blood 2011, 117, 5827–5834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Knipstein, J.; Gore, L. Entinostat for treatment of solid tumors and hematologic malignancies. Expert Opin. Investig. Drugs 2011, 20, 1455–1467. [Google Scholar] [CrossRef]
  136. Younes, A.; Oki, Y.; Bociek, R.G.; Kuruvilla, J.; Fanale, M.; Neelapu, S.; Copeland, A.; Buglio, D.; Galal, A.; Besterman, J.; et al. Mocetinostat for relapsed classical Hodgkin’s lymphoma: An open-label, single-arm, phase 2 trial. Lancet Oncol 2011, 12, 1222–1228. [Google Scholar] [CrossRef] [Green Version]
  137. Abaza, Y.M.; Kadia, T.M.; Jabbour, E.J.; Konopleva, M.Y.; Borthakur, G.; Ferrajoli, A.; Estrov, Z.; Wierda, W.G.; Alfonso, A.; Chong, T.H.; et al. Phase 1 dose escalation multicenter trial of pracinostat alone and in combination with azacitidine in patients with advanced hematologic malignancies. Cancer 2017, 123, 4851–4859. [Google Scholar] [CrossRef] [Green Version]
  138. Lee, H.Z.; Kwitkowski, V.E.; Del Valle, P.L.; Ricci, M.S.; Saber, H.; Habtemariam, B.A.; Bullock, J.; Bloomquist, E.; Li Shen, Y.; Chen, X.H.; et al. FDA Approval: Belinostat for the Treatment of Patients with Relapsed or Refractory Peripheral T-cell Lymphoma. Clin. Cancer Res. 2015, 21, 2666–2670. [Google Scholar] [CrossRef] [Green Version]
  139. von Tresckow, B.; Gundermann, S.; Eichenauer, D.A.; Aulitzky, W.E.; Göbeler, M.; Sayehli, C.; Bacchus, L.; Hauns, B.; Mais, A.; Hentsch, B.; et al. First-in-human study of 4SC-202, a novel oral HDAC inhibitor in advanced hematologic malignancies (TOPAS study). J. Clin. Oncol. 2014, 32, 8559. [Google Scholar] [CrossRef]
  140. Evens, A.M.; Balasubramanian, S.; Vose, J.M.; Harb, W.; Gordon, L.I.; Langdon, R.; Sprague, J.; Sirisawad, M.; Mani, C.; Yue, J.; et al. A Phase I/II Multicenter, Open-Label Study of the Oral Histone Deacetylase Inhibitor Abexinostat in Relapsed/Refractory Lymphoma. Clin. Cancer Res. 2016, 22, 1059–1066. [Google Scholar] [CrossRef]
  141. Kim, K.P.; Park, S.J.; Kim, J.E.; Hong, Y.S.; Lee, J.L.; Bae, K.S.; Cha, H.; Kwon, S.K.; Ro, S.; Cho, J.; et al. First-in-human study of the toxicity, pharmacokinetics, and pharmacodynamics of CG200745, a pan-HDAC inhibitor, in patients with refractory solid malignancies. Investig. New Drugs 2015, 33, 1048–1057. [Google Scholar] [CrossRef]
  142. Banerji, U.; van Doorn, L.; Papadatos-Pastos, D.; Kristeleit, R.; Debnam, P.; Tall, M.; Stewart, A.; Raynaud, F.; Garrett, M.D.; Toal, M.; et al. A phase I pharmacokinetic and pharmacodynamic study of CHR-3996, an oral class I selective histone deacetylase inhibitor in refractory solid tumors. Clin. Cancer Res. 2012, 18, 2687–2694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Galloway, T.J.; Wirth, L.J.; Colevas, A.D.; Gilbert, J.; Bauman, J.E.; Saba, N.F.; Raben, D.; Mehra, R.; Ma, A.W.; Atoyan, R.; et al. A Phase I Study of CUDC-101, a Multitarget Inhibitor of HDACs, EGFR, and HER2, in Combination with Chemoradiation in Patients with Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2015, 21, 1566–1573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Qian, C.; Lai, C.J.; Bao, R.; Wang, D.G.; Wang, J.; Xu, G.X.; Atoyan, R.; Qu, H.; Yin, L.; Samson, M.; et al. Cancer network disruption by a single molecule inhibitor targeting both histone deacetylase activity and phosphatidylinositol 3-kinase signaling. Clin. Cancer Res. 2012, 18, 4104–4113. [Google Scholar] [CrossRef] [Green Version]
  145. Furlan, A.; Monzani, V.; Reznikov, L.L.; Leoni, F.; Fossati, G.; Modena, D.; Mascagni, P.; Dinarello, C.A. Pharmacokinetics, safety and inducible cytokine responses during a phase 1 trial of the oral histone deacetylase inhibitor ITF2357 (givinostat). Mol. Med. 2011, 17, 353–362. [Google Scholar] [CrossRef] [PubMed]
  146. Galli, M.; Salmoiraghi, S.; Golay, J.; Gozzini, A.; Crippa, C.; Pescosta, N.; Rambaldi, A. A phase II multiple dose clinical trial of histone deacetylase inhibitor ITF2357 in patients with relapsed or progressive multiple myeloma. Ann. Hematol. 2010, 89, 185–190. [Google Scholar] [CrossRef] [Green Version]
  147. Zwergel, C.; Valente, S.; Jacob, C.; Mai, A. Emerging approaches for histone deacetylase inhibitor drug discovery. Expert Opin. Drug Discov. 2015, 10, 599–613. [Google Scholar] [CrossRef] [PubMed]
  148. Venugopal, B.; Baird, R.; Kristeleit, R.S.; Plummer, R.; Cowan, R.; Stewart, A.; Fourneau, N.; Hellemans, P.; Elsayed, Y.; McClue, S.; et al. A phase I study of quisinostat (JNJ-26481585), an oral hydroxamate histone deacetylase inhibitor with evidence of target modulation and antitumor activity, in patients with advanced solid tumors. Clin. Cancer Res. 2013, 19, 4262–4272. [Google Scholar] [CrossRef] [Green Version]
  149. Shi, Y.; Dong, M.; Hong, X.; Zhang, W.; Feng, J.; Zhu, J.; Yu, L.; Ke, X.; Huang, H.; Shen, Z.; et al. Results from a multicenter, open-label, pivotal phase II study of chidamide in relapsed or refractory peripheral T-cell lymphoma. Ann. Oncol. 2015, 26, 1766–1771. [Google Scholar] [CrossRef]
  150. Santo, L.; Hideshima, T.; Kung, A.L.; Tseng, J.C.; Tamang, D.; Yang, M.; Jarpe, M.; van Duzer, J.H.; Mazitschek, R.; Ogier, W.C.; et al. Preclinical activity, pharmacodynamic, and pharmacokinetic properties of a selective HDAC6 inhibitor, ACY-1215, in combination with bortezomib in multiple myeloma. Blood 2012, 119, 2579–2589. [Google Scholar] [CrossRef]
  151. Pauer, L.R.; Olivares, J.; Cunningham, C.; Williams, A.; Grove, W.; Kraker, A.; Olson, S.; Nemunaitis, J. Phase I study of oral CI-994 in combination with carboplatin and paclitaxel in the treatment of patients with advanced solid tumors. Cancer Investig. 2004, 22, 886–896. [Google Scholar] [CrossRef]
  152. Guzman, M.L.; Yang, N.; Sharma, K.K.; Balys, M.; Corbett, C.A.; Jordan, C.T.; Becker, M.W.; Steidl, U.; Abdel-Wahab, O.; Levine, R.L.; et al. Selective activity of the histone deacetylase inhibitor AR-42 against leukemia stem cells: A novel potential strategy in acute myelogenous leukemia. Mol. Cancer Ther. 2014, 13, 1979–1990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Iannitti, T.; Palmieri, B. Clinical and experimental applications of sodium phenylbutyrate. Drugs R D 2011, 11, 227–249. [Google Scholar] [CrossRef] [PubMed]
  154. Reid, T.; Valone, F.; Lipera, W.; Irwin, D.; Paroly, W.; Natale, R.; Sreedharan, S.; Keer, H.; Lum, B.; Scappaticci, F.; et al. Phase II trial of the histone deacetylase inhibitor pivaloyloxymethyl butyrate (Pivanex, AN-9) in advanced non-small cell lung cancer. Lung Cancer 2004, 45, 381–386. [Google Scholar] [CrossRef] [PubMed]
  155. Bilen, M.A.; Fu, S.; Falchook, G.S.; Ng, C.S.; Wheler, J.J.; Abdelrahim, M.; Erguvan-Dogan, B.; Hong, D.S.; Tsimberidou, A.M.; Kurzrock, R.; et al. Phase I trial of valproic acid and lenalidomide in patients with advanced cancer. Cancer Chemother. Pharmacol. 2015, 75, 869–874. [Google Scholar] [CrossRef]
  156. Zheng, Y.C.; Yu, B.; Jiang, G.Z.; Feng, X.J.; He, P.X.; Chu, X.Y.; Zhao, W.; Liu, H.M. Irreversible LSD1 Inhibitors: Application of Tranylcypromine and Its Derivatives in Cancer Treatment. Curr. Top. Med. Chem. 2016, 16, 2179–2188. [Google Scholar] [CrossRef]
  157. Maes, T.; Carceller, E.; Salas, J.; Ortega, A.; Buesa, C. Advances in the development of histone lysine demethylase inhibitors. Curr. Opin. Pharmacol. 2015, 23, 52–60. [Google Scholar] [CrossRef]
  158. Sugino, N.; Kawahara, M.; Tatsumi, G.; Kanai, A.; Matsui, H.; Yamamoto, R.; Nagai, Y.; Fujii, S.; Shimazu, Y.; Hishizawa, M.; et al. A novel LSD1 inhibitor NCD38 ameliorates MDS-related leukemia with complex karyotype by attenuating leukemia programs via activating super-enhancers. Leukemia 2017, 31, 2303–2314. [Google Scholar] [CrossRef]
  159. Heinemann, B.; Nielsen, J.M.; Hudlebusch, H.R.; Lees, M.J.; Larsen, D.V.; Boesen, T.; Labelle, M.; Gerlach, L.-O.; Birk, P.; Helin, K. Inhibition of demethylases by GSK-J1/J4. Nature 2014, 514, E1–E2. [Google Scholar] [CrossRef]
  160. Hancock, R.L.; Dunne, K.; Walport, L.J.; Flashman, E.; Kawamura, A. Epigenetic regulation by histone demethylases in hypoxia. Epigenomics 2015, 7, 791–811. [Google Scholar] [CrossRef] [PubMed]
  161. Dawson, M.A.; Prinjha, R.K.; Dittmann, A.; Giotopoulos, G.; Bantscheff, M.; Chan, W.I.; Robson, S.C.; Chung, C.W.; Hopf, C.; Savitski, M.M.; et al. Inhibition of BET recruitment to chromatin as an effective treatment for MLL-fusion leukaemia. Nature 2011, 478, 529–533. [Google Scholar] [CrossRef] [Green Version]
  162. Chaidos, A.; Caputo, V.; Gouvedenou, K.; Liu, B.; Marigo, I.; Chaudhry, M.S.; Rotolo, A.; Tough, D.F.; Smithers, N.N.; Bassil, A.K.; et al. Potent antimyeloma activity of the novel bromodomain inhibitors I-BET151 and I-BET762. Blood 2014, 123, 697–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Nepali, K.; Liou, J.-P. Recent developments in epigenetic cancer therapeutics: Clinical advancement and emerging trends. J. Biomed. Sci. 2021, 28, 27. [Google Scholar] [CrossRef]
  164. Valdez, B.C.; Li, Y.; Murray, D.; Corn, P.; Champlin, R.E.; Andersson, B.S. 5-Aza-2’-deoxycytidine sensitizes busulfan-resistant myeloid leukemia cells by regulating expression of genes involved in cell cycle checkpoint and apoptosis. Leuk. Res. 2010, 34, 364–372. [Google Scholar] [CrossRef] [Green Version]
  165. Fu, S.; Hu, W.; Iyer, R.; Kavanagh, J.J.; Coleman, R.L.; Levenback, C.F.; Sood, A.K.; Wolf, J.K.; Gershenson, D.M.; Markman, M.; et al. Phase 1b-2a study to reverse platinum resistance through use of a hypomethylating agent, azacitidine, in patients with platinum-resistant or platinum-refractory epithelial ovarian cancer. Cancer 2011, 117, 1661–1669. [Google Scholar] [CrossRef]
  166. Singal, R.; Ramachandran, K.; Gordian, E.; Quintero, C.; Zhao, W.; Reis, I.M. Phase I/II study of azacitidine, docetaxel, and prednisone in patients with metastatic castration-resistant prostate cancer previously treated with docetaxel-based therapy. Clin. Genitourin. Cancer 2015, 13, 22–31. [Google Scholar] [CrossRef] [PubMed]
  167. Issa, J.P.; Garcia-Manero, G.; Giles, F.J.; Mannari, R.; Thomas, D.; Faderl, S.; Bayar, E.; Lyons, J.; Rosenfeld, C.S.; Cortes, J.; et al. Phase 1 study of low-dose prolonged exposure schedules of the hypomethylating agent 5-aza-2’-deoxycytidine (decitabine) in hematopoietic malignancies. Blood 2004, 103, 1635–1640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Covre, A.; Coral, S.; Di Giacomo, A.M.; Taverna, P.; Azab, M.; Maio, M. Epigenetics meets immune checkpoints. Semin. Oncol. 2015, 42, 506–513. [Google Scholar] [CrossRef] [PubMed]
  169. Maio, M.; Covre, A.; Fratta, E.; Di Giacomo, A.M.; Taverna, P.; Natali, P.G.; Coral, S.; Sigalotti, L. Molecular Pathways: At the Crossroads of Cancer Epigenetics and Immunotherapy. Clin. Cancer Res. 2015, 21, 4040–4047. [Google Scholar] [CrossRef] [Green Version]
  170. Rasmussen, K.D.; Helin, K. Role of TET enzymes in DNA methylation, development, and cancer. Genes Dev. 2016, 30, 733–750. [Google Scholar] [CrossRef]
  171. Dang, L.; Yen, K.; Attar, E.C. IDH mutations in cancer and progress toward development of targeted therapeutics. Ann. Oncol. 2016, 27, 599–608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Damaschke, N.A.; Gawdzik, J.; Avilla, M.; Yang, B.; Svaren, J.; Roopra, A.; Luo, J.H.; Yu, Y.P.; Keles, S.; Jarrard, D.F. CTCF loss mediates unique DNA hypermethylation landscapes in human cancers. Clin. Epigenet. 2020, 12, 80. [Google Scholar] [CrossRef] [PubMed]
  173. Bailey, C.G.; Metierre, C.; Feng, Y.; Baidya, K.; Filippova, G.N.; Loukinov, D.I.; Lobanenkov, V.V.; Semaan, C.; Rasko, J.E. CTCF Expression is Essential for Somatic Cell Viability and Protection Against Cancer. Int. J. Mol. Sci. 2018, 19, 3832. [Google Scholar] [CrossRef] [Green Version]
  174. Li, Y.; Seto, E. HDACs and HDAC Inhibitors in Cancer Development and Therapy. Cold Spring Harb. Perspect. Med. 2016, 6, a026831. [Google Scholar] [CrossRef] [Green Version]
  175. Campbell, P.; Thomas, C.M. Belinostat for the treatment of relapsed or refractory peripheral T-cell lymphoma. J. Oncol. Pharm. Pract. 2017, 23, 143–147. [Google Scholar] [CrossRef]
  176. Richon, V.M.; Sandhoff, T.W.; Rifkind, R.A.; Marks, P.A. Histone deacetylase inhibitor selectively induces p21WAF1 expression and gene-associated histone acetylation. Proc. Natl. Acad. Sci. USA 2000, 97, 10014–10019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Xue, K.; Gu, J.J.; Zhang, Q.; Mavis, C.; Hernandez-Ilizaliturri, F.J.; Czuczman, M.S.; Guo, Y. Vorinostat, a histone deacetylase (HDAC) inhibitor, promotes cell cycle arrest and re-sensitizes rituximab- and chemo-resistant lymphoma cells to chemotherapy agents. J. Cancer Res. Clin. Oncol. 2016, 142, 379–387. [Google Scholar] [CrossRef]
  178. Garnock-Jones, K.P. Panobinostat: First global approval. Drugs 2015, 75, 695–704. [Google Scholar] [CrossRef] [PubMed]
  179. Jain, N.; Odenike, O. Emerging role of the histone deacetylase inhibitor romidepsin in hematologic malignancies. Expert Opin. Pharmacother. 2010, 11, 3073–3084. [Google Scholar] [CrossRef]
  180. Wang, Y.M.; Gu, M.L.; Meng, F.S.; Jiao, W.R.; Zhou, X.X.; Yao, H.P.; Ji, F. Histone acetyltransferase p300/CBP inhibitor C646 blocks the survival and invasion pathways of gastric cancer cell lines. Int. J. Oncol. 2017, 51, 1860–1868. [Google Scholar] [CrossRef]
  181. Oike, T.; Komachi, M.; Ogiwara, H.; Amornwichet, N.; Saitoh, Y.; Torikai, K.; Kubo, N.; Nakano, T.; Kohno, T. C646, a selective small molecule inhibitor of histone acetyltransferase p300, radiosensitizes lung cancer cells by enhancing mitotic catastrophe. Radiother. Oncol. 2014, 111, 222–227. [Google Scholar] [CrossRef] [PubMed]
  182. Heimbruch, K.E.; Fisher, J.B.; Stelloh, C.T.; Phillips, E.; Reimer, M.H., Jr.; Wargolet, A.J.; Meyer, A.E.; Pulakanti, K.; Viny, A.D.; Loppnow, J.J.; et al. DOT1L inhibitors block abnormal self-renewal induced by cohesin loss. Sci. Rep. 2021, 11, 7288. [Google Scholar] [CrossRef] [PubMed]
  183. Mould, D.P.; McGonagle, A.E.; Wiseman, D.H.; Williams, E.L.; Jordan, A.M. Reversible inhibitors of LSD1 as therapeutic agents in acute myeloid leukemia: Clinical significance and progress to date. Med. Res. Rev. 2015, 35, 586–618. [Google Scholar] [CrossRef] [PubMed]
  184. Feng, Z.; Yao, Y.; Zhou, C.; Chen, F.; Wu, F.; Wei, L.; Liu, W.; Dong, S.; Redell, M.; Mo, Q.; et al. Pharmacological inhibition of LSD1 for the treatment of MLL-rearranged leukemia. J. Hematol. Oncol. 2016, 9, 24. [Google Scholar] [CrossRef] [Green Version]
  185. Agboyibor, C.; Dong, J.; Effah, C.Y.; Drokow, E.K.; Pervaiz, W.; Liu, H.M. LSD1 as a Biomarker and the Outcome of Its Inhibitors in the Clinical Trial: The Therapy Opportunity in Tumor. J. Oncol. 2021, 2021, 5512524. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Techniques for studying chromatin organization in three-dimensional (3D) space. Steps involved in studying the 3D organization of genomes, (a) A common step in 3C-related techniques is crosslinking chromosomes with formaldehyde and their fragmentation by restriction digestion. The organization of chromatin can be visualized in 3D space using several different detection approaches such as (b) 3C (chromosome conformation capture), (c) 4C (chromosome conformation capture-on-Chip), (d) 5C (chromosome conformation capture carbon copy), (e) Hi-C, (f) Micro-C, (g) ChIA-PET (chromatin interaction analysis paired-end tag sequencing) and (h) Capture-C. This figure was created using BioRender.com access on 14 December 2022.
Figure 1. Techniques for studying chromatin organization in three-dimensional (3D) space. Steps involved in studying the 3D organization of genomes, (a) A common step in 3C-related techniques is crosslinking chromosomes with formaldehyde and their fragmentation by restriction digestion. The organization of chromatin can be visualized in 3D space using several different detection approaches such as (b) 3C (chromosome conformation capture), (c) 4C (chromosome conformation capture-on-Chip), (d) 5C (chromosome conformation capture carbon copy), (e) Hi-C, (f) Micro-C, (g) ChIA-PET (chromatin interaction analysis paired-end tag sequencing) and (h) Capture-C. This figure was created using BioRender.com access on 14 December 2022.
Cancers 15 00215 g001
Figure 2. Factors involved in regulating epigenetic code. Histones undergoes different PTMs such as methylation, acetylation, phosphorylation, ubiquitination and ADP-ribosylation. Writers and erasers are involved in adding and removing PTM, respectively. Readers with specific domains recognizes the PTMs. This figure was created using BioRender.com (accessed on 14 December 2022).
Figure 2. Factors involved in regulating epigenetic code. Histones undergoes different PTMs such as methylation, acetylation, phosphorylation, ubiquitination and ADP-ribosylation. Writers and erasers are involved in adding and removing PTM, respectively. Readers with specific domains recognizes the PTMs. This figure was created using BioRender.com (accessed on 14 December 2022).
Cancers 15 00215 g002
Figure 3. Several levels of organization are present in genomes. Inside the nucleus, DNA (black) is wrapped around nucleosomes (orange), which consist of octamers of core histone proteins H2A, H2B, H3, and H4. Cohesin (red) and CTCF (green) colocalize in the mammalian genomes to form high-order chromatin structures. Chromatin fibers are folded into loops, bringing upstream regulatory elements close to gene promoters. TADs (brown) are formed when chromatin is extruded and insulated in a physical compartment. DNA resides inside chromosome territories (multiple colors), generating non-random arrangements of chromosomes and genes within the nucleus. This figure was created using BioRender.com (accessed on 14 December 2022).
Figure 3. Several levels of organization are present in genomes. Inside the nucleus, DNA (black) is wrapped around nucleosomes (orange), which consist of octamers of core histone proteins H2A, H2B, H3, and H4. Cohesin (red) and CTCF (green) colocalize in the mammalian genomes to form high-order chromatin structures. Chromatin fibers are folded into loops, bringing upstream regulatory elements close to gene promoters. TADs (brown) are formed when chromatin is extruded and insulated in a physical compartment. DNA resides inside chromosome territories (multiple colors), generating non-random arrangements of chromosomes and genes within the nucleus. This figure was created using BioRender.com (accessed on 14 December 2022).
Cancers 15 00215 g003
Table 1. Proteins involved in chromatin organization, adapted from [90,91,92].
Table 1. Proteins involved in chromatin organization, adapted from [90,91,92].
ModificationGene/ComponentsEnzyme/Action
WritersDNMT1DNA methyltransferase
DNMT3aDNA methyltransferase
DNMT3bDNA methyltransferase
EZH1/2Histone methyltransferase
DOT1L/KMT4Histone methyltransferase
EP300 (P300/KAT3B)Histone acetyltransferase
CREBBP (CBP/KAT3A)Histone acetyltransferase
GNAT family (GCN5)Histone acetyltransferase
MYST family (TIP60/KAT5)Histone acetyltransferase
EHMT1/2Histone methyltransferase
SUV39H1/2Histone methyltransferase
PRMT1/2/4/5/6/7Histone methyltransferase
KMT7Histone methyltransferase
AuroraB (AURKB)Kinases
WSTF (BAZ1B)Kinases
BRCA1-BARD1Ubiquitin-ligases
PARP1ADP-ribosyltransferases
ErasersTET1/TET2DNA demethylation
HDAC 1-3,8 (Class I)Histone deacetylase
HDAC 4,5,7,9 (Class IIa)Histone deacetylase
HDAC 6,10 (Class IIb)Histone deacetylase
HDAC Class III: SIRT 1-7Histone deacetylase
HDAC 11 (Class IV)Histone deacetylase
KDM1A/LSD1, KDM1B/LSD2Histone demethylase
KDM2A/2B, KDM3A/3BHistone demethylase
KDM4A/4B/4C/4D/4EHistone demethylase
KDM5A/5B/5C/5DHistone demethylase
KDM6A/6B/7/8Histone demethylase
PP1 (PPP1CA)/PP2 (PPP2CA)Phosphatases
PPgamma (PPARG)Phosphatases
EYA1/3Phosphatases
OTUB1/2, BRCC36, USP3/16/26/44Isopeptidases
PARG, MDO1/2, TARGADP-ribosylhydrolases
ReadersMORF, MRG15 (chromodomain)Reads methylation
MBT, PHF1/19, TDRD7 (Tudor domain)Reads methylation
BRPF1, NSD1-3 (PWWP domain)Reads methylation
G9a/GLP (Ankyrin repeats)Reads methylation
BRD2/3/4/T (bromodomain)Reads acetylation
XRCC1, NBS1, BARD1 (BIR domain)Reads Phosphorylation
14-3-3β/γ/η/ε/μ(14-3-3 proteins BRCT domain)Reads Phosphorylation
53BP1Ubiquitin lysine reader
RNF146 (Macrodomains)Reads ADP-ribosylation
APLF, CHFR (PBZ)Reads ADP-ribosylation
Table 2. Epigenetic inhibitors in different cancer types (Adapted from [108]).
Table 2. Epigenetic inhibitors in different cancer types (Adapted from [108]).
CategoryEpigenetic RegulationTargetCancer TypesKnown CompoundReference/https://clinicaltrials.gov Identifier
WritersDNMTDNMT1MDS, AML5-aza-2′-deoxycytidine (5-aza-CdR; decitabine, Dacogen®)[109]
DNMT1MDS, AML5-azacytidine (5-aza-CR; Aza; Vidaza®)[110]
DNMT1, DNMT3,Hematological malignancies, and solid tumorsZebularine (NSC309132; 4-deoxyuridine)[111]
DNMT1Hematological malignancies and solid tumorsGuadecitabine (SGI-110)NCT01261312, NCT01752933
HMTDOT1LHematological malignanciesEPZ00477[112]
DOT1LHematological malignanciesPinometostat (EPZ-5676)[113]
DOT1LLeukemiaSGC0946[114]
EZH1DLBCLUNC1999[115]
EZH2AML3-Deazaneplanocin A (DZnep)[116]
EZH2B-cell lymphomaGSK126[117]
EZH2B-cell lymphomaTazemetostat (EPZ-6348)[118]
EZH2DLBCLEI1[119]
EZH2Non-Hodgkin lymphomaEPZ005687[120]
EZH2Ovarian cancerGSK343[117]
EZH2Breast, colon, prostate cancerDZNep[116]
G9a/EHMT2Leukemia, bladder cancerBIX-01294[121]
G9a/EHMT2Pancreatic cancerBRD4770[122]
G9a/EHMT2AML, breast cancerUNC0638[123]
SUV39H1LymphomasChaetocin[124]
PRMT5Solid tumors, non-Hodgkin lymphomaGSK3326595NCT02783300
PRMT1AMLAMI-408[125]
PRMT1, 3, 4, 6, 8TBDMS023[126]
HATp300Prostate cancerC646[127]
p300, CBPMM, breast cancer, pancreatic cancerCurcumin[128]
ErasersHDACClass I, II, IVLeukemia, colorectal cancer, prostate cancer, and other solid tumorsSulforaphane (SFN)[129]
Class ILeukemia, colorectal cancerDomatinostat (4SC-202)[130]
Class I, II, IVLeukemia, colorectal cancer, head and neck cancer, hepatocellular carcinomaResminostat (4SC-201, RAS2410)[131]
Class I, II, IVCTCL, Hodgkin’s lymphoma, breast cancer, head and neck cancer, prostate cancer, colorectal cancer, thyroid cancerPanobinostat (LBH589)[132]
Class I, II, IVCTCL, leukemia, prostate cancer, bladder cancer, breast cancerVorinostat (SAHA, Zolinza®)[133]
Class I, II, IVCTCLRomidepsin (depsipeptide, FK228)[134]
Class I (HDAC1, 9, 11)Hodgkin lymphoma, kidney cancer, breast cancerEntinostat (MS-275, SNDX-275)[135]
Class I, IVFollicular lymphoma, Hodgkin’s lymphoma and AML, CLL, MDS, solid tumorsMocetinostat (MGCD0103)[136]
Class I, II, IVMDS, AMLPracinostat (SB939)[137]
Class I, II, IVLeukemia, colorectal cancer, lung cancer, pancreatic cancerBelinostat (PXD101)[138]
HDAC-LSD1Hematological malignanciesHDAC-LSD1 4SC-202[139]
Class I, IIClinical trials: Hodgkin lymphoma, non-Hodgkin lymphoma, CLLAbexinostat[140]
Class I, II, IVClinical trials: solid tumorsCG200745[141]
Class IClinical trials: solid tumorsCHR-3996[142]
Class I, IIClinical trials: SCCCUDC-101[143]
Class I, IIClinical trials: MM, lymphoma, solid tumorsCUDC-907[144]
Class I, IIClinical trials: CLL, MM, Hodgkin lymphomaGivinostat[145,146]
Class I, II (HDAC1, 2, 6)Clinical trials: solid tumorsMPT0E028[147]
Class I, IIClinical trials: solid tumors, lymphoma, CTCLQuisinostat[148]
Class I, II (HDAC1, 2, 3, 10)Clinical trials: breast cancer, NSCLCChidamide[149]
Class II (HDAC6)Clinical trials: MM, lymphomaRicolinostat[150]
Class IClinical trials: MM, lung cancer, pancreatic cancerTacedinaline[151]
Class I (HDAC1, 2)Approved: CTCL, PTCLRomidepsin[134]
Class I, IIClinical trials: AML, MMAR-42[152]
Class I, IIClinical trials: solid tumors, hematological malignanciesPhenylbutyrate[153]
Class I, IIClinical trials: NSCLC, MM, CLLPivanex[154]
Class I, IIClinical trials: solid tumors, hematological malignanciesValproic acid[155]
HDMLSD1AML, small cell lung cancerGSK2879552[156]
LSD1AMLGSK354, GSK690[157]
LSD1MDSNCD25, NCD38[158]
LSD1Acute leukemiaORY-1001[156]
LSD1MDS, AMLTranylcypromine[156]
HDAC-LSD1Hematological malignancies4SC-202[139]
JmjC domain proteinsTBDGSK-J1, GSK-J4[159]
KDM5BHematological malignancies and solid tumorsEPT-103182[160]
ReadersBETBET proteinsProstate cancer, AML with mixed lineage leukemia translocations, MM, NUT midline carcinomaJQ1[161]
BET proteinsClinical trials: Hematological malignancies, NUT midline carcinoma, solid tumorsI-BET762NCT01943851, NCT01587703, [162]
BRD2, 3, 4Clinical trials: AMLOTX015NCT01713582
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mohammad, A.; Jha, S. Epimutations and Their Effect on Chromatin Organization: Exciting Avenues for Cancer Treatment. Cancers 2023, 15, 215. https://doi.org/10.3390/cancers15010215

AMA Style

Mohammad A, Jha S. Epimutations and Their Effect on Chromatin Organization: Exciting Avenues for Cancer Treatment. Cancers. 2023; 15(1):215. https://doi.org/10.3390/cancers15010215

Chicago/Turabian Style

Mohammad, Asad, and Sudhakar Jha. 2023. "Epimutations and Their Effect on Chromatin Organization: Exciting Avenues for Cancer Treatment" Cancers 15, no. 1: 215. https://doi.org/10.3390/cancers15010215

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop