Next Article in Journal
Toxicity Study and Binding Analysis of Newly Synthesized Antifungal N-(4-aryl/cyclohexyl)-2-(pyridine-4-yl carbonyl) hydrazinecarbothioamide Derivative with Bovine Serum Albumin
Next Article in Special Issue
Roles of Histone H2A Variants in Cancer Development, Prognosis, and Treatment
Previous Article in Journal
Unravelling the Mechanism and Governing Factors in Lewis Acid and Non-Covalent Diels–Alder Catalysis: Different Perspectives
Previous Article in Special Issue
The Roles of Histone Post-Translational Modifications in the Formation and Function of a Mitotic Chromosome
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Histone Modification in DNA Replication-Coupled Nucleosome Assembly and Cancer

State Key Laboratory of Biotherapy and Cancer Center, and Frontiers Science Center for Disease-Related Molecular Network, West China Hospital, Sichuan University, Chengdu 610041, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(5), 4939; https://doi.org/10.3390/ijms24054939
Submission received: 12 December 2022 / Revised: 28 January 2023 / Accepted: 29 January 2023 / Published: 3 March 2023

Abstract

:
Histone modification regulates replication-coupled nucleosome assembly, DNA damage repair, and gene transcription. Changes or mutations in factors involved in nucleosome assembly are closely related to the development and pathogenesis of cancer and other human diseases and are essential for maintaining genomic stability and epigenetic information transmission. In this review, we discuss the role of different types of histone posttranslational modifications in DNA replication-coupled nucleosome assembly and disease. In recent years, histone modification has been found to affect the deposition of newly synthesized histones and the repair of DNA damage, further affecting the assembly process of DNA replication-coupled nucleosomes. We summarize the role of histone modification in the nucleosome assembly process. At the same time, we review the mechanism of histone modification in cancer development and briefly describe the application of histone modification small molecule inhibitors in cancer therapy.

1. Introduction

The nucleosome is the basic unit of chromatin. It is an octamer composed of 4 core histones (H3, H4, H2A, H2B), including one H3-H4 tetramer and two H2A-H2B dimers, surrounded by 147 pairs of DNA base pairs [1]. The core histones form a spherical core particle, and their N-terminal tails are free from the core particle, which helps the modification occur. Posttranslational modifications (PTMs) are involved in a variety of cellular processes, such as transcription, DNA damage, apoptosis, and cell cycle regulation. Mass spectrometry is a powerful tool for finding and verifying histone PTMs [2]. In addition, new proteomic, genomic, and functional solid-phase chemistry tools have been developed to detect the function of PTMs [3,4].

2. Posttranslational Modifications of Histones

Histone acetylation was first identified by biologist Vincent Allfrey in the 1960s and has been associated with mammalian gene activity [5,6]. Since then, histone PTMs have been discovered and well described. Currently, more than 10 different covalent modifications have been found on different amino acid residues of core histones, including acetylation of lysine, methylation of arginine and lysine, ubiquitination, ADP ribosylation, citrullination, phosphorylation of serine and tyrosine, isomerization of proline, sumoylation, carbonylation, and controversial biotinylation. Moreover, new modification sites and patterns are continuously discovered (Figure 1). Covalent modification can occur not only in the N-terminal tail protruding from the nucleosome but also in the core region. Different combinations of modification sites and forms can encode very rich information, which can be transmitted to daughter cells as epigenetic markers [7]. Therefore, the “histone code” hypothesis has been proposed by Allis et al. in the early 21st century and supposes that the functions derived from rich histone language are extremely extensive and fine, involving all aspects of cell fate determination, such as replication of genetic code, cell adaptation to internal and external environmental changes, regulation of gene expression, and others. Interpreting these messages and elucidating their functions is the main content of epigenetics [8,9,10].
Common sites of histone acetylation are K5, K9, and K13 on histone H2A [11,12]; K5, K12, K15, and K20 on H2B [13]; K9, K14, K18, K23, K27, K56, and K79 on H3 [11]; and K5, K8, K12, K16, and K91 on H4 [14]. The various acetylation sites of histones correspond to different functions. For example, H3K56 acetylation is involved in the regulation of nucleosome assembly, while H4K16 acetylation is involved in the regulation of nucleosome-mediated chromatin compaction, activation or inhibition of gene transcription, DNA damage repair, and other processes. Methylation occurs mainly at K4, K9, K27, K36, and K79 on histone H3 and at K20 of H4 [12]. Most histones are monoubiquitinated rather than polyubiquitinated, which occurs mainly on H2A and H2B [15,16].

3. Nucleosome Assembly

A key feature of chromatin assembly during DNA replication is that it occurs immediately after DNA synthesis, with the first deposited nucleosome detected approximately 250 bp behind the replication fork. Histone deposition and chromatin assembly are important processes throughout S phase DNA synthesis and are indispensable for gene expression [17]. Histone chaperones, including Nap1, Vps75, NASP, FACT, CAF-1, Rtt106, Spt6, Asf1, and DAXX, play a central role in both histone deposition and chromatin assembly and are therefore involved in the regulation of cellular processes [18]. Nucleosomes can block DNA access during the S phase of the cell cycle [19]. Nucleosomes located in front of replication forks are depolymerized so that DNA replication elements can bind to DNA. During DNA replication, parental histones undergo depolymerization before replication forks, and newly synthesized histones are deposited on DNA with the help of histone chaperones to reform nucleosomes and assemble chromatin [20]. Newly synthesized and preexisting histones are randomly and sequentially deposited to assemble “new” nucleosomes; the H3-H4 tetramer and H2A-H2B dimers are deposited on DNA in a chaperone-dependent manner, and this nucleosome assembly is called replication-coupled nucleosome assembly (RCNA) (Figure 2A) [21]. The RCNA process is important for both epigenetic information transmission and genome integrity [22]. In some DNA damage reactions, parental histones must be removed. The removal of histones at DNA damage sites is considered to have many similarities with the RCNA process. In addition, nucleosome assembly during gene transcription and histone exchange can occur throughout the cell cycle and is named replication-independent nucleosome assembly (RINA) (Figure 2B) [21].

4. The Role of Histone Modification in Nucleosome Assembly

4.1. Influence of Histone Modification on the Deposition of Newly Synthesized Histones

Packaging DNA into the chromatin structure is an important step in DNA replication, which not only ensures the high compaction of DNA but also the correct transmission of epigenetic information to daughter cells [23,24,25,26]. The replication of DNA to the correct packaging of chromatin structure depends on the precise modification, transportation, and assembly of newly synthesized histones. In addition to recycling parental histones, chromatin assembly on replication forks requires the deposition of newly synthesized histones. The expression of canonical histones is activated in the late G1/early S phase to ensure the rapid supply of histones during replication and is inhibited in G1, G2, and early mitosis to prevent the adverse consequences of excess histones on DNA metabolism [27,28].
Newly synthesized histones travel with molecular chaperones from the cytoplasm to the nucleus and are modified after translation to facilitate chromatin deposition. In particular, acetylation of the amino terminus tail of H3 and H4 plays an important role in chromatin assembly [29]. The histone acetyltransferase Hat1 can form the Hat1-Hat2 complex with the histone partner Hat2 to acetylate lysine 5 (H4K5) and 12 (H4K12) of histone H4 [30]. In budding yeast, lysine 9 and 27 of histone H3 are acetylated by the acetyltransferases Rtt109 and Gcn5 [31,32], and the acetylation of H4K91 by Hat1 and H3K56 by Rtt109 are both important in replication-coupled chromatin assembly [33,34,35]. H3 can also be acetylated at lysine 14 and 18 in some mammals [36]. In yeast, most newly synthesized histones H3 are acetylated at lysine 56, which is also a marker of newly synthesized histones. However, the acetylation of H3K56 in humans is less than 1.5%, and SETDB1-mediated monomethylation of H3.1K9 marks newly synthesized histones [33]. The histone chaperone Asf1 binds to newly synthesized H3-H4 dimers and presents them to Rtt109 for acetylation [37]. H3K56 acetylation increases the binding affinity of dimer H3-H4 for histone deposition factors CAF-1 and Rtt106, as well as the binding of CAF-1 to chromatin [38]. The Rtt101Mms1/Mms22 complex can facilitate this process by preferentially binding and ubiquitinating H3K56ac at lysine 121/122/125, which weakens the interaction of Asf1-H3-H4 and promotes the transfer of H3-H4 to downstream chromatin assembly factors, such as Rtt106 (Figure 3) [39]. Notably, H3K56 acetylation is not essential, in part because GCN5-mediated acetylation of the amino terminus of H3 also increases the binding affinity of the dimer H3-H4 for CAF-1 and Rtt106 [32].
CAF-1 is a highly conserved histone chaperone that plays a role in the deposition of newly synthesized histones through the interaction between its subunits and H3-H4 dimers [40,41,42]. CAF-1 is recruited to the replication fork through direct interaction with PCNA [43,44], and promotes nucleosome assembly through interaction with Asf1. The CAF-1-Asf1 histone deposition complex binds to a single H3-H4 dimer, and CAF-1’s ability to form homodimers may provide the second H3-H4 dimer required for the deposition of (H3-H4)2 tetramer [45]. In addition, in vitro interaction analysis has shown that CAF-1 can also bind (H3-H4)2 tetramers in monomer form [46]. Interestingly, the binding of H3-H4 to Asf1 stimulated the binding of Asf1 and CAF-1, while the binding of H3-H4 to CAF-1 was mutually exclusive with Asf1, indicating that the H3-H4 transfer process from Asf1 to DNA occurs through CAF-1 [45]. In yeast, the histone chaperone Rtt106 interacts with CAF-1. Rtt106 can also form homodimers, which interact with the K56 region of histone H3 via the double pleckstrin homology domain and bind directly to the newly synthesized (H3-H4)2 tetramer. After the acetylation of H3K56, the affinity between Rtt106 and H3-H4 was enhanced [47,48]. The roles of CAF-1 and Rtt106 in new histone deposition are redundant, and only the simultaneous deletion of both complexes will affect histone deposition. Additionally, FACT participates in the deposition of new histones by forming complexes with CAF-1 or Rtt106 and H3K56Ac-H4 (but not Asf1) [22,49].
Replication protein A complex (RPA) can regulate DNA metabolism [50], including three subunits of Rfa1, Rfa2, and Rfa3 in yeast, which bind single-stranded DNA to replication forks and mediate replication movement. The initial replication-coupled nucleosome assembly begins with the deposition of histones H3-H4 onto the replicated DNA, followed by the rapid incorporation of histones H2A-H2B. RPA can directly bind the unmodified H3-H4 histone complex. In vitro experiments have shown that RPA can promote the formation of single-stranded DNA-(H3-H4) complex and can quickly connect to double-stranded DNA. In this process, a series of H3-H4 chaperones are recruited (RPA subunit Rfa2 can bind the three histone chaperones CAF-1, FACT, and Rtt106) to assist in the assembly of new nucleosomes [51]. The above research shows that the main function of RPA is to provide a “platform” for the incorporation of histones into the replication fork through the coupled nucleosome accompanied by DNA replication. It provides a good model for explaining how epigenetic information is assembled and transmitted during chromatin replication.
In humans, two subtypes of NASP have been identified: sNASP (somatic NASP) and tNASP (testicular NASP) [52]. HSC70 binds to the new histone H3.1, promoting its folding. The newly synthesized histone H3.1 is then presented to HSP90, which, together with the helper chaperone tNASP, promotes the formation of the H3.1-H4 dimer [44]. sNASP is a H3-H4 histone chaperone in cytoplasmic histone processing. sNASP binds the H3.1-H4 heterodimer and presents it to RbAp46 [53]. RbAp46 recruits Hat1 and catalyzes the acetylation of H4K5/12. The histone chaperone Asf1 binds to H3-H4K5/12ac and promotes new histone nuclear import with Importin-4 [54]. In the nucleus, p300/CBP binds and catalyzes acetylation of H3K56, which can further promote Cul4A/DDB1 to catalyze ubiquitination of H3K122 [55]. Acetylation and ubiquitination enable H3.1-H4 to dissociate from Asf1 and be presented to the histone chaperone CAF-1, which is eventually deposited on replicated DNA and involved in the assembly of replication-coupled nucleosomes (Figure 4).

4.2. Role of Histone Modification in DNA Damage Repair during Nucleosome Assembly

DNA replication stress poses a threat to the transmission of genetic information [56]. For example, replication stress may contribute to the development of tumors by promoting changes in histone-related epigenetic marker patterns. When the replication fork encounters obstacles, it inevitably enters a stagnant state, which is prone to collapse, leading to DNA damage or genomic instability [57]. Therefore, these obstacles must be repaired or bypassed to restore normal DNA replication [58]. This replication fork damage bypass occurs through different mechanisms, either by break-induced replication using the DNA polymerase Polδ subunit, yeast Pol32 or human POLD3, or by switching to a sister chromatid template to bypass the damage site [59,60,61]. These mechanisms all occur in the context of nucleosome assembly; thus, histone modification plays an important role in DNA damage repair (DDR) and is one of the criteria for selecting damage repair pathways. After DNA damage, the damaged site is marked by histone modification to regulate the signaling pathway in a timely manner and provide support for the assembly of effector proteins [62].
During DNA replication, the MCM2-MCM7 complex is loaded at replication initiation under the regulation of the origin recognition complex (ORC), Cdc6, and Cdt1 to form a prereplication complex (pre-RC). Then, the MCM complex is phosphorylated by DDK and CDK. Cdc45 is recruited to the MCM in a SLD3-dependent manner, and the GINS complex is recruited to the MCM complex by phosphorylated Sld2 and Sld3 together with Dbp11 to assemble the CMG (CDC45-MCM-Gins) complex [63,64]. Double-stranded DNA unwinds into single-stranded DNA (ssDNA), and RPA rapidly binds to the newly formed ssDNA, protecting it from damage and generating secondary structures. Subsequently, DNA polymerase α (Polα) initiates DNA synthesis, DNA polymerase ε (Polε) continuously synthesizes the leading strand, and DNA polymerase δ (Polδ) synthesizes the legging strand [65]. Ctf4, a yeast homolog of human AND1, links CMG helicase to Polα polymerase to form trimers involved in DNA replication [66].
The human MMS22L-TONSL complex is located in the replication fork and increases enrichment when DNA damage occurs [67]. It has a similar function to the Rtt101/Mms1/Mms22 complex in yeast. It has been found that the absence of the MMS22L-TONSL complex affects replication fork stability in the context of CPT stimulation [68]. The proportion of H3K56ac modification in human cells is less than 1.5% of the total H3 [36], and this proportion varies little throughout the cell cycle [69]. However, the unmethylated H3-H4K20 histone is methylated at the late G2/M stage [70]. In particular, MMS22L-TONSL is able to bind not only to newly synthesized histones as part of a predeposited complex with MCM2 and ASF1, but also to H4K20me0 on nascent chromatin. MMS22L and TONSL were necessary for the recruitment and homologous recombination of RAD51 [70,71] (Figure 5).
Here, we describe the role of a typical histone modification in DDR. H3K56ac is one of the earliest core modifications described in yeast [72] and is deacetylated by the deacetylases Hst3 and Hst4 at the end of S/G2. During DNA damage, Hst3/Hst4 is downregulated in a checkpoint-dependent manner [73], suggesting that H3K56ac modification is crucial in the DNA damage reaction. In fact, yeast cells with defective H3K56 acetylation are highly sensitive to DNA damage agents such as MMS and CPT [74]. Genetic analysis showed that H3K56ac was upstream of the Rtt101/Mms1/Mms22 ubiquitin ligase complex signaling pathway, which is resistant to genotoxic agents [75,76]. Direct evidence for the involvement of H3K56ac in DDR is that fully acetylated H3K56 in vitro increases exposure to DNA sites [77]. This function is unlikely to be related to the role of chromatin assembly in replication fork stability, as cells lacking CAF-1 and Rtt106 are much less sensitive to MMS and CPT than H3K56 acetylated mutants [78]. Members of the Asf1/Rtt109/H3K56ac/Rtt101Mms1/Mms22 pathway are required in the process of MMS and CPT-induced DNA damage [79,80]. Deletion mutations in this pathway can disrupt checkpoint recovery after drug therapy, demonstrating that this pathway plays an important role in the mechanism of DDR template conversion. H3K56ac deposition appears to promote the ubiquitination of some unknown substrate to uncouple the replicating helicase with the polymerase as a prerequisite for blocking the recombinant bypass of the lesion (Figure 6). Consistent with this model, the interaction of Rtt101Mms1/Mms22 with Ctf4 via the amino terminal tail of Mms22 is necessary for the function of H3K56ac in tolerating replication stress. Ubiquitination of unknown factors is used for Mrc1 and Ctf4 to uncouple the helicase CMG with the polymerase and facilitate recombination repair bypass [81].
Histone chaperone FACT is ubiquitinated by Rtt101 in a manner independent of Mms1/Mms22 [82]. Histones are also potential targets for ubiquitination. Studies have found that human histones are ubiquitinated by Cul4A/DDB1 in UV-induced photodimers, and this modification weakens their interaction with DNA and promotes the recruitment of repair proteins [83,84]. FACT is mainly involved in the deposition of newly synthesized H3-H4. In yeast, Spt16 interacts with Pob3 to form FACT, which is a conserved histone chaperone that plays an important role in DNA regulation [85,86]. Deletion of FACT disrupts the chromatin structure of the gene-coding region [87]. Studies have shown that Spt16 is involved in chromatin remodeling in DDR through ubiquitination of H2B [88,89,90]. At the same time, FACT can regulate DDR mediated by homologous recombination (HR) and base excision repair (BER), which proves that FACT is essential for damage repair [91]. Asf1 is critical for histone modification, histone deposition, and DNA replication. Its function in heterochromatin silencing was first identified in yeast [92] and later found as a replication-coupled assembly factor (RCAF) in Drosophila [93]. Asf1 can bind to the histone H3-H4 dimer with Mcm2-7. Histone H3-H4 was modified with specific parental labeling (H4K16ac and H3K9me3) under hydroxyurea treatment, leading to the accumulation of replication forks [94,95]. It is suggested that Mcm2-H3-H4-Asf1 is an intermediate in parental histone assembly and may promote DNA unwinding through its ability to transfer histones during chromatin assembly. CAF-1 promotes a Rad51-dependent replication fork bypass repair pathway [96]. The interaction between CAF-1 and the RecQ helicase Bloom in human cells is conserved, and both factors accumulate in the DNA replication center through replication stress and promote cell survival [97].

5. Histone Modification and Cancer

Histone modification is involved in chromatin remodeling [98], thereby altering chromatin status and gene expression [99] and is very important for gene regulation and genomic stability. Abnormal histone modification can cause abnormal chromatin status or genomic instability, which is often believed to be closely related to the occurrence and development of cancer [100,101,102].

5.1. Histone Methylation in Cancer

Histone methylation, including monomethylation, demethylation, and trimethylation, is regulated by methyltransferases and demethylases and occurs mainly on lysine residues of H3 and H4 [103]. H3K4me1/2/3, H3K36me1/2/3, and H3K79me1/2/3 are transcriptionally active marks, while H3K9me1/2/3 and H3K27me3 are transcriptionally repressive marks [9]. Histone methylation disorder leads to the destruction of gene expression and genomic stability, and the abnormal modification of histone methylation in tumor cells can alter cancer development (Table 1). For example, decreased H3K27me3 and increased H3K4me3 activate the Wnt/β-catenin signaling pathway to promote colorectal cancer cell development [104]. Mutations in histone methylation sites (H3K27M, H3K27I, etc.) are present in approximately 30% of children with glioblastoma [105]. NSD2 maintains genome integrity and reduces disease incidence by methylating H3K36 and DOT1-mediated H3K79 methylation in response to UV radiation-induced DDR [106]. Under normal physiological conditions, the number of H3K9me3 increases dramatically over time at the site of DNA double-strand break damage and participates in DDR. In contrast, in the environment of abnormal tumor cell metabolism, abnormal H3K9me3 inhibits DNA repair [107].
Histone methylation has been used as a promising target for cancer therapy. A large number of methyltransferase inhibitors have been developed and entered clinical trials, mainly against H3K27 and H3K79 methyltransferase, and arginine methyltransferase [108]. EZH2, a methyltransferase of H3K27, is involved in tumor occurrence, metabolism, drug resistance, and immune regulation [109]. Therefore, targeting EZH2 for cancer therapy has become a hot research topic. Strategies for EZH2 inhibitors include targeting methyltransferase activity (GSK126, GSK343, EPZ011989, et al.), breaking PRC2’s structure (SAH-EZH2, Astemizole, MAK683, et al.), or triggering EZH2 degradation (GNA022, ANCR, FBW7, et al.) [109]. For example, EZH2 inhibitor GSK343 can decrease self-renewal and increase sensitivity to chemotherapy in colorectal cancer cells [110]. Some studies have also shown that EZH2 has an antitumor effect [111,112]. For example, GSK126 can increase the number of myeloid-derived suppressor cells (MDSC) and decrease the number of IFNγ+CD8+ T cells, leading to the failure of antitumor therapy. Interestingly, when combined with neutralizing antibodies against the myeloid differentiation antigen GR-1, MDSC-mediated immunosuppression was mitigated and increased the therapeutic effect of GSK126 [113]. Developing a multi-drug combination therapy strategy may address the limitations of single drug therapy. These studies indicate that histone methylation modification plays an important role in the development and prevention of cancer.
Table 1. The role of genes encoding histone methyltransferase in human cancer.
Table 1. The role of genes encoding histone methyltransferase in human cancer.
GeneTumorRoleReference
SETD1AColorectal cancer, Lung cancer, Gastric cancerOncogenic[114,115,116]
SETD1BPancancerSuppressor[117]
MLL1Breast cancer, Cervical carcinoma, Acute myeloid leukemiaOncogenic[118,119,120]
MLL2Bland cancer, Prostate cancerSuppressor[121,122]
MLL3Nasopharyngeal carcinoma, Breast cancer, Pancreatic cancer, Colorectal cancerAmbiguous[123,124,125,126]
MLL4Lung cancer, Medulloblastoma,Suppressor[127,128]
SMYD2Lung cancer, Cervical cancerOncogene[129,130]
SET7Glioma, Colorectal cancer, Lung cancerSuppressor[131,132,133]
SET9Glioma, Lung cancer, Breast cancerSuppressor[131,133,134]
SMYD3Pancreatic cancer, Lung cancer, Breast cancerOverexpression[135,136,137]
SUV39H1Cervical cancer, Prostate cancer, MelanomaOncogene[138,139,140]
SUV39H2Colorectal cancer, Osteosarcoma, Lung cancerOncogene[141,142,143]
G9AColorectal cancer, Bladder cancer, Lung cancer, Breast cancerOncogene[144,145,146,147]
SETDB1Lung cancer, Gastric cancer, Colorectal cancerOncogenic[148,149,150]
PRDM3Pancreatic ductal adenocarcinomaSuppressor[151]
PRDM16Kidney cancer, Prostrate cancer, Lung cancerSuppressor[152,153,154]
EZH1Breast cancer, Hepatocellular carcinomaOncogenic[155,156]
EZH2Lung cancer, Hepatocellular carcinoma, Breast cancer, Gastric cancer, Colorectal cancerOncogene[156,157,158,159,160]
SETD2Prostate cancer, Pancreatic cancer, Leukemogenesis, Hepatocellular carcinomaSuppressor[161,162,163,164]
NSD1Pancreatic cancer, Laryngeal cancerOncogenic[165,166]
NSD2Lung cancer, Colorectal cancer, Breast cancer, Renal cancer, Osteosarcoma, Prostate cancerOncogene[167,168,169,170,171,172]
NSD3Lung cancer, Breast cancer, Colorectal cancer, Pancreatic cancerOncogene[173,174,175]
SETD3Breast cancer, Hepatocellular carcinomaOverexpression[176,177]
ASH1LProstate cancer, LeukemiaOncogenic[178,179]
SETMARAcute myeloid leukemiaSuppressor[180]
PRDM9PancancerOverexpression[181]
DOT1LProstate cancer, Colorectal cancer, Gastric cancer, Ovarian cancer, Breast cancer, Acute myeloid leukemiaOverexpression[119,182,183,184,185,186]
SET8Prostate cancer, Hepatocellular carcinoma, Breast cancerOncogenic[187,188,189]
SUV4-20H2Hepatocellular carcinoma, Breast cancerSuppressor[190,191]

5.2. Histone Acetylation in Cancer

Acetylation is one of the main modifications of histones and is strictly regulated by histone acetyltransferases (HAT) and histone deacetylases (HDAC) to maintain the normal acetylation state, thus controlling the initiation and shutdown of gene transcription. HATs transfer the acetyl group from acetyl-CoA to the amino terminal of the specific lysine residue of the histone, generating an acetate bond. Acetylation is a key epigenetic mechanism in gene regulation [192] and regulates chromatin structure and function through transcriptional capacity [193,194]. Abnormal histone acetylation can disrupt cell homeostasis and affect cell metabolism and gene regulation [195]. Cumulative evidence suggests that abnormal expression of histone modification enzymes is closely related to tumor development (Table 2). The current antitumor treatment of histone acetylation as a therapeutic target is expected to be achieved through the development of HAT and HDAC inhibitors. The first HDAC inhibitor approved for clinical treatment was suberoylanilide hydroxamic acid (SAHA), and more drugs are being developed, such as YF479, which has good antitumor activity and can inhibit the recurrence and metastasis of breast cancer [196,197]. Thus, histone acetylation modification plays a significant role in the occurrence and development of cancer.

5.3. Histone Ubiquitination in Cancer

Ubiquitin (Ub) exists widely in eukaryotes, and ubiquitination is also one of the main posttranscriptional modifications. Posttranslational modification of proteins is a reversible, dynamic process. Histone ubiquitination is dynamically regulated by ubiquitination enzymes and deubiquitination enzymes and can participate in most intracellular processes, including protein degradation, intracellular signaling, endocytosis, and DNA damage reactions [216,217,218]. Histone ubiquitination is the core event of DDR, and DNA damage requires a large number of ubiquitin molecules, which are crucial for preventing abnormal DNA repair and maintaining genomic stability [219]. Histone H3 ubiquitination enzymes mainly include NEDD4 and CUL4A. NEDD4 ubiquitinates histone H3 on lysine 23/36/37 residues in a glucose-dependent manner, specifically recruiting the histone acetyltransferase GCN5 for subsequent H3 acetylation. This mechanism can regulate gene transcription and tumorigenesis in cancer [220]. The RNA demethylase ALKBH5 and the USP22/RNF40 axis regulate histone H2AK119 monoubiquitination to regulate the expression of key genes involved in DNA repair, thus playing a crucial role in the development of osteosarcoma [221]. Rad6 and Bre1 form a well-characterized H2B monoubiquitin enzyme to degrade histones in DDR reactions [222]. USP11 can deubiquitinate H2AK119 and H2BK120 to separate ubiquitin molecules from histones and maintain genomic stability [223]. It is worth noting that the existing studies on histone ubiquitination mainly focus on histone H2A/H2B, and the discovery of histone H3 ubiquitination and the study of its mechanism are also gradually deepening. However, the regulation of histone H3 deubiquitination remains unclear.

5.4. Histone Phosphorylation in Cancer

Histone phosphorylation occurs on serine and tyrosine residues of histones and has been shown to be involved in many cellular life activities, including DNA damage repair, gene transcription, chromatin maintenance and aging, through histone methylation [224,225]. For example, PRK-mediated H3T11 phosphorylation (H3T11ph) hastens the removal of repressive histone H3 lysine 9 (H3K9) methylation by JMJD2C, demonstrating a unique mechanism by which histone phosphorylation activates gene expression. Importantly, the level of H3T11ph correlates with prostate cancer malignancy, suggesting that inhibition of H3T11ph may be a promising therapeutic target [226]. Phosphorylated H3.3 (H3.3S31ph) enhances the binding of the methyltransferase SETD2 to histone proteins, thus promoting gene transcription and highlighting the causal role of H3.3 phosphorylation in tumor metastasis [227]. H3.3S31ph is also involved in the regulation of heterochromatin regions and reduces the demethylation of H3K9me3 to maintain chromatin integrity by downregulating the activity of KDM4B [228]. Pyk1-catalyzed H3T11ph can weaken the binding of Dot1 to chromatin and reduce Dot1-mediated H3K79me3, leading to suppression of autophagy-related gene transcription and uncovering histone modification crosstalk in response to cell metabolism [229]. Additionally, a recent study showed that phosphorylation of histone H3 at serine 10 inhibits methylation of histone H3 at adjacent arginine 8, providing a framework for understanding the effects of phosphoserine on the methylation of adjacent amino acid residues and arginine [230]. In order to function, histone phosphorylation may antagonize its methylation.

6. Conclusions and Perspectives

Nucleosome assembly is a complex and highly regulated process in eukaryotes. Nucleosome assembly requires precise regulation by histone-modifying enzymes, histone chaperones, and histone modifications. In recent years, yeast-based experiments have provided new insights into how to regulate the assembly of novel H3-H4 through histone modification and molecular chaperones. Additionally, nucleosome reassembly maintains replication fork stability, but the mechanism remains elusive. Similarly, histone epigenetic modifications affect the complexity and correlation of newly assembled chromatin structures during DDR [231]. In fact, some aspects of the replication-dependent chromatin assembly process are not discussed here because there is still no evidence that they are associated with the progression and stability of replication forks. Finally, it is worth mentioning that although we focus on how chromatin assembly regulates DNA replication, the effects are mutual. For example, replication stress resulting from replication disorders promotes heterochromatin formation. The combination of genetic and biochemical approaches with genome-wide analysis may help reveal the dynamics of chromosomal remodeling in these different scenarios and understand the molecular mechanism of how defects in replication-coupled chromatin assembly lead to genetic diseases, cancer, and aging.
Epigenetic modification of histones is closely related to disease pathogenesis and can be a molecular signature in cancer [232]. The diversity of histone modifications provides new molecular targets for the treatment of various diseases [233]. It is worth noting that many drugs targeting histone modifications have been developed and used in clinical research in the past decade, which is sufficient to show that histone modification plays a very important role in disease treatment. Therefore, understanding the basic mechanisms controlling epigenetic modification changes will bring new breakthroughs and advances in drug development and treatment of cancer and other human diseases [234]. The development of epigenetic drugs creates a new avenue for the treatment of diseases, which is a huge leap forward in the extension of basic scientific research to clinical drug development. Meanwhile, numerous studies have found that histone modification tandem is closely related to the development and pathogenesis of various diseases, indicating a new direction for the research and development of histone modification inhibitors.

Author Contributions

Conceptualization, Y.Z. (Yaguang Zhang), Q.Z., Y.Z. (Yang Zhang) and J.H.; writing—original draft preparation, Y.Z. (Yaguang Zhang), Q.Z. and Y.Z. (Yang Zhang); writing—review and editing, J.H.; project administration, J.H.; funding acquisition, Y.Z. (Yaguang Zhang) and J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (82203447), 1.3.5 Project for Disciplines of Excellence, West China Hospital (ZYJC21021), the Natural Science Foundation of Sichuan, China (2022NSFSC1424) and the Postdoctoral Research Project, West China Hospital, Sichuan University (18HXBH068).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful to the funding supporters including the National Natural Science Foundation of China, the Natural Science Foundation of Sichuan, and West China Hospital.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Luger, K.; Mader, A.W.; Richmond, R.K.; Sargent, D.F.; Richmond, T.J. Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 1997, 389, 251–260. [Google Scholar] [CrossRef] [PubMed]
  2. Lu, C.; Coradin, M.; Porter, E.G.; Garcia, B.A. Accelerating the Field of Epigenetic Histone Modification Through Mass Spectrometry-Based Approaches. Mol. Cell Proteom. 2021, 20, 100006. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, Y.; Wang, Y. Mapping histone modification-dependent protein interactions with chemical proteomics. Trends Biochem. Sci. 2022, 47, 189–193. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, Y.; Zhang, C.; Jiang, H.; Yang, P.; Lu, H. Fishing the PTM proteome with chemical approaches using functional solid phases. Chem. Soc. Rev. 2015, 44, 8260–8287. [Google Scholar] [CrossRef]
  5. Allfrey, V.G.; Faulkner, R.; Mirsky, A.E. Acetylation and Methylation of Histones and Their Possible Role in the Regulation of Rna Synthesis. Proc. Natl. Acad. Sci. USA 1964, 51, 786–794. [Google Scholar] [CrossRef] [Green Version]
  6. Grunstein, M.; Allis, C.D. Genetics, Biochemistry, and "Simple" Organisms Converge to Unlock Secrets in Histone Biology: The 2018 Albert Lasker Basic Medical Research Award. JAMA 2018, 320, 1233–1234. [Google Scholar] [CrossRef]
  7. Yun, M.; Wu, J.; Workman, J.L.; Li, B. Readers of histone modifications. Cell Res. 2011, 21, 564–578. [Google Scholar] [CrossRef] [Green Version]
  8. Prakash, K.; Fournier, D. Histone Code and Higher-Order Chromatin Folding: A Hypothesis. Genom. Comput. Biol. 2017, 3, e41. [Google Scholar] [CrossRef] [Green Version]
  9. Jenuwein, T.; Allis, C.D. Translating the histone code. Science 2001, 293, 1074–1080. [Google Scholar] [CrossRef] [Green Version]
  10. Soffers, J.H.; Li, X.; Abmayr, S.M.; Workman, J.L. Reading and Interpreting the Histone Acylation Code. Genom. Proteom. Bioinform. 2016, 14, 329–332. [Google Scholar] [CrossRef]
  11. Arnaudo, A.M.; Garcia, B.A. Proteomic characterization of novel histone post-translational modifications. Epigenet. Chromatin 2013, 6, 24. [Google Scholar] [CrossRef] [Green Version]
  12. Blackledge, N.P.; Klose, R.J. Histone lysine methylation: An epigenetic modification? Epigenomics 2010, 2, 151–161. [Google Scholar] [CrossRef]
  13. Wang, X.; Hayes, J.J. Site-specific binding affinities within the H2B tail domain indicate specific effects of lysine acetylation. J. Biol. Chem. 2007, 282, 32867–32876. [Google Scholar] [CrossRef] [Green Version]
  14. Ge, Z.; Nair, D.; Guan, X.; Rastogi, N.; Freitas, M.A.; Parthun, M.R. Sites of acetylation on newly synthesized histone H4 are required for chromatin assembly and DNA damage response signaling. Mol. Cell. Biol. 2013, 33, 3286–3298. [Google Scholar] [CrossRef] [Green Version]
  15. Osley, M.A. Regulation of histone H2A and H2B ubiquitylation. Brief. Funct. Genom. 2006, 5, 179–189. [Google Scholar] [CrossRef]
  16. Weake, V.M.; Workman, J.L. Histone ubiquitination: Triggering gene activity. Mol. Cell 2008, 29, 653–663. [Google Scholar] [CrossRef]
  17. Burgess, R.J.; Zhang, Z. Histones, histone chaperones and nucleosome assembly. Protein Cell 2010, 1, 607–612. [Google Scholar] [CrossRef] [Green Version]
  18. Gill, J.; Kumar, A.; Sharma, A. Structural comparisons reveal diverse binding modes between nucleosome assembly proteins and histones. Epigenet. Chromatin 2022, 15, 20. [Google Scholar] [CrossRef]
  19. Sauer, P.V.; Gu, Y.; Liu, W.H.; Mattiroli, F.; Panne, D.; Luger, K.; Churchill, M.E. Mechanistic insights into histone deposition and nucleosome assembly by the chromatin assembly factor-1. Nucleic Acids Res. 2018, 46, 9907–9917. [Google Scholar] [CrossRef] [Green Version]
  20. Groth, A.; Rocha, W.; Verreault, A.; Almouzni, G. Chromatin challenges during DNA replication and repair. Cell 2007, 128, 721–733. [Google Scholar] [CrossRef] [Green Version]
  21. Burgess, R.J.; Zhang, Z. Histone chaperones in nucleosome assembly and human disease. Nat. Struct. Mol. Biol. 2013, 20, 14–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yang, J.; Zhang, X.; Feng, J.; Leng, H.; Li, S.; Xiao, J.; Liu, S.; Xu, Z.; Xu, J.; Li, D.; et al. The Histone Chaperone FACT Contributes to DNA Replication-Coupled Nucleosome Assembly. Cell Rep. 2016, 16, 3414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Das, C.; Tyler, J.K.; Churchill, M.E. The histone shuffle: Histone chaperones in an energetic dance. Trends Biochem. Sci. 2010, 35, 476–489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Alabert, C.; Groth, A. Chromatin replication and epigenome maintenance. Nat. Rev. Mol. Cell Biol. 2012, 13, 153–167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Annunziato, A.T. Assembling chromatin: The long and winding road. Biochim. Biophys. Acta 2013, 1819, 196–210. [Google Scholar] [CrossRef]
  26. Gurard-Levin, Z.A.; Quivy, J.P.; Almouzni, G. Histone chaperones: Assisting histone traffic and nucleosome dynamics. Annu. Rev. Biochem. 2014, 83, 487–517. [Google Scholar] [CrossRef]
  27. Osley, M.A. The regulation of histone synthesis in the cell cycle. Annu. Rev. Biochem. 1991, 60, 827–861. [Google Scholar] [CrossRef]
  28. Marzluff, W.F.; Duronio, R.J. Histone mRNA expression: Multiple levels of cell cycle regulation and important developmental consequences. Curr. Opin. Cell Biol. 2002, 14, 692–699. [Google Scholar] [CrossRef]
  29. Ling, X.; Harkness, T.A.; Schultz, M.C.; Fisher-Adams, G.; Grunstein, M. Yeast histone H3 and H4 amino termini are important for nucleosome assembly in vivo and in vitro: Redundant and position-independent functions in assembly but not in gene regulation. Genes Dev. 1996, 10, 686–699. [Google Scholar] [CrossRef] [Green Version]
  30. Yue, Y.; Yang, W.S.; Zhang, L.; Liu, C.P.; Xu, R.M. Topography of histone H3-H4 interaction with the Hat1-Hat2 acetyltransferase complex. Genes Dev. 2022, 36, 408–413. [Google Scholar] [CrossRef]
  31. Fillingham, J.; Recht, J.; Silva, A.C.; Suter, B.; Emili, A.; Stagljar, I.; Krogan, N.J.; Allis, C.D.; Keogh, M.C.; Greenblatt, J.F. Chaperone control of the activity and specificity of the histone H3 acetyltransferase Rtt109. Mol. Cell. Biol. 2008, 28, 4342–4353. [Google Scholar] [CrossRef] [Green Version]
  32. Burgess, R.J.; Zhou, H.; Han, J.; Zhang, Z. A role for Gcn5 in replication-coupled nucleosome assembly. Mol. Cell 2010, 37, 469–480. [Google Scholar] [CrossRef] [Green Version]
  33. Ye, J.; Ai, X.; Eugeni, E.E.; Zhang, L.; Carpenter, L.R.; Jelinek, M.A.; Freitas, M.A.; Parthun, M.R. Histone H4 lysine 91 acetylation a core domain modification associated with chromatin assembly. Mol. Cell 2005, 18, 123–130. [Google Scholar] [CrossRef] [Green Version]
  34. Han, J.; Zhou, H.; Li, Z.; Xu, R.M.; Zhang, Z. Acetylation of lysine 56 of histone H3 catalyzed by RTT109 and regulated by ASF1 is required for replisome integrity. J. Biol. Chem. 2007, 282, 28587–28596. [Google Scholar] [CrossRef] [Green Version]
  35. Han, J.; Zhou, H.; Horazdovsky, B.; Zhang, K.; Xu, R.M.; Zhang, Z. Rtt109 acetylates histone H3 lysine 56 and functions in DNA replication. Science 2007, 315, 653–655. [Google Scholar] [CrossRef]
  36. Jasencakova, Z.; Scharf, A.N.; Ask, K.; Corpet, A.; Imhof, A.; Almouzni, G.; Groth, A. Replication stress interferes with histone recycling and predeposition marking of new histones. Mol. Cell 2010, 37, 736–743. [Google Scholar] [CrossRef]
  37. Tsubota, T.; Berndsen, C.E.; Erkmann, J.A.; Smith, C.L.; Yang, L.; Freitas, M.A.; Denu, J.M.; Kaufman, P.D. Histone H3-K56 acetylation is catalyzed by histone chaperone-dependent complexes. Mol. Cell 2007, 25, 703–712. [Google Scholar] [CrossRef] [Green Version]
  38. Li, Q.; Zhou, H.; Wurtele, H.; Davies, B.; Horazdovsky, B.; Verreault, A.; Zhang, Z. Acetylation of histone H3 lysine 56 regulates replication-coupled nucleosome assembly. Cell 2008, 134, 244–255. [Google Scholar] [CrossRef] [Green Version]
  39. Han, J.; Zhang, H.; Zhang, H.; Wang, Z.; Zhou, H.; Zhang, Z. A Cul4 E3 ubiquitin ligase regulates histone hand-off during nucleosome assembly. Cell 2013, 155, 817–829. [Google Scholar] [CrossRef] [Green Version]
  40. Rodriges Blanko, E.; Kadyrova, L.Y.; Kadyrov, F.A. DNA Mismatch Repair Interacts with CAF-1- and ASF1A-H3-H4-dependent Histone (H3-H4)2 Tetramer Deposition. J. Biol. Chem. 2016, 291, 9203–9217. [Google Scholar] [CrossRef] [Green Version]
  41. Liu, W.H.; Roemer, S.C.; Zhou, Y.; Shen, Z.J.; Dennehey, B.K.; Balsbaugh, J.L.; Liddle, J.C.; Nemkov, T.; Ahn, N.G.; Hansen, K.C.; et al. The Cac1 subunit of histone chaperone CAF-1 organizes CAF-1-H3/H4 architecture and tetramerizes histones. eLife 2016, 5, e18023. [Google Scholar] [CrossRef] [PubMed]
  42. Kim, D.; Setiaputra, D.; Jung, T.; Chung, J.; Leitner, A.; Yoon, J.; Aebersold, R.; Hebert, H.; Yip, C.K.; Song, J.J. Molecular Architecture of Yeast Chromatin Assembly Factor 1. Sci. Rep. 2016, 6, 26702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zhang, K.; Gao, Y.; Li, J.; Burgess, R.; Han, J.; Liang, H.; Zhang, Z.; Liu, Y. A DNA binding winged helix domain in CAF-1 functions with PCNA to stabilize CAF-1 at replication forks. Nucleic Acids Res. 2016, 44, 5083–5094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Rolef Ben-Shahar, T.; Castillo, A.G.; Osborne, M.J.; Borden, K.L.; Kornblatt, J.; Verreault, A. Two fundamentally distinct PCNA interaction peptides contribute to chromatin assembly factor 1 function. Mol. Cell. Biol. 2009, 29, 6353–6365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Liu, W.H.; Roemer, S.C.; Port, A.M.; Churchill, M.E.A. CAF-1-induced oligomerization of histones H3/H4 and mutually exclusive interactions with Asf1 guide H3/H4 transitions among histone chaperones and DNA. Nucleic Acids Res. 2017, 45, 9809. [Google Scholar] [CrossRef] [Green Version]
  46. Soniat, M.; Cagatay, T.; Chook, Y.M. Recognition Elements in the Histone H3 and H4 Tails for Seven Different Importins. J. Biol. Chem. 2016, 291, 21171–21183. [Google Scholar] [CrossRef] [Green Version]
  47. Su, D.; Hu, Q.; Li, Q.; Thompson, J.R.; Cui, G.; Fazly, A.; Davies, B.A.; Botuyan, M.V.; Zhang, Z.; Mer, G. Structural basis for recognition of H3K56-acetylated histone H3-H4 by the chaperone Rtt106. Nature 2012, 483, 104–107. [Google Scholar] [CrossRef] [Green Version]
  48. Fazly, A.; Li, Q.; Hu, Q.; Mer, G.; Horazdovsky, B.; Zhang, Z. Histone chaperone Rtt106 promotes nucleosome formation using (H3-H4)2 tetramers. J. Biol. Chem. 2012, 287, 10753–10760. [Google Scholar] [CrossRef] [Green Version]
  49. Piquet, S.; Le Parc, F.; Bai, S.K.; Chevallier, O.; Adam, S.; Polo, S.E. The Histone Chaperone FACT Coordinates H2A.X-Dependent Signaling and Repair of DNA Damage. Mol. Cell 2018, 72, 888–901.e7. [Google Scholar] [CrossRef] [Green Version]
  50. Marechal, A.; Zou, L. RPA-coated single-stranded DNA as a platform for post-translational modifications in the DNA damage response. Cell Res. 2015, 25, 9–23. [Google Scholar] [CrossRef] [Green Version]
  51. Liu, S.; Xu, Z.; Leng, H.; Zheng, P.; Yang, J.; Chen, K.; Feng, J.; Li, Q. RPA binds histone H3-H4 and functions in DNA replication-coupled nucleosome assembly. Science 2017, 355, 415–420. [Google Scholar] [CrossRef]
  52. Osakabe, A.; Tachiwana, H.; Matsunaga, T.; Shiga, T.; Nozawa, R.S.; Obuse, C.; Kurumizaka, H. Nucleosome formation activity of human somatic nuclear autoantigenic sperm protein (sNASP). J. Biol. Chem. 2010, 285, 11913–11921. [Google Scholar] [CrossRef] [Green Version]
  53. Apta-Smith, M.J.; Hernandez-Fernaud, J.R.; Bowman, A.J. Evidence for the nuclear import of histones H3.1 and H4 as monomers. EMBO J. 2018, 37, e98714. [Google Scholar] [CrossRef]
  54. Ask, K.; Jasencakova, Z.; Menard, P.; Feng, Y.; Almouzni, G.; Groth, A. Codanin-1, mutated in the anaemic disease CDAI, regulates Asf1 function in S-phase histone supply. EMBO J. 2012, 31, 2013–2023. [Google Scholar] [CrossRef]
  55. Tang, Y.; Holbert, M.A.; Wurtele, H.; Meeth, K.; Rocha, W.; Gharib, M.; Jiang, E.; Thibault, P.; Verreault, A.; Cole, P.A.; et al. Fungal Rtt109 histone acetyltransferase is an unexpected structural homolog of metazoan p300/CBP. Nat. Struct. Mol. Biol. 2008, 15, 738–745. [Google Scholar] [CrossRef]
  56. Saxena, S.; Zou, L. Hallmarks of DNA replication stress. Mol. Cell 2022, 82, 2298–2314. [Google Scholar] [CrossRef]
  57. Ashour, M.E.; Mosammaparast, N. Mechanisms of damage tolerance and repair during DNA replication. Nucleic Acids Res. 2021, 49, 3033–3047. [Google Scholar] [CrossRef]
  58. Uckelmann, M.; Sixma, T.K. Histone ubiquitination in the DNA damage response. DNA Repair 2017, 56, 92–101. [Google Scholar] [CrossRef]
  59. Cortez, D. Replication-Coupled DNA Repair. Mol. Cell 2019, 74, 866–876. [Google Scholar] [CrossRef] [Green Version]
  60. Prado, F. Homologous recombination maintenance of genome integrity during DNA damage tolerance. Mol. Cell Oncol. 2014, 1, e957039. [Google Scholar] [CrossRef] [Green Version]
  61. Sale, J.E.; Lehmann, A.R.; Woodgate, R. Y-family DNA polymerases and their role in tolerance of cellular DNA damage. Nat. Rev. Mol. Cell Biol. 2012, 13, 141–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Sharma, A.K.; Hendzel, M.J. The relationship between histone posttranslational modification and DNA damage signaling and repair. Int. J. Radiat. Biol. 2019, 95, 382–393. [Google Scholar] [CrossRef] [PubMed]
  63. Du, W.; Shi, G.; Shan, C.M.; Li, Z.; Zhu, B.; Jia, S.; Li, Q.; Zhang, Z. Mechanisms of chromatin-based epigenetic inheritance. Sci. China Life Sci. 2022, 65, 2162–2190. [Google Scholar] [CrossRef] [PubMed]
  64. Burgers, P.M.J.; Kunkel, T.A. Eukaryotic DNA Replication Fork. Annu. Rev. Biochem. 2017, 86, 417–438. [Google Scholar] [CrossRef]
  65. Guilliam, T.A.; Yeeles, J.T.P. An updated perspective on the polymerase division of labor during eukaryotic DNA replication. Crit. Rev. Biochem. Mol. Biol. 2020, 55, 469–481. [Google Scholar] [CrossRef]
  66. Buser, R.; Kellner, V.; Melnik, A.; Wilson-Zbinden, C.; Schellhaas, R.; Kastner, L.; Piwko, W.; Dees, M.; Picotti, P.; Maric, M.; et al. The Replisome-Coupled E3 Ubiquitin Ligase Rtt101Mms22 Counteracts Mrc1 Function to Tolerate Genotoxic Stress. PLoS Genet. 2016, 12, e1005843. [Google Scholar] [CrossRef] [Green Version]
  67. Piwko, W.; Mlejnkova, L.J.; Mutreja, K.; Ranjha, L.; Stafa, D.; Smirnov, A.; Brodersen, M.M.; Zellweger, R.; Sturzenegger, A.; Janscak, P.; et al. The MMS22L-TONSL heterodimer directly promotes RAD51-dependent recombination upon replication stress. EMBO J. 2016, 35, 2584–2601. [Google Scholar] [CrossRef] [Green Version]
  68. Daboussi, F.; Courbet, S.; Benhamou, S.; Kannouche, P.; Zdzienicka, M.Z.; Debatisse, M.; Lopez, B.S. A homologous recombination defect affects replication-fork progression in mammalian cells. J. Cell Sci. 2008, 121, 162–166. [Google Scholar] [CrossRef] [Green Version]
  69. Tjeertes, J.V.; Miller, K.M.; Jackson, S.P. Screen for DNA-damage-responsive histone modifications identifies H3K9Ac and H3K56Ac in human cells. EMBO J. 2009, 28, 1878–1889. [Google Scholar] [CrossRef] [Green Version]
  70. Saredi, G.; Huang, H.; Hammond, C.M.; Alabert, C.; Bekker-Jensen, S.; Forne, I.; Reveron-Gomez, N.; Foster, B.M.; Mlejnkova, L.; Bartke, T.; et al. H4K20me0 marks post-replicative chromatin and recruits the TONSL-MMS22L DNA repair complex. Nature 2016, 534, 714–718. [Google Scholar] [CrossRef] [Green Version]
  71. O’Donnell, L.; Panier, S.; Wildenhain, J.; Tkach, J.M.; Al-Hakim, A.; Landry, M.C.; Escribano-Diaz, C.; Szilard, R.K.; Young, J.T.; Munro, M.; et al. The MMS22L-TONSL complex mediates recovery from replication stress and homologous recombination. Mol. Cell 2010, 40, 619–631. [Google Scholar] [CrossRef] [Green Version]
  72. Ozdemir, A.; Spicuglia, S.; Lasonder, E.; Vermeulen, M.; Campsteijn, C.; Stunnenberg, H.G.; Logie, C. Characterization of lysine 56 of histone H3 as an acetylation site in Saccharomyces cerevisiae. J. Biol. Chem. 2005, 280, 25949–25952. [Google Scholar] [CrossRef] [Green Version]
  73. Maas, N.L.; Miller, K.M.; DeFazio, L.G.; Toczyski, D.P. Cell cycle and checkpoint regulation of histone H3 K56 acetylation by Hst3 and Hst4. Mol. Cell 2006, 23, 109–119. [Google Scholar] [CrossRef]
  74. Masumoto, H.; Hawke, D.; Kobayashi, R.; Verreault, A. A role for cell-cycle-regulated histone H3 lysine 56 acetylation in the DNA damage response. Nature 2005, 436, 294–298. [Google Scholar] [CrossRef]
  75. Collins, S.R.; Miller, K.M.; Maas, N.L.; Roguev, A.; Fillingham, J.; Chu, C.S.; Schuldiner, M.; Gebbia, M.; Recht, J.; Shales, M.; et al. Functional dissection of protein complexes involved in yeast chromosome biology using a genetic interaction map. Nature 2007, 446, 806–810. [Google Scholar] [CrossRef]
  76. Wurtele, H.; Kaiser, G.S.; Bacal, J.; St-Hilaire, E.; Lee, E.H.; Tsao, S.; Dorn, J.; Maddox, P.; Lisby, M.; Pasero, P.; et al. Histone H3 lysine 56 acetylation and the response to DNA replication fork damage. Mol. Cell. Biol. 2012, 32, 154–172. [Google Scholar] [CrossRef] [Green Version]
  77. Tessarz, P.; Kouzarides, T. Histone core modifications regulating nucleosome structure and dynamics. Nat. Rev. Mol. Cell Biol. 2014, 15, 703–708. [Google Scholar] [CrossRef]
  78. Clemente-Ruiz, M.; Gonzalez-Prieto, R.; Prado, F. Histone H3K56 acetylation, CAF1, and Rtt106 coordinate nucleosome assembly and stability of advancing replication forks. PLoS Genet. 2011, 7, e1002376. [Google Scholar] [CrossRef] [Green Version]
  79. Luke, B.; Versini, G.; Jaquenoud, M.; Zaidi, I.W.; Kurz, T.; Pintard, L.; Pasero, P.; Peter, M. The cullin Rtt101p promotes replication fork progression through damaged DNA and natural pause sites. Curr. Biol. 2006, 16, 786–792. [Google Scholar] [CrossRef] [Green Version]
  80. Duro, E.; Vaisica, J.A.; Brown, G.W.; Rouse, J. Budding yeast Mms22 and Mms1 regulate homologous recombination induced by replisome blockage. DNA Repair 2008, 7, 811–818. [Google Scholar] [CrossRef]
  81. Erkmann, J.A.; Kaufman, P.D. A negatively charged residue in place of histone H3K56 supports chromatin assembly factor association but not genotoxic stress resistance. DNA Repair 2009, 8, 1371–1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Han, J.; Li, Q.; McCullough, L.; Kettelkamp, C.; Formosa, T.; Zhang, Z. Ubiquitylation of FACT by the cullin-E3 ligase Rtt101 connects FACT to DNA replication. Genes Dev. 2010, 24, 1485–1490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. 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] [PubMed]
  84. Hu, J.; McCall, C.M.; Ohta, T.; Xiong, Y. Targeted ubiquitination of CDT1 by the DDB1-CUL4A-ROC1 ligase in response to DNA damage. Nat. Cell Biol. 2004, 6, 1003–1009. [Google Scholar] [CrossRef]
  85. Formosa, T. The role of FACT in making and breaking nucleosomes. Biochim. Biophys. Acta 2013, 1819, 247–255. [Google Scholar] [CrossRef] [Green Version]
  86. Gambus, A.; Jones, R.C.; Sanchez-Diaz, A.; Kanemaki, M.; van Deursen, F.; Edmondson, R.D.; Labib, K. GINS maintains association of Cdc45 with MCM in replisome progression complexes at eukaryotic DNA replication forks. Nat. Cell Biol. 2006, 8, 358–366. [Google Scholar] [CrossRef]
  87. Murawska, M.; Schauer, T.; Matsuda, A.; Wilson, M.D.; Pysik, T.; Wojcik, F.; Muir, T.W.; Hiraoka, Y.; Straub, T.; Ladurner, A.G. The Chaperone FACT and Histone H2B Ubiquitination Maintain S. pombe Genome Architecture through Genic and Subtelomeric Functions. Mol. Cell 2020, 77, 501–513.e7. [Google Scholar] [CrossRef]
  88. Moyal, L.; Lerenthal, Y.; Gana-Weisz, M.; Mass, G.; So, S.; Wang, S.Y.; Eppink, B.; Chung, Y.M.; Shalev, G.; Shema, E.; et al. Requirement of ATM-dependent monoubiquitylation of histone H2B for timely repair of DNA double-strand breaks. Mol. Cell 2011, 41, 529–542. [Google Scholar] [CrossRef]
  89. Chandrasekharan, M.B.; Huang, F.; Sun, Z.W. Ubiquitination of histone H2B regulates chromatin dynamics by enhancing nucleosome stability. Proc. Natl. Acad. Sci. USA 2009, 106, 16686–16691. [Google Scholar] [CrossRef] [Green Version]
  90. Fleming, A.B.; Kao, C.F.; Hillyer, C.; Pikaart, M.; Osley, M.A. H2B ubiquitylation plays a role in nucleosome dynamics during transcription elongation. Mol. Cell 2008, 31, 57–66. [Google Scholar] [CrossRef]
  91. Wang, P.; Yang, W.; Zhao, S.; Nashun, B. Regulation of chromatin structure and function: Insights into the histone chaperone FACT. Cell Cycle 2021, 20, 465–479. [Google Scholar] [CrossRef]
  92. Le, S.; Davis, C.; Konopka, J.B.; Sternglanz, R. Two new S-phase-specific genes from Saccharomyces cerevisiae. Yeast 1997, 13, 1029–1042. [Google Scholar] [CrossRef]
  93. Tyler, J.K.; Adams, C.R.; Chen, S.R.; Kobayashi, R.; Kamakaka, R.T.; Kadonaga, J.T. The RCAF complex mediates chromatin assembly during DNA replication and repair. Nature 1999, 402, 555–560. [Google Scholar] [CrossRef]
  94. Groth, A.; Corpet, A.; Cook, A.J.; Roche, D.; Bartek, J.; Lukas, J.; Almouzni, G. Regulation of replication fork progression through histone supply and demand. Science 2007, 318, 1928–1931. [Google Scholar] [CrossRef]
  95. Sanematsu, F.; Takami, Y.; Barman, H.K.; Fukagawa, T.; Ono, T.; Shibahara, K.I.; Nakayama, T. Asf1 is required for viability and chromatin assembly during DNA replication in vertebrate cells. J. Biol. Chem. 2006, 281, 13817–13827. [Google Scholar] [CrossRef] [Green Version]
  96. Pietrobon, V.; Freon, K.; Hardy, J.; Costes, A.; Iraqui, I.; Ochsenbein, F.; Lambert, S.A. The chromatin assembly factor 1 promotes Rad51-dependent template switches at replication forks by counteracting D-loop disassembly by the RecQ-type helicase Rqh1. PLoS Biol. 2014, 12, e1001968. [Google Scholar] [CrossRef] [Green Version]
  97. Jiao, R.; Bachrati, C.Z.; Pedrazzi, G.; Kuster, P.; Petkovic, M.; Li, J.L.; Egli, D.; Hickson, I.D.; Stagljar, I. Physical and functional interaction between the Bloom’s syndrome gene product and the largest subunit of chromatin assembly factor 1. Mol. Cell. Biol. 2004, 24, 4710–4719. [Google Scholar] [CrossRef] [Green Version]
  98. Lazo, P.A. Targeting Histone Epigenetic Modifications and DNA Damage Responses in Synthetic Lethality Strategies in Cancer? Cancers 2022, 14, 4050. [Google Scholar] [CrossRef]
  99. Kouzarides, T. Chromatin modifications and their function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [Green Version]
  100. Nacev, B.A.; Feng, L.; Bagert, J.D.; Lemiesz, A.E.; Gao, J.; Soshnev, A.A.; Kundra, R.; Schultz, N.; Muir, T.W.; Allis, C.D. The expanding landscape of ‘oncohistone’ mutations in human cancers. Nature 2019, 567, 473–478. [Google Scholar] [CrossRef]
  101. Espiritu, D.; Gribkova, A.K.; Gupta, S.; Shaytan, A.K.; Panchenko, A.R. Molecular Mechanisms of Oncogenesis through the Lens of Nucleosomes and Histones. J. Phys. Chem. B 2021, 125, 3963–3976. [Google Scholar] [CrossRef] [PubMed]
  102. Neganova, M.E.; Klochkov, S.G.; Aleksandrova, Y.R.; Aliev, G. Histone modifications in epigenetic regulation of cancer: Perspectives and achieved progress. Semin. Cancer Biol. 2022, 83, 452–471. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, Y.; Sun, Z.; Jia, J.; Du, T.; Zhang, N.; Tang, Y.; Fang, Y.; Fang, D. Overview of Histone Modification. In Histone Mutations and Cancer; Advances in Experimental Medicine and Biology; Springer: Singapore, 2021; Volume 1283, pp. 1–16. [Google Scholar] [CrossRef] [PubMed]
  104. McCleland, M.L.; Soukup, T.M.; Liu, S.D.; Esensten, J.H.; de Sousa e Melo, F.; Yaylaoglu, M.; Warming, S.; Roose-Girma, M.; Firestein, R. Cdk8 deletion in the Apc(Min) murine tumour model represses EZH2 activity and accelerates tumourigenesis. J. Pathol. 2015, 237, 508–519. [Google Scholar] [CrossRef]
  105. Mohammad, F.; Helin, K. Oncohistones: Drivers of pediatric cancers. Genes Dev. 2017, 31, 2313–2324. [Google Scholar] [CrossRef] [Green Version]
  106. Gsell, C.; Richly, H.; Coin, F.; Naegeli, H. A chromatin scaffold for DNA damage recognition: How histone methyltransferases prime nucleosomes for repair of ultraviolet light-induced lesions. Nucleic Acids Res. 2020, 48, 1652–1668. [Google Scholar] [CrossRef] [Green Version]
  107. Ji, H.; Zhou, Y.; Zhuang, X.; Zhu, Y.; Wu, Z.; Lu, Y.; Li, S.; Zeng, Y.; Lu, Q.R.; Huo, Y.; et al. HDAC3 Deficiency Promotes Liver Cancer through a Defect in H3K9ac/H3K9me3 Transition. Cancer Res. 2019, 79, 3676–3688. [Google Scholar] [CrossRef]
  108. Michalak, E.M.; Burr, M.L.; Bannister, A.J.; Dawson, M.A. The roles of DNA, RNA and histone methylation in ageing and cancer. Nat. Rev. Mol. Cell Biol. 2019, 20, 573–589. [Google Scholar] [CrossRef]
  109. Duan, R.; Du, W.; Guo, W. EZH2: A novel target for cancer treatment. J. Hematol. Oncol. 2020, 13, 104. [Google Scholar] [CrossRef]
  110. Lima-Fernandes, E.; Murison, A.; da Silva Medina, T.; Wang, Y.; Ma, A.; Leung, C.; Luciani, G.M.; Haynes, J.; Pollett, A.; Zeller, C.; et al. Targeting bivalency de-represses Indian Hedgehog and inhibits self-renewal of colorectal cancer-initiating cells. Nat. Commun. 2019, 10, 1436. [Google Scholar] [CrossRef] [Green Version]
  111. Ntziachristos, P.; Tsirigos, A.; Van Vlierberghe, P.; Nedjic, J.; Trimarchi, T.; Flaherty, M.S.; Ferres-Marco, D.; da Ros, V.; Tang, Z.; Siegle, J.; et al. Genetic inactivation of the polycomb repressive complex 2 in T cell acute lymphoblastic leukemia. Nat. Med. 2012, 18, 298–301. [Google Scholar] [CrossRef]
  112. Shimizu, T.; Kubovcakova, L.; Nienhold, R.; Zmajkovic, J.; Meyer, S.C.; Hao-Shen, H.; Geier, F.; Dirnhofer, S.; Guglielmelli, P.; Vannucchi, A.M.; et al. Loss of Ezh2 synergizes with JAK2-V617F in initiating myeloproliferative neoplasms and promoting myelofibrosis. J. Exp. Med. 2016, 213, 1479–1496. [Google Scholar] [CrossRef] [Green Version]
  113. Huang, S.; Wang, Z.; Zhou, J.; Huang, J.; Zhou, L.; Luo, J.; Wan, Y.Y.; Long, H.; Zhu, B. EZH2 Inhibitor GSK126 Suppresses Antitumor Immunity by Driving Production of Myeloid-Derived Suppressor Cells. Cancer Res. 2019, 79, 2009–2020. [Google Scholar] [CrossRef]
  114. Fang, L.; Teng, H.; Wang, Y.; Liao, G.; Weng, L.; Li, Y.; Wang, X.; Jin, J.; Jiao, C.; Chen, L.; et al. SET1A-Mediated Mono-Methylation at K342 Regulates YAP Activation by Blocking Its Nuclear Export and Promotes Tumorigenesis. Cancer Cell 2018, 34, 103–118.e9. [Google Scholar] [CrossRef] [Green Version]
  115. Wang, R.; Liu, J.; Li, K.; Yang, G.; Chen, S.; Wu, J.; Xie, X.; Ren, H.; Pang, Y. An SETD1A/Wnt/beta-catenin feedback loop promotes NSCLC development. J. Exp. Clin. Cancer Res. 2021, 40, 318. [Google Scholar] [CrossRef]
  116. Wu, J.; Chai, H.; Xu, X.; Yu, J.; Gu, Y. Histone methyltransferase SETD1A interacts with HIF1alpha to enhance glycolysis and promote cancer progression in gastric cancer. Mol. Oncol. 2020, 14, 1397–1409. [Google Scholar] [CrossRef] [Green Version]
  117. Cheng, J.; Demeulemeester, J.; Wedge, D.C.; Vollan, H.K.M.; Pitt, J.J.; Russnes, H.G.; Pandey, B.P.; Nilsen, G.; Nord, S.; Bignell, G.R.; et al. Pan-cancer analysis of homozygous deletions in primary tumours uncovers rare tumour suppressors. Nat. Commun. 2017, 8, 1221. [Google Scholar] [CrossRef] [Green Version]
  118. Hu, A.; Hong, F.; Li, D.; Jin, Y.; Kon, L.; Xu, Z.; He, H.; Xie, Q. Long non-coding RNA ROR recruits histone transmethylase MLL1 to up-regulate TIMP3 expression and promote breast cancer progression. J. Transl. Med. 2021, 19, 95. [Google Scholar] [CrossRef]
  119. Riedel, S.S.; Haladyna, J.N.; Bezzant, M.; Stevens, B.; Pollyea, D.A.; Sinha, A.U.; Armstrong, S.A.; Wei, Q.; Pollock, R.M.; Daigle, S.R.; et al. MLL1 and DOT1L cooperate with meningioma-1 to induce acute myeloid leukemia. J. Clin. Investig. 2016, 126, 1438–1450. [Google Scholar] [CrossRef] [Green Version]
  120. Qiang, R.; Cai, N.; Wang, X.; Wang, L.; Cui, K.; Wang, X.; Li, X. MLL1 promotes cervical carcinoma cell tumorigenesis and metastasis through interaction with β-catenin. OncoTargets Ther. 2016, 9, 6631–6640. [Google Scholar] [CrossRef] [Green Version]
  121. Yang, Z.; Li, C.; Fan, Z.; Liu, H.; Zhang, X.; Cai, Z.; Xu, L.; Luo, J.; Huang, Y.; He, L.; et al. Single-cell Sequencing Reveals Variants in ARID1A, GPRC5A and MLL2 Driving Self-renewal of Human Bladder Cancer Stem Cells. Eur. Urol. 2017, 71, 8–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Grasso, C.S.; Wu, Y.M.; Robinson, D.R.; Cao, X.; Dhanasekaran, S.M.; Khan, A.P.; Quist, M.J.; Jing, X.; Lonigro, R.J.; Brenner, J.C.; et al. The mutational landscape of lethal castration-resistant prostate cancer. Nature 2012, 487, 239–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Bian, S.; Wang, Z.; Chen, Y.; Li, R. SPLUNC1 and MLL3 regulate cancer stem cells in nasopharyngeal carcinoma. J. BUON 2019, 24, 1700–1705. [Google Scholar] [PubMed]
  124. Kim, S.S.; Lee, M.H.; Lee, M.O. Histone methyltransferases regulate the transcriptional expression of ERalpha and the proliferation of tamoxifen-resistant breast cancer cells. Breast Cancer Res. Treat. 2020, 180, 45–54. [Google Scholar] [CrossRef] [Green Version]
  125. Liu, Y.; Li, J.; Yue, B.; Liang, L.; Zhang, S.; Chen, Y. Long non-coding RNA DANCR regulate MLL3 and thereby it determines the progression of pancreatic cancer. J. BUON 2020, 25, 1954–1959. [Google Scholar]
  126. Larsson, C.; Cordeddu, L.; Siggens, L.; Pandzic, T.; Kundu, S.; He, L.; Ali, M.A.; Pristovsek, N.; Hartman, K.; Ekwall, K.; et al. Restoration of KMT2C/MLL3 in human colorectal cancer cells reinforces genome-wide H3K4me1 profiles and influences cell growth and gene expression. Clin. Epigenet. 2020, 12, 74. [Google Scholar] [CrossRef]
  127. Alam, H.; Tang, M.; Maitituoheti, M.; Dhar, S.S.; Kumar, M.; Han, C.Y.; Ambati, C.R.; Amin, S.B.; Gu, B.; Chen, T.Y.; et al. KMT2D Deficiency Impairs Super-Enhancers to Confer a Glycolytic Vulnerability in Lung Cancer. Cancer Cell 2020, 37, 599–617.e7. [Google Scholar] [CrossRef]
  128. Dhar, S.S.; Zhao, D.; Lin, T.; Gu, B.; Pal, K.; Wu, S.J.; Alam, H.; Lv, J.; Yun, K.; Gopalakrishnan, V.; et al. MLL4 Is Required to Maintain Broad H3K4me3 Peaks and Super-Enhancers at Tumor Suppressor Genes. Mol. Cell 2018, 70, 825–841.e6. [Google Scholar] [CrossRef] [Green Version]
  129. Wu, L.; Kou, F.; Ji, Z.; Li, B.; Zhang, B.; Guo, Y.; Yang, L. SMYD2 promotes tumorigenesis and metastasis of lung adenocarcinoma through RPS7. Cell Death Dis. 2021, 12, 439. [Google Scholar] [CrossRef]
  130. Wang, Y.; Jin, G.; Guo, Y.; Cao, Y.; Niu, S.; Fan, X.; Zhang, J. SMYD2 suppresses p53 activity to promote glucose metabolism in cervical cancer. Exp. Cell Res. 2021, 404, 112649. [Google Scholar] [CrossRef]
  131. Li, C.; Feng, S.Y.; Chen, L. SET7/9 promotes H3K4me3 at lncRNA DRAIC promoter to modulate growth and metastasis of glioma. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 12241–12250. [Google Scholar] [CrossRef]
  132. Zhang, S.L.; Du, X.; Tan, L.N.; Deng, F.H.; Zhou, B.Y.; Zhou, H.J.; Zhu, H.Y.; Chu, Y.; Liu, D.L.; Tan, Y.Y. SET7 interacts with HDAC6 and suppresses the development of colon cancer through inactivation of HDAC6. Am. J. Transl. Res. 2020, 12, 602–611. [Google Scholar]
  133. Daks, A.; Mamontova, V.; Fedorova, O.; Petukhov, A.; Shuvalov, O.; Parfenyev, S.; Netsvetay, S.; Venina, A.; Kizenko, A.; Imyanitov, E.; et al. Set7/9 controls proliferation and genotoxic drug resistance of NSCLC cells. Biochem. Biophys. Res. Commun. 2021, 572, 41–48. [Google Scholar] [CrossRef]
  134. Montenegro, M.F.; Sanchez-Del-Campo, L.; Gonzalez-Guerrero, R.; Martinez-Barba, E.; Pinero-Madrona, A.; Cabezas-Herrera, J.; Rodriguez-Lopez, J.N. Tumor suppressor SET9 guides the epigenetic plasticity of breast cancer cells and serves as an early-stage biomarker for predicting metastasis. Oncogene 2016, 35, 6143–6152. [Google Scholar] [CrossRef]
  135. Li, J.; Zhao, L.; Pan, Y.; Ma, X.; Liu, L.; Wang, W.; You, W. SMYD3 overexpression indicates poor prognosis and promotes cell proliferation, migration and invasion in non-small cell lung cancer. Int. J. Oncol. 2020, 57, 756–766. [Google Scholar] [CrossRef]
  136. Zhu, C.L.; Huang, Q. Overexpression of the SMYD3 Promotes Proliferation, Migration, and Invasion of Pancreatic Cancer. Dig. Dis. Sci. 2020, 65, 489–499. [Google Scholar] [CrossRef]
  137. Fenizia, C.; Bottino, C.; Corbetta, S.; Fittipaldi, R.; Floris, P.; Gaudenzi, G.; Carra, S.; Cotelli, F.; Vitale, G.; Caretti, G. SMYD3 promotes the epithelial-mesenchymal transition in breast cancer. Nucleic Acids Res. 2019, 47, 1278–1293. [Google Scholar] [CrossRef] [Green Version]
  138. Zhang, L.; Tian, S.; Zhao, M.; Yang, T.; Quan, S.; Yang, Q.; Song, L.; Yang, X. SUV39H1-DNMT3A-mediated epigenetic regulation of Tim-3 and galectin-9 in the cervical cancer. Cancer Cell Int. 2020, 20, 325. [Google Scholar] [CrossRef]
  139. Yu, T.; Wang, C.; Yang, J.; Guo, Y.; Wu, Y.; Li, X. Metformin inhibits SUV39H1-mediated migration of prostate cancer cells. Oncogenesis 2017, 6, e324. [Google Scholar] [CrossRef] [Green Version]
  140. Kim, G.; Kim, J.Y.; Lim, S.C.; Lee, K.Y.; Kim, O.; Choi, H.S. SUV39H1/DNMT3A-dependent methylation of the RB1 promoter stimulates PIN1 expression and melanoma development. FASEB J. 2018, 32, 5647–5660. [Google Scholar] [CrossRef]
  141. Shuai, W.; Wu, J.; Chen, S.; Liu, R.; Ye, Z.; Kuang, C.; Fu, X.; Wang, G.; Li, Y.; Peng, Q.; et al. SUV39H2 promotes colorectal cancer proliferation and metastasis via tri-methylation of the SLIT1 promoter. Cancer Lett. 2018, 422, 56–69. [Google Scholar] [CrossRef] [PubMed]
  142. Miao, Y.; Liu, G.; Liu, L. Histone methyltransferase SUV39H2 regulates LSD1-dependent CDH1 expression and promotes epithelial mesenchymal transition of osteosarcoma. Cancer Cell Int. 2021, 21, 2. [Google Scholar] [CrossRef] [PubMed]
  143. Zheng, Y.; Li, B.; Wang, J.; Xiong, Y.; Wang, K.; Qi, Y.; Sun, H.; Wu, L.; Yang, L. Identification of SUV39H2 as a potential oncogene in lung adenocarcinoma. Clin. Epigenet. 2018, 10, 129. [Google Scholar] [CrossRef] [PubMed]
  144. Bergin, C.J.; Zouggar, A.; Haebe, J.R.; Masibag, A.N.; Desrochers, F.M.; Reilley, S.Y.; Agrawal, G.; Benoit, Y.D. G9a controls pluripotent-like identity and tumor-initiating function in human colorectal cancer. Oncogene 2021, 40, 1191–1202. [Google Scholar] [CrossRef]
  145. Segovia, C.; San Jose-Eneriz, E.; Munera-Maravilla, E.; Martinez-Fernandez, M.; Garate, L.; Miranda, E.; Vilas-Zornoza, A.; Lodewijk, I.; Rubio, C.; Segrelles, C.; et al. Inhibition of a G9a/DNMT network triggers immune-mediated bladder cancer regression. Nat. Med. 2019, 25, 1073–1081. [Google Scholar] [CrossRef]
  146. Pangeni, R.P.; Yang, L.; Zhang, K.; Wang, J.; Li, W.; Guo, C.; Yun, X.; Sun, T.; Wang, J.; Raz, D.J. G9a regulates tumorigenicity and stemness through genome-wide DNA methylation reprogramming in non-small cell lung cancer. Clin. Epigenet. 2020, 12, 88. [Google Scholar] [CrossRef]
  147. Casciello, F.; Al-Ejeh, F.; Kelly, G.; Brennan, D.J.; Ngiow, S.F.; Young, A.; Stoll, T.; Windloch, K.; Hill, M.M.; Smyth, M.J.; et al. G9a drives hypoxia-mediated gene repression for breast cancer cell survival and tumorigenesis. Proc. Natl. Acad. Sci. USA 2017, 114, 7077–7082. [Google Scholar] [CrossRef] [Green Version]
  148. Wang, G.; Long, J.; Gao, Y.; Zhang, W.; Han, F.; Xu, C.; Sun, L.; Yang, S.C.; Lan, J.; Hou, Z.; et al. SETDB1-mediated methylation of Akt promotes its K63-linked ubiquitination and activation leading to tumorigenesis. Nat. Cell Biol. 2019, 21, 214–225. [Google Scholar] [CrossRef]
  149. Shang, W.; Wang, Y.; Liang, X.; Li, T.; Shao, W.; Liu, F.; Cui, X.; Wang, Y.; Lv, L.; Chai, L.; et al. SETDB1 promotes gastric carcinogenesis and metastasis via upregulation of CCND1 and MMP9 expression. J. Pathol. 2021, 253, 148–159. [Google Scholar] [CrossRef]
  150. Hou, Z.; Sun, L.; Xu, F.; Hu, F.; Lan, J.; Song, D.; Feng, Y.; Wang, J.; Luo, X.; Hu, J.; et al. Blocking histone methyltransferase SETDB1 inhibits tumorigenesis and enhances cetuximab sensitivity in colorectal cancer. Cancer Lett. 2020, 487, 63–73. [Google Scholar] [CrossRef]
  151. Ye, J.; Huang, A.; Wang, H.; Zhang, A.M.Y.; Huang, X.; Lan, Q.; Sato, T.; Goyama, S.; Kurokawa, M.; Deng, C.; et al. PRDM3 attenuates pancreatitis and pancreatic tumorigenesis by regulating inflammatory response. Cell Death Dis. 2020, 11, 187. [Google Scholar] [CrossRef] [Green Version]
  152. Kundu, A.; Nam, H.; Shelar, S.; Chandrashekar, D.S.; Brinkley, G.; Karki, S.; Mitchell, T.; Livi, C.B.; Buckhaults, P.; Kirkman, R.; et al. PRDM16 suppresses HIF-targeted gene expression in kidney cancer. J. Exp. Med. 2020, 217, e20191005. [Google Scholar] [CrossRef] [Green Version]
  153. Yin, G.; Yan, C.; Hao, J.; Zhang, C.; Wang, P.; Zhao, C.; Cai, S.; Meng, B.; Zhang, A.; Li, L. PRDM16, negatively regulated by miR-372-3p, suppresses cell proliferation and invasion in prostate cancer. Andrologia 2022, e14529. [Google Scholar] [CrossRef]
  154. Fei, L.R.; Huang, W.J.; Wang, Y.; Lei, L.; Li, Z.H.; Zheng, Y.W.; Wang, Z.; Yang, M.Q.; Liu, C.C.; Xu, H.T. PRDM16 functions as a suppressor of lung adenocarcinoma metastasis. J. Exp. Clin. Cancer Res. 2019, 38, 35. [Google Scholar] [CrossRef] [Green Version]
  155. Zeng, Z.; Yang, Y.; Wu, H. MicroRNA-765 alleviates the malignant progression of breast cancer via interacting with EZH1. Am. J. Transl. Res. 2019, 11, 4500–4507. [Google Scholar]
  156. Kusakabe, Y.; Chiba, T.; Oshima, M.; Koide, S.; Rizq, O.; Aoyama, K.; Ao, J.; Kaneko, T.; Kanzaki, H.; Kanayama, K.; et al. EZH1/2 inhibition augments the anti-tumor effects of sorafenib in hepatocellular carcinoma. Sci. Rep. 2021, 11, 21396. [Google Scholar] [CrossRef]
  157. Wan, L.; Li, X.; Shen, H.; Bai, X. Quantitative analysis of EZH2 expression and its correlations with lung cancer patients’ clinical pathological characteristics. Clin. Transl. Oncol. 2013, 15, 132–138. [Google Scholar] [CrossRef]
  158. Li, Z.; Wang, D.; Lu, J.; Huang, B.; Wang, Y.; Dong, M.; Fan, D.; Li, H.; Gao, Y.; Hou, P.; et al. Methylation of EZH2 by PRMT1 regulates its stability and promotes breast cancer metastasis. Cell Death Differ. 2020, 27, 3226–3242. [Google Scholar] [CrossRef]
  159. Pan, Y.M.; Wang, C.G.; Zhu, M.; Xing, R.; Cui, J.T.; Li, W.M.; Yu, D.D.; Wang, S.B.; Zhu, W.; Ye, Y.J.; et al. STAT3 signaling drives EZH2 transcriptional activation and mediates poor prognosis in gastric cancer. Mol. Cancer 2016, 15, 79. [Google Scholar] [CrossRef] [Green Version]
  160. Cheraghi, S.; Asadzadeh, H.; Javadi, G. Dysregulated Expression of Long Non-Coding RNA MINCR and EZH2 in Colorectal Cancer. Iran. Biomed. J. 2022, 26, 64–69. [Google Scholar] [CrossRef]
  161. Yuan, H.; Han, Y.; Wang, X.; Li, N.; Liu, Q.; Yin, Y.; Wang, H.; Pan, L.; Li, L.; Song, K.; et al. SETD2 Restricts Prostate Cancer Metastasis by Integrating EZH2 and AMPK Signaling Pathways. Cancer Cell 2020, 38, 350–365.e7. [Google Scholar] [CrossRef] [PubMed]
  162. Niu, N.; Lu, P.; Yang, Y.; He, R.; Zhang, L.; Shi, J.; Wu, J.; Yang, M.; Zhang, Z.G.; Wang, L.W.; et al. Loss of Setd2 promotes Kras-induced acinar-to-ductal metaplasia and epithelia-mesenchymal transition during pancreatic carcinogenesis. Gut 2020, 69, 715–726. [Google Scholar] [CrossRef] [PubMed]
  163. Chen, B.Y.; Song, J.; Hu, C.L.; Chen, S.B.; Zhang, Q.; Xu, C.H.; Wu, J.C.; Hou, D.; Sun, M.; Zhang, Y.L.; et al. SETD2 deficiency accelerates MDS-associated leukemogenesis via S100a9 in NHD13 mice and predicts poor prognosis in MDS. Blood 2020, 135, 2271–2285. [Google Scholar] [CrossRef] [PubMed]
  164. Kim, I.K.; McCutcheon, J.N.; Rao, G.; Liu, S.V.; Pommier, Y.; Skrzypski, M.; Zhang, Y.W.; Giaccone, G. Acquired SETD2 mutation and impaired CREB1 activation confer cisplatin resistance in metastatic non-small cell lung cancer. Oncogene 2019, 38, 180–193. [Google Scholar] [CrossRef]
  165. Ettel, M.; Zhao, L.; Schechter, S.; Shi, J. Expression and prognostic value of NSD1 and SETD2 in pancreatic ductal adenocarcinoma and its precursor lesions. Pathology 2019, 51, 392–398. [Google Scholar] [CrossRef]
  166. Cancer Genome Atlas Network. Comprehensive genomic characterization of head and neck squamous cell carcinomas. Nature 2015, 517, 576–582. [Google Scholar] [CrossRef] [Green Version]
  167. Sengupta, D.; Zeng, L.; Li, Y.; Hausmann, S.; Ghosh, D.; Yuan, G.; Nguyen, T.N.; Lyu, R.; Caporicci, M.; Morales Benitez, A.; et al. NSD2 dimethylation at H3K36 promotes lung adenocarcinoma pathogenesis. Mol. Cell 2021, 81, 4481–4492.e9. [Google Scholar] [CrossRef]
  168. Zhao, L.H.; Li, Q.; Huang, Z.J.; Sun, M.X.; Lu, J.J.; Zhang, X.H.; Li, G.; Wu, F. Identification of histone methyltransferase NSD2 as an important oncogenic gene in colorectal cancer. Cell Death Dis. 2021, 12, 974. [Google Scholar] [CrossRef]
  169. Gao, B.; Liu, X.; Li, Z.; Zhao, L.; Pan, Y. Overexpression of EZH2/NSD2 Histone Methyltransferase Axis Predicts Poor Prognosis and Accelerates Tumor Progression in Triple-Negative Breast Cancer. Front. Oncol. 2020, 10, 600514. [Google Scholar] [CrossRef]
  170. Lu, M.H.; Fan, M.F.; Yu, X.D. NSD2 promotes osteosarcoma cell proliferation and metastasis by inhibiting E-cadherin expression. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 928–936. [Google Scholar]
  171. Han, X.; Piao, L.; Xu, X.; Luo, F.; Liu, Z.; He, X. NSD2 Promotes Renal Cancer Progression Through Stimulating Akt/Erk Signaling. Cancer Manag. Res. 2020, 12, 375–383. [Google Scholar] [CrossRef] [Green Version]
  172. Aytes, A.; Giacobbe, A.; Mitrofanova, A.; Ruggero, K.; Cyrta, J.; Arriaga, J.; Palomero, L.; Farran-Matas, S.; Rubin, M.A.; Shen, M.M.; et al. NSD2 is a conserved driver of metastatic prostate cancer progression. Nat. Commun. 2018, 9, 5201. [Google Scholar] [CrossRef] [Green Version]
  173. Jeong, G.Y.; Park, M.K.; Choi, H.J.; An, H.W.; Park, Y.U.; Choi, H.J.; Park, J.; Kim, H.Y.; Son, T.; Lee, H.; et al. NSD3-Induced Methylation of H3K36 Activates NOTCH Signaling to Drive Breast Tumor Initiation and Metastatic Progression. Cancer Res. 2021, 81, 77–90. [Google Scholar] [CrossRef]
  174. Yi, L.; Yi, L.; Liu, Q.; Li, C. Downregulation of NSD3 (WHSC1L1) inhibits cell proliferation and migration via ERK1/2 deactivation and decreasing CAPG expression in colorectal cancer cells. OncoTargets Ther. 2019, 12, 3933–3943. [Google Scholar] [CrossRef] [Green Version]
  175. Sun, Y.; Xie, J.; Cai, S.; Wang, Q.; Feng, Z.; Li, Y.; Lu, J.J.; Chen, W.; Ye, Z. Elevated expression of nuclear receptor-binding SET domain 3 promotes pancreatic cancer cell growth. Cell Death Dis. 2021, 12, 913. [Google Scholar] [CrossRef]
  176. Hassan, N.; Rutsch, N.; Gyorffy, B.; Espinoza-Sanchez, N.A.; Gotte, M. SETD3 acts as a prognostic marker in breast cancer patients and modulates the viability and invasion of breast cancer cells. Sci. Rep. 2020, 10, 2262. [Google Scholar] [CrossRef] [Green Version]
  177. Zou, T.; Wang, Y.; Dong, L.; Che, T.; Zhao, H.; Yan, X.; Lin, Z. Stabilization of SETD3 by deubiquitinase USP27 enhances cell proliferation and hepatocellular carcinoma progression. Cell. Mol. Life Sci. 2022, 79, 70. [Google Scholar] [CrossRef]
  178. Yu, M.; Jia, Y.; Ma, Z.; Ji, D.; Wang, C.; Liang, Y.; Zhang, Q.; Yi, H.; Zeng, L. Structural insight into ASH1L PHD finger recognizing methylated histone H3K4 and promoting cell growth in prostate cancer. Front. Oncol. 2022, 12, 906807. [Google Scholar] [CrossRef]
  179. Zhu, L.; Li, Q.; Wong, S.H.; Huang, M.; Klein, B.J.; Shen, J.; Ikenouye, L.; Onishi, M.; Schneidawind, D.; Buechele, C.; et al. ASH1L Links Histone H3 Lysine 36 Dimethylation to MLL Leukemia. Cancer Discov. 2016, 6, 770–783. [Google Scholar] [CrossRef] [Green Version]
  180. Jeyaratnam, D.C.; Baduin, B.S.; Hansen, M.C.; Hansen, M.; Jorgensen, J.M.; Aggerholm, A.; Ommen, H.B.; Hokland, P.; Nyvold, C.G. Delineation of known and new transcript variants of the SETMAR (Metnase) gene and the expression profile in hematologic neoplasms. Exp. Hematol. 2014, 42, 448–456.e4. [Google Scholar] [CrossRef]
  181. Houle, A.A.; Gibling, H.; Lamaze, F.C.; Edgington, H.A.; Soave, D.; Fave, M.J.; Agbessi, M.; Bruat, V.; Stein, L.D.; Awadalla, P. Aberrant PRDM9 expression impacts the pan-cancer genomic landscape. Genome Res. 2018, 28, 1611–1620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Vatapalli, R.; Sagar, V.; Rodriguez, Y.; Zhao, J.C.; Unno, K.; Pamarthy, S.; Lysy, B.; Anker, J.; Han, H.; Yoo, Y.A.; et al. Histone methyltransferase DOT1L coordinates AR and MYC stability in prostate cancer. Nat. Commun. 2020, 11, 4153. [Google Scholar] [CrossRef] [PubMed]
  183. Liu, C.; Yang, Q.; Zhu, Q.; Lu, X.; Li, M.; Hou, T.; Li, Z.; Tang, M.; Li, Y.; Wang, H.; et al. CBP mediated DOT1L acetylation confers DOT1L stability and promotes cancer metastasis. Theranostics 2020, 10, 1758–1776. [Google Scholar] [CrossRef] [PubMed]
  184. Song, Z.; Wei, Z.; Wang, Q.; Zhang, X.; Tao, X.; Wu, N.; Liu, X.; Qian, J. The role of DOT1L in the proliferation and prognosis of gastric cancer. Biosci. Rep. 2020, 40, BSR20193515. [Google Scholar] [CrossRef] [Green Version]
  185. Wang, X.; Wang, H.; Xu, B.; Jiang, D.; Huang, S.; Yu, H.; Wu, Z.; Wu, Q. Depletion of H3K79 methyltransferase Dot1L promotes cell invasion and cancer stem-like cell property in ovarian cancer. Am. J. Transl. Res. 2019, 11, 1145–1153. [Google Scholar]
  186. Kurani, H.; Razavipour, S.F.; Harikumar, K.B.; Dunworth, M.; Ewald, A.J.; Nasir, A.; Pearson, G.; Van Booven, D.; Zhou, Z.; Azzam, D.; et al. DOT1L Is a Novel Cancer Stem Cell Target for Triple-Negative Breast Cancer. Clin. Cancer Res. 2022, 28, 1948–1965. [Google Scholar] [CrossRef]
  187. Hou, L.; Li, Q.; Yu, Y.; Li, M.; Zhang, D. SET8 induces epithelial-mesenchymal transition and enhances prostate cancer cell metastasis by cooperating with ZEB1. Mol. Med. Rep. 2016, 13, 1681–1688. [Google Scholar] [CrossRef] [Green Version]
  188. Wu, J.; Qiao, K.; Du, Y.; Zhang, X.; Cheng, H.; Peng, L.; Guo, Z. Downregulation of histone methyltransferase SET8 inhibits progression of hepatocellular carcinoma. Sci. Rep. 2020, 10, 4490. [Google Scholar] [CrossRef] [Green Version]
  189. Liu, B.; Zhang, X.; Song, F.; Liu, Q.; Dai, H.; Zheng, H.; Cui, P.; Zhang, L.; Zhang, W.; Chen, K. A functional single nucleotide polymorphism of SET8 is prognostic for breast cancer. Oncotarget 2016, 7, 34277–34287. [Google Scholar] [CrossRef] [Green Version]
  190. Pogribny, I.P.; Ross, S.A.; Tryndyak, V.P.; Pogribna, M.; Poirier, L.A.; Karpinets, T.V. Histone H3 lysine 9 and H4 lysine 20 trimethylation and the expression of Suv4-20h2 and Suv-39h1 histone methyltransferases in hepatocarcinogenesis induced by methyl deficiency in rats. Carcinogenesis 2006, 27, 1180–1186. [Google Scholar] [CrossRef]
  191. Tryndyak, V.P.; Kovalchuk, O.; Pogribny, I.P. Loss of DNA methylation and histone H4 lysine 20 trimethylation in human breast cancer cells is associated with aberrant expression of DNA methyltransferase 1, Suv4-20h2 histone methyltransferase and methyl-binding proteins. Cancer Biol. Ther. 2006, 5, 65–70. [Google Scholar] [CrossRef] [Green Version]
  192. Wu, Z.; Guan, K.L. Acetyl-CoA, protein acetylation, and liver cancer. Mol. Cell 2022, 82, 4196–4198. [Google Scholar] [CrossRef]
  193. Verdone, L.; Caserta, M.; Di Mauro, E. Role of histone acetylation in the control of gene expression. Biochem. Cell Biol. 2005, 83, 344–353. [Google Scholar] [CrossRef]
  194. Klein, B.J.; Jang, S.M.; Lachance, C.; Mi, W.; Lyu, J.; Sakuraba, S.; Krajewski, K.; Wang, W.W.; Sidoli, S.; Liu, J.; et al. Histone H3K23-specific acetylation by MORF is coupled to H3K14 acylation. Nat. Commun. 2019, 10, 4724. [Google Scholar] [CrossRef] [Green Version]
  195. Dang, F.; Wei, W. Targeting the acetylation signaling pathway in cancer therapy. Semin. Cancer Biol. 2022, 85, 209–218. [Google Scholar] [CrossRef]
  196. Zhang, T.; Chen, Y.; Li, J.; Yang, F.; Wu, H.; Dai, F.; Hu, M.; Lu, X.; Peng, Y.; Liu, M.; et al. Antitumor action of a novel histone deacetylase inhibitor, YF479, in breast cancer. Neoplasia 2014, 16, 665–677. [Google Scholar] [CrossRef] [Green Version]
  197. Guo, P.; Chen, W.; Li, H.; Li, M.; Li, L. The Histone Acetylation Modifications of Breast Cancer and their Therapeutic Implications. Pathol. Oncol. Res. 2018, 24, 807–813. [Google Scholar] [CrossRef]
  198. Yang, G.; Yuan, Y.; Yuan, H.; Wang, J.; Yun, H.; Geng, Y.; Zhao, M.; Li, L.; Weng, Y.; Liu, Z.; et al. Histone acetyltransferase 1 is a succinyltransferase for histones and non-histones and promotes tumorigenesis. EMBO Rep. 2021, 22, e50967. [Google Scholar] [CrossRef]
  199. Yin, Y.W.; Jin, H.J.; Zhao, W.; Gao, B.; Fang, J.; Wei, J.; Zhang, D.D.; Zhang, J.; Fang, D. The Histone Acetyltransferase GCN5 Expression Is Elevated and Regulated by c-Myc and E2F1 Transcription Factors in Human Colon Cancer. Gene Expr. 2015, 16, 187–196. [Google Scholar] [CrossRef] [Green Version]
  200. Stemmler, M.P. PCAF, ISX, and BRD4: A maleficent alliance serving lung cancer malignancy. EMBO Rep. 2020, 21, e49766. [Google Scholar] [CrossRef]
  201. Cheng, Y.W.; Zeng, F.M.; Li, D.J.; Wang, S.H.; He, J.Z.; Guo, Z.C.; Nie, P.J.; Wu, Z.Y.; Shi, W.Q.; Wen, B.; et al. P300/CBP-associated factor (PCAF)-mediated acetylation of Fascin at lysine 471 inhibits its actin-bundling activity and tumor metastasis in esophageal cancer. Cancer Commun. 2021, 41, 1398–1416. [Google Scholar] [CrossRef] [PubMed]
  202. Tan, K.N.; Avery, V.M.; Carrasco-Pozo, C. Metabolic Roles of Androgen Receptor and Tip60 in Androgen-Dependent Prostate Cancer. Int. J. Mol. Sci. 2020, 21, 6622. [Google Scholar] [CrossRef] [PubMed]
  203. McGuire, A.; Casey, M.C.; Shalaby, A.; Kalinina, O.; Curran, C.; Webber, M.; Callagy, G.; Holian, E.; Bourke, E.; Kerin, M.J.; et al. Quantifying Tip60 (Kat5) stratifies breast cancer. Sci. Rep. 2019, 9, 3819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zhu, H.; Wang, Y.; Wei, T.; Zhao, X.; Li, F.; Li, Y.; Wang, F.; Cai, Y.; Jin, J. KAT8/MOF-Mediated Anti-Cancer Mechanism of Gemcitabine in Human Bladder Cancer Cells. Biomol. Ther. 2021, 29, 184–194. [Google Scholar] [CrossRef]
  205. Wu, Y.; Zeng, K.; Wang, C.; Wang, S.; Sun, H.; Liu, W.; Wang, X.; Niu, J.; Cong, S.Y.; Zhou, X.; et al. Histone acetyltransferase MOF is involved in suppression of endometrial cancer and maintenance of ERalpha stability. Biochem. Biophys. Res. Commun. 2019, 509, 541–548. [Google Scholar] [CrossRef]
  206. Guo, R.; Liang, Y.; Zou, B.; Li, D.; Wu, Z.; Xie, F.; Zhang, X.; Li, X. The Histone Acetyltransferase MOF Regulates SIRT1 Expression to Suppress Renal Cell Carcinoma Progression. Front. Oncol. 2022, 12, 842967. [Google Scholar] [CrossRef]
  207. Hemming, M.L.; Benson, M.R.; Loycano, M.A.; Anderson, J.A.; Andersen, J.L.; Taddei, M.L.; Krivtsov, A.V.; Aubrey, B.J.; Cutler, J.A.; Hatton, C.; et al. MOZ and Menin-MLL Complexes Are Complementary Regulators of Chromatin Association and Transcriptional Output in Gastrointestinal Stromal Tumor. Cancer Discov. 2022, 12, 1804–1823. [Google Scholar] [CrossRef]
  208. Yokoyama, A. Role of the MOZ/MLL-mediated transcriptional activation system for self-renewal in normal hematopoiesis and leukemogenesis. FEBS J. 2021, 289, 7987–8002. [Google Scholar] [CrossRef]
  209. Baell, J.B.; Leaver, D.J.; Hermans, S.J.; Kelly, G.L.; Brennan, M.S.; Downer, N.L.; Nguyen, N.; Wichmann, J.; McRae, H.M.; Yang, Y.; et al. Inhibitors of histone acetyltransferases KAT6A/B induce senescence and arrest tumour growth. Nature 2018, 560, 253–257. [Google Scholar] [CrossRef]
  210. Chen, T.F.; Hao, H.F.; Zhang, Y.; Chen, X.Y.; Zhao, H.S.; Yang, R.; Li, P.; Qiu, L.X.; Sang, Y.H.; Xu, C.; et al. HBO1 induces histone acetylation and is important for non-small cell lung cancer cell growth. Int. J. Biol. Sci. 2022, 18, 3313–3323. [Google Scholar] [CrossRef]
  211. Gao, Y.Y.; Ling, Z.Y.; Zhu, Y.R.; Shi, C.; Wang, Y.; Zhang, X.Y.; Zhang, Z.Q.; Jiang, Q.; Chen, M.B.; Yang, S.; et al. The histone acetyltransferase HBO1 functions as a novel oncogenic gene in osteosarcoma. Theranostics 2021, 11, 4599–4615. [Google Scholar] [CrossRef]
  212. Zhong, W.; Liu, H.; Deng, L.; Chen, G.; Liu, Y. HBO1 overexpression is important for hepatocellular carcinoma cell growth. Cell Death Dis. 2021, 12, 549. [Google Scholar] [CrossRef]
  213. Iizuka, M.; Susa, T.; Takahashi, Y.; Tamamori-Adachi, M.; Kajitani, T.; Okinaga, H.; Fukusato, T.; Okazaki, T. Histone acetyltransferase Hbo1 destabilizes estrogen receptor α by ubiquitination and modulates proliferation of breast cancers. Cancer Sci. 2013, 104, 1647–1655. [Google Scholar] [CrossRef]
  214. Gruber, M.; Ferrone, L.; Puhr, M.; Santer, F.R.; Furlan, T.; Eder, I.E.; Sampson, N.; Schäfer, G.; Handle, F.; Culig, Z. p300 is upregulated by docetaxel and is a target in chemoresistant prostate cancer. Endocr. Relat. Cancer 2020, 27, 187–198. [Google Scholar] [CrossRef] [Green Version]
  215. Hou, X.; Gong, R.; Zhan, J.; Zhou, T.; Ma, Y.; Zhao, Y.; Zhang, Y.; Chen, G.; Zhang, Z.; Ma, S.; et al. p300 promotes proliferation, migration, and invasion via inducing epithelial-mesenchymal transition in non-small cell lung cancer cells. BMC Cancer 2018, 18, 641. [Google Scholar] [CrossRef]
  216. Clague, M.J.; Urbe, S. Integration of cellular ubiquitin and membrane traffic systems: Focus on deubiquitylases. FEBS J. 2017, 284, 1753–1766. [Google Scholar] [CrossRef] [Green Version]
  217. Heideker, J.; Wertz, I.E. DUBs, the regulation of cell identity and disease. Biochem. J. 2015, 467, 191. [Google Scholar] [CrossRef]
  218. Li, Y.; Yuan, J. Role of deubiquitinating enzymes in DNA double-strand break repair. J. Zhejiang Univ. Sci. B 2021, 22, 63–72. [Google Scholar] [CrossRef]
  219. Mattiroli, F.; Penengo, L. Histone Ubiquitination: An Integrative Signaling Platform in Genome Stability. Trends Genet. 2021, 37, 566–581. [Google Scholar] [CrossRef]
  220. Zhang, X.; Li, B.; Rezaeian, A.H.; Xu, X.; Chou, P.C.; Jin, G.; Han, F.; Pan, B.S.; Wang, C.Y.; Long, J.; et al. H3 ubiquitination by NEDD4 regulates H3 acetylation and tumorigenesis. Nat. Commun. 2017, 8, 14799. [Google Scholar] [CrossRef] [Green Version]
  221. Yadav, P.; Subbarayalu, P.; Medina, D.; Nirzhor, S.; Timilsina, S.; Rajamanickam, S.; Eedunuri, V.K.; Gupta, Y.; Zheng, S.; Abdelfattah, N.; et al. M6A RNA Methylation Regulates Histone Ubiquitination to Support Cancer Growth and Progression. Cancer Res. 2022, 82, 1872–1889. [Google Scholar] [CrossRef] [PubMed]
  222. Challa, K.; Schmid, C.D.; Kitagawa, S.; Cheblal, A.; Iesmantavicius, V.; Seeber, A.; Amitai, A.; Seebacher, J.; Hauer, M.H.; Shimada, K.; et al. Damage-induced chromatome dynamics link Ubiquitin ligase and proteasome recruitment to histone loss and efficient DNA repair. Mol. Cell 2021, 81, 811–829.e6. [Google Scholar] [CrossRef] [PubMed]
  223. Ting, X.; Xia, L.; Yang, J.; He, L.; Si, W.; Shang, Y.; Sun, L. USP11 acts as a histone deubiquitinase functioning in chromatin reorganization during DNA repair. Nucleic Acids Res. 2019, 47, 9721–9740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Cerutti, H.; Casas-Mollano, J.A. Histone H3 phosphorylation: Universal code or lineage specific dialects? Epigenetics 2009, 4, 71–75. [Google Scholar] [CrossRef] [Green Version]
  225. Zhang, S.; Yu, X.; Zhang, Y.; Xue, X.; Yu, Q.; Zha, Z.; Gogol, M.; Workman, J.L.; Li, S. Metabolic regulation of telomere silencing by SESAME complex-catalyzed H3T11 phosphorylation. Nat. Commun. 2021, 12, 594. [Google Scholar] [CrossRef]
  226. Metzger, E.; Yin, N.; Wissmann, M.; Kunowska, N.; Fischer, K.; Friedrichs, N.; Patnaik, D.; Higgins, J.M.; Potier, N.; Scheidtmann, K.H.; et al. Phosphorylation of histone H3 at threonine 11 establishes a novel chromatin mark for transcriptional regulation. Nat. Cell Biol. 2008, 10, 53–60. [Google Scholar] [CrossRef]
  227. Armache, A.; Yang, S.; Martinez de Paz, A.; Robbins, L.E.; Durmaz, C.; Cheong, J.Q.; Ravishankar, A.; Daman, A.W.; Ahimovic, D.J.; Klevorn, T.; et al. Histone H3.3 phosphorylation amplifies stimulation-induced transcription. Nature 2020, 583, 852–857. [Google Scholar] [CrossRef]
  228. Udugama, M.; Vinod, B.; Chan, F.L.; Hii, L.; Garvie, A.; Collas, P.; Kalitsis, P.; Steer, D.; Das, P.P.; Tripathi, P.; et al. Histone H3.3 phosphorylation promotes heterochromatin formation by inhibiting H3K9/K36 histone demethylase. Nucleic Acids Res. 2022, 50, 4500–4514. [Google Scholar] [CrossRef]
  229. He, F.; Yu, Q.; Wang, M.; Wang, R.; Gong, X.; Ge, F.; Yu, X.; Li, S. SESAME-catalyzed H3T11 phosphorylation inhibits Dot1-catalyzed H3K79me3 to regulate autophagy and telomere silencing. Nat. Commun. 2022, 13, 7526. [Google Scholar] [CrossRef]
  230. Leal, J.A.; Estrada-Tobar, Z.M.; Wade, F.; Mendiola, A.J.P.; Meza, A.; Mendoza, M.; Nerenberg, P.S.; Zurita-Lopez, C.I. Phosphoserine inhibits neighboring arginine methylation in the RKS motif of histone H3. Arch. Biochem. Biophys. 2021, 698, 108716. [Google Scholar] [CrossRef]
  231. Kim, J.J.; Lee, S.Y.; Miller, K.M. Preserving genome integrity and function: The DNA damage response and histone modifications. Crit. Rev. Biochem. Mol. Biol. 2019, 54, 208–241. [Google Scholar] [CrossRef]
  232. Zhao, Z.; Shilatifard, A. Epigenetic modifications of histones in cancer. Genome Biol. 2019, 20, 245. [Google Scholar] [CrossRef] [Green Version]
  233. Fults, D.W. Stemming the growth of pediatric gliomas through histone modification. Neuro Oncol. 2021, 23, 341–342. [Google Scholar] [CrossRef]
  234. Allis, C.D. Pursuing the Secrets of Histone Proteins: An Amazing Journey with a Remarkable Supporting Cast. Cell 2018, 175, 18–21. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of histone modification landscape. The nucleosome consists of one H3-H4 tetramer, two H2A-H2B dimers, and the surrounding DNA. There are many modifications on the histone tail, including histone methylation (Me), acetylation (Ac), ubiquitination (Ub), and phosphorylation (P).
Figure 1. Schematic diagram of histone modification landscape. The nucleosome consists of one H3-H4 tetramer, two H2A-H2B dimers, and the surrounding DNA. There are many modifications on the histone tail, including histone methylation (Me), acetylation (Ac), ubiquitination (Ub), and phosphorylation (P).
Ijms 24 04939 g001
Figure 2. Diagram of RCNA and RINA. (A) Replication-coupled nucleosome assembly (RCNA). Nucleosomes must be disassembled to make way for DNA replication machinery. Nucleosomes are then reassembled in close concert with DNA replication on the leading and lagging strands. (B) Replication-independent nucleosome assembly (RINA). Many of the same core principles of nucleosome disassembly, DNA access and nucleosome assembly are likely applicable to replication-independent processes such as gene transcription.
Figure 2. Diagram of RCNA and RINA. (A) Replication-coupled nucleosome assembly (RCNA). Nucleosomes must be disassembled to make way for DNA replication machinery. Nucleosomes are then reassembled in close concert with DNA replication on the leading and lagging strands. (B) Replication-independent nucleosome assembly (RINA). Many of the same core principles of nucleosome disassembly, DNA access and nucleosome assembly are likely applicable to replication-independent processes such as gene transcription.
Ijms 24 04939 g002
Figure 3. The function of histone modification in nucleosome assembly in yeast. The new H3--H4 is acetylated by the acetyltransferase Hat1/Hat2, and Asf1 promotes nuclear import of H3-H4K5/12ac. Subsequently, acetylation of H3K56 by Rtt109-Vps75 promotes ubiquitination of H3K121/122/125 by Rtt101Mms1/Mms22. The modified H3-H4 dissociates from Asf1 and is presented to the chaperones Caf1 and Rtt106 to deposit histones on DNA for nucleosome assembly.
Figure 3. The function of histone modification in nucleosome assembly in yeast. The new H3--H4 is acetylated by the acetyltransferase Hat1/Hat2, and Asf1 promotes nuclear import of H3-H4K5/12ac. Subsequently, acetylation of H3K56 by Rtt109-Vps75 promotes ubiquitination of H3K121/122/125 by Rtt101Mms1/Mms22. The modified H3-H4 dissociates from Asf1 and is presented to the chaperones Caf1 and Rtt106 to deposit histones on DNA for nucleosome assembly.
Ijms 24 04939 g003
Figure 4. Histone modifications function in nucleosome assembly in mammals. After H4 acetylation on K5 and K12 by HAT1, Asf1 and Importin-4 help nuclear import of H3-H4K5/12ac. Subsequently, acetyltransferase p300/CBP catalyzes the acetylation of H3K56 to promote the ubiquitination of H3K122 by Cul4ADDB1. The modified H3-H4 is presented to the histone chaperone CAF-1 to assemble the nucleosome.
Figure 4. Histone modifications function in nucleosome assembly in mammals. After H4 acetylation on K5 and K12 by HAT1, Asf1 and Importin-4 help nuclear import of H3-H4K5/12ac. Subsequently, acetyltransferase p300/CBP catalyzes the acetylation of H3K56 to promote the ubiquitination of H3K122 by Cul4ADDB1. The modified H3-H4 is presented to the histone chaperone CAF-1 to assemble the nucleosome.
Ijms 24 04939 g004
Figure 5. DNA damage repair by chromatin assembly in human. Encounter of a replication fork with DNA lesions that hamper its advance triggers the DNA damage repair (DDR) response. In human cells, MMS22L-TONSL mediates the deposition of newly synthesized histones, binds to unmethylated H3-H4K20 (H3-H4K20me0), and promotes RAD51 recruitment and replication fork recombination repair.
Figure 5. DNA damage repair by chromatin assembly in human. Encounter of a replication fork with DNA lesions that hamper its advance triggers the DNA damage repair (DDR) response. In human cells, MMS22L-TONSL mediates the deposition of newly synthesized histones, binds to unmethylated H3-H4K20 (H3-H4K20me0), and promotes RAD51 recruitment and replication fork recombination repair.
Ijms 24 04939 g005
Figure 6. DNA damage tolerance by chromatin assembly in yeast. In yeast, replication-coupled, newly synthesized histones are deposited at the fork to mark chromatin with acetylated H3K56, which in turn activates ubiquitin ligase activity in the Rtt101Mms1/Mms22 complex. Ubiquitination of Mrc1 and Ctf4 by unknown factors causes uncoupling of helicase CMG with polymerase and promotes recombination repair of bypass. In addition, the chromatin assembly factor CAF-1 interacts with the D-loop of the RecQ helicase by assembling the nucleosome onto the D-loop and abolishing the D-loop dissociation activity.
Figure 6. DNA damage tolerance by chromatin assembly in yeast. In yeast, replication-coupled, newly synthesized histones are deposited at the fork to mark chromatin with acetylated H3K56, which in turn activates ubiquitin ligase activity in the Rtt101Mms1/Mms22 complex. Ubiquitination of Mrc1 and Ctf4 by unknown factors causes uncoupling of helicase CMG with polymerase and promotes recombination repair of bypass. In addition, the chromatin assembly factor CAF-1 interacts with the D-loop of the RecQ helicase by assembling the nucleosome onto the D-loop and abolishing the D-loop dissociation activity.
Ijms 24 04939 g006
Table 2. The role of genes encoding histone acetyltransferase in human cancer.
Table 2. The role of genes encoding histone acetyltransferase in human cancer.
GeneTumorRoleReference
HAT1Liver cancer, Pancreatic cancer, CholangiocarcinomaOverexpression[198]
GCN5Colorectal cancerOverexpression[199]
PCAFLung cancer, Esophageal cancerAmbiguous [200,201]
Tip60Prostate cancer, Breast cancerAmbiguous[202,203]
MOFBladder cancer, Endometrial cancer, Renal cell carcinomaSuppressor[204,205,206]
MOZGastrointestinal stromal tumor, Acute myeloid leukemiaOncogene[207,208]
MORFlymphomaOncogene [209]
HBO1Lung cancer, Osteosarcoma, Hepatocellular carcinoma, Breast cancerOncogenic [210,211,212,213]
p300Esophageal cancer, Prostate cancer, Lung cancerOncogenic[201,214,215]
CBPColorectal cancer Oncogenic[183]
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

Zhang, Y.; Zhang, Q.; Zhang, Y.; Han, J. The Role of Histone Modification in DNA Replication-Coupled Nucleosome Assembly and Cancer. Int. J. Mol. Sci. 2023, 24, 4939. https://doi.org/10.3390/ijms24054939

AMA Style

Zhang Y, Zhang Q, Zhang Y, Han J. The Role of Histone Modification in DNA Replication-Coupled Nucleosome Assembly and Cancer. International Journal of Molecular Sciences. 2023; 24(5):4939. https://doi.org/10.3390/ijms24054939

Chicago/Turabian Style

Zhang, Yaguang, Qin Zhang, Yang Zhang, and Junhong Han. 2023. "The Role of Histone Modification in DNA Replication-Coupled Nucleosome Assembly and Cancer" International Journal of Molecular Sciences 24, no. 5: 4939. https://doi.org/10.3390/ijms24054939

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