Skip to main content
American Journal of Translational Research logoLink to American Journal of Translational Research
. 2015 Nov 15;7(11):2159–2175.

Histone methyltransferases: novel targets for tumor and developmental defects

Xin Yi 1,2, Xue-Jun Jiang 1,2, Xiao-Yan Li 1,2, Ding-Sheng Jiang 3,4,5
PMCID: PMC4697697  PMID: 26807165

Abstract

Histone lysine methylation plays a critical role in epigenetic regulation of eukaryotes. To date, studies have shown that lysine residues of K4, K9, K27, K36 and K79 in histone H3 and K20 in histone H4 can be modified by histone methyltransferases (HMTs). Such histone methylation can specifically activate or repress the transcriptional activity to play a key role in gene expression/regulation and biological genetics. Importantly, abnormities of patterns or levels of histone methylation in higher eukaryotes may result in tumorigenesis and developmental defects, suggesting histone methylation will be one of the important targets or markers for treating these diseases. This review will outline the structural characteristics, active sites and specificity of HMTs, correlation between histone methylation and human diseases and lay special emphasis on the progress of the research on H3K36 methylation.

Keywords: Histone lysine methylation, histone methyltransferases, epigenetic modification

Introduction

Nucleosome, a complex formed by repeated winding and folding of both DNAs and histones, is a basic structural unit of chromosome. Formation of nucleosome firstly needs a histone octamer (as surrounded by a segment of approximately 200 bp DNA) consisting of 2 copies each of the core histones H2A, H2B, H3, and H4 [1,2]. Then, approximately ~147 bp of the aforementioned 200 bp DNA directly wrapped around the histone octamer comprises a core particle, called “core DNA”, which is difficult to be digested and decomposed by nucleases; whereas approximately 20-80 bp DNA acts on the “linker DNA” connecting two neighboring nucleosomes where H1 (a linker histone) binds to [3-5]. The terminal tails of the nucleosomal core histone, N-terminal tails, are subject to various external modifications because they are freed externally. Modifications known to date include methylation, acetylation, phosphorylation, ubiquitination and ADP-ribosylation, which can modulate the affinity between DNAs and histones to alter the chromatin structure conditions (including causing looser or tighter chromatin) or function as a regulator of gene expression similar to genetic codes in DNA (now termed “histone code”) by regulating the binding characteristics of transcription factors to DNA sequences. However, such histone modifications, together with DNA methylation and RNA modification, constitute the epigenetic modification [6-8]. In recent years, histone methylation has been a research hotspot in epigenetics, and also a focus of molecular biology, genetics and oncology [9-12]. However, histone methylation mainly occurs on arginine and lysine residues of histones H3 and H4 that are mainly regulated by histone methyltransferases (HMTs). Of the 24 sites of histone methylation ever discovered, there are 17 lysine and 7 arginine residues. Lysine residues can be modified by mono-, di- and tri-methylation, whereas arginine residues by mono- and di-methylation [13,14]. This article is trying to outline histone lysine methyltransferases with respect to structures, active sites and specificity, and lay special emphasis on the relevant research progress of histone H3 lysine 36 (H3K36) methylation so far.

HMTs

Structural characteristics of HMTs

Histone methylation is mainly regulated by a series of HMTs containing highly conserved core SET, cysteine-rich pre- and post-SET domains. SET domains were named after the initials of the three genes first discovered which express such domains, namely, Suppressor of variegation 3-9 (Su(var) 3-9), Enhancer of zeste (E(z)) and Trithorax (Trx) [15-18]. The catalytic domain in the SET domain is in charge of determining the catalytic activity of HMT; the pre-SET domain functions as a maintainer of the structural stability of the protein; whereas the post-SET domain offers a hydrophobic channel to participate in composition of parts of active sites of the enzyme [19-22]. Because the presence of SET domain makes HMTs different from other methyltransferases, it has its own unique folding structure. The SET domain is a peptide chain containing 130 amino acid residues, with high conservation. In this domain, N- and C-termini separately coil up and circle round to constitute two non-adjacent spatial conformations with 3-4 short folds; then a short helix containing 9 rings links with the spatial conformations of N- and C-termini. The vast majority of C-termini of the SET domain coil up into a “pseudoknot-like” structure. This topological structure passes through a short helix containing 9 rings and then links to other sides in the side chain of the C-terminus of the SET domain, forming the core SET domain which contains the most conservative motifs (ELXF/YDY and NHS/CXXPN) in the SET domain [23-25]. In brief, the “pseudoknot-like” structure at the C-terminus of the SET domain is constituted by the fragments of the C-terminus passing through the terminal sequence of the protein and extending forward to form ring structures. Moreover, there are also inserts in the SET domain, termed iSETs, which differ in length and specifically recognize the active substrates and cofactors of HMTs [26,27]. Furthermore, unlike the SET domain, pre- and post-SET domains on both sides of it are not highly conserved [28,29].

Types of HMTs

SET domain, an important domain of the HMTs, is a 130 amino acid-long peptide chain, which was first defined in 1998 [30]. To date, the SET domain has been found in the great majority of studies of organs of eukaryotes. There are 157 human-derived SET domain containing proteins in the SMART database and 93 in the Pfam database [31-35]. It is well known so far that SET domain containing proteins are majorly divided into seven families (SUV39, SET1, SET2, EZ, RIZ, SMYD and SUV4-20) and other rare members (i.e., SET7/9 and SET8) [36-39]. SET domain containing proteins in the same family share not only highly similar SET domains but also similar motifs beside this domain.

Target sites of HMTs

Histone methylation commonly occurs on lysine and arginine residues of H3 and H4 (Table 1). Of the 24 sites of histone methylation ever discovered, there are 17 lysine and 7 arginine residues, which are regulated by histone lysine methyltransferases (KMTs) and arginine methyltransferases (RMTs). Lysine residues can be modified by mono-, di- and tri-methylation [13,14]. Post-methylation biological effects vary from sites of histone lysine residues, by which the gene transcription can be either activated or repressed. For example, methylation modifications at histone H3 lysine 4 (H3K4) and H3K36 can activate the gene transcription, whereas those at histone H3 lysine 9 (H3K9), histone H3 lysine 27 (H3K27), histone H3 lysine 79 (H3K79) and histone H4 lysine 20 (H4K20) can repress it [39-43].

Table 1.

Histone methyltransferases, target sites, binding domains and biological functions

Target sites Histone methyltransferases Binding domain Biological functions
H3K4 SET1 Chd1 Transcriptional activation
MLL WDR5 Transcriptional elogation
SET7/9 JMJD2A Transcriptional scilencing
SMYD2 MBT
SMYD3 PHD finger
EZH2
ASH1
H3K9 SUV39h1 HP1 Transcriptional activation
SUV39h2 CDY1 Transcriptional repression
G9a JMJD2A DNA methylation
EZH2 Heterochromatic silencing
GLP Euchromatic silencing
RIZ
SETDB1
ASH1
H3K27 EZH1 Pc Transcriptional scilencing
EZH2 Euchromatic silencing
NSD2 X-inactivation
G9a
H3K36 NSD1 Eaf3 Transcriptional scilencing
NSD2 Transcriptional elogation
NSD3 Transcriptional regulation
ASH1
SetD2
H3K79 DOT1L 53BP1 DNA repair
Demarcation of euchromatin
H4K20 SUV4-20H1 Crb2 Transcriptional scilencing
SUV4-20H2 JMJD2A Transcriptional activation
NSD1 Transcriptional regulation
ASH1 Heterochromatic silencing
SET9 Cell cycle-dependent silencing
SET8/PR-SET7 Mitosis and cytokinesis

Diversity and specificity of HMTs

Histone methylation is a very complicated process. Because its sites are diverse, it can either specifically activate or repress the transcriptional activity of the gene. Methylation modification comes in many forms, including mono-, di- and tri-methylation. Thus it can be seen that diverse sites and patterns of methylation lead to tens of thousands of patterns of methylation modifications, increasing the complexity and diversity of gene expression regulated by histone methylation and also highlighting the importance of histone methylation in epigenetic regulation. Because HMT is also specific to the active site of the enzyme, it can possess its own specific substrate and target site. As for its specificity of the active site, the current mainstream view is that iSET, the insert in the SET domain, specifically recognizes the substrate and cofactor of the HMTs and function as an antigenic determinant, causing different HMTs to specifically recognize diverse substrates [26,27]. In addition, a few studies have reported HMTs may also recognize the same motif in different substrates, i.e., SET7/9 which can methylate not only H3K4 but p53 and TAF10 because these three substrates share the same motif, K/R-S/T-K [44]. Thus, the unique three-dimensional structure of the HMTs is characterized by specific recognition of the substrate and also exhibits the diversity of histone methylation.

Biological functions of histone methylation

Histone methylation is a part of epigenetic modification, but it possesses an extremely wide range of biological functions and plays important roles in transcriptional regulation, gene expression, heterochromatinization, genomic imprinting and X-inactivation [45-50]. Methylation at H3K4 and H3K36 can activate the gene transcription and that at H3K9, H3K27 and H4K20 can repress it; however, that at H3K79 is not involved in gene transcription and regulation directly but DNA damage/repair, where error localization occurs in P53BP1 and Crb around the DNA damage area and thus affects the process of DNA damage/repair when H3K79 methylation is repressed [51]. To date, it has been found that two members of the SUV39 family HMTs, SUV39h1 and SUV39h2, play a key role in heterochromatinization. When they mutate simultaneously, the methylation level of H3K9 will decrease by 50% or so and lead to disorders and deficits of separation of chromatin in neonatal mice during mitosis [52]. Moreover, SUV39h1 is a transcriptional repressor in mammals, which is co-localized in the transcriptionally silenced heterochromatin with the transcriptional repressor. When H3K9 is methylated by SUV39h1, HP1 can bind to the histone H3. It follows that SUV39h1 can recruit HP1 to the site of the heterochromatin during methylation of H3K9, playing a vital role in heterochromatinization [53]. Genomic imprinting is a phenomenon independent of the Mendelian inheritance, by which certain homologous alleles exhibit monoallelic expression during expression in a parent-of-origin-specific manner. Such genetic model is controlled monophyly-dependently. However, HMTs play a vital role in such genomic imprinting. It will lead to a loss of genomic imprinting of certain homologous alleles in genetic processing where there is a lack of functions of HMTs [54]. To date, scholars have found in studying the rat chromosome 7 that in order to maintain the genomic imprinting there also involves an important pathway-histone methylation-besides the important regulatory effect on DNA methylation [55]. In the dosage compensation mechanism for sex determination in organism, sex determination mainly depends on Xist, a non-coding RNA, and X-inactivation. Trimethylation of H3K27 (H3K27me3) and monomethylation of H4K20 (H4K20me1) are closely related to the Xist expression; DNA methylationis closely associated with X-inactivation, whereas maintenance of X-inactivation status is strongly related to the Xist expression. Thus it follows that X-inactivation is co-regulated by methylation of both histones and DNAs [56]. To sum up, histone methylation is strongly related to genetics and molecular biology.

Correlation between histone methylation and diseases

The structural feature, namely, HMTs containing SET domains, has determined a very close relationship between HMTs and human tumorigenesis (Table 2). It is found that members of the RIZ family mutate in hepatocellular carcinoma, breast carcinoma, spinal cord tumor, neuroblastoma, lung carcinoma, colon carcinoma and bone tumor, resulting in loss of activity of histone H3K9 methyltransferase. Loss of enzyme activity of RIZ family results in prolonged cell cycle G2/M phase and repression of apoptosis. It can be speculated that methylation at H3K9 by the RIZ family can repress the tumorigenesis [57-59]. Also, some studies have found that two members of the SUV39 family, SUV39h1 and SUV39h2, are closely related to geneses of B-cell lymphomas and non-Hodgkin’s lymphomas (NHLs), and that unstable gene expression of SUV39h1- and SUV39h2-knockout mice results in improper chromosome segregation and final genesis of B-cell lymphomas in mice. However, decrease in SUV39h1 and SUV39h2 methylation affects the interaction between SUV39h1/SUV39h2 and proteins of glioma retinae, thereby failing to regulate the gene expression of cyclinE in a normal manner and resulting in uncontrolled cell proliferation and thus promotion of carcinogenesis [60]. Moreover, mutation of mixed-lineage leukemia (MLL), the SET1 family member, results in the pathogenesis of leukemia; expression of the histone methyltransferase EZH2 increases aberrantly in patients with breast cancer, lymphoma and prostate cancer; histone methyltransferase SMYD3 also increases aberrantly in hepatocellular carcinoma and colorectal carcinoma cells [61]. Besides the fact that HMTs have a close relationship with human tumorigenesis, its role in the spectrum of human diseasehas attracted more and more attention. Recent studies also found that compared with normal healthy controls, histone methyltransferase Set7 was overexpressed in the peripheral blood mononuclear cells (PBMCs) of patients with type 2 diabetes and was closely linked to chronic inflammatory responses in organism, suggesting that overexpression of Set7 has a key effect on dysfunction of blood vessels in patients with type 2 diabetes [62]. In addition, foundational research found that mutation of H3K4 methyltransferase MLL2 may result in hyperglycemia and insulin resistance and progress to nonalcoholic fatty liver [63]. To sum up, with the gradual improvement of the research on histone methylation and the definition of functions of different sites of histone methylation, relationships between the aberrant methylation at all sites and tumors or other diseases will be demonstrated gradually. This will provide novel strategies and ideas for clinical diagnosis and treatment of tumors and other human diseases.

Table 2.

Histone lysine methyltransferases and related cancer or related developmental defects

H3K4 methyltransferases Related cancer Related developmental defects

SET1 prostate cancer, T-cell acute lymphoblastic leukemia, leukemia retarded growth
MLL acute myeloid leukemia, acute lymphoblastic leukemia, acute promyelocytic leukemia Wiedemann-Steiner Syndrome, Kabuki syndrome, hematopoiesis defects
SET7/9 gallbladder cancer impaired muscle differentiation
SMYD2 renal cell tumors, breast cancer, gastric cancer, acute lymphoblastic leukemia, esophageal squamous cell carcinoma impaired skeletal and cardiac muscles development
SMYD3 glioma, gastric cancer, prostate cancer, breast cancer, medullary thyroid carcinomas, hepatocellular cancer impaired early embryonic lineage commitment, impaired heart morphogenesis, impaired skeletal and cardiac muscles development
EZH2 prostate cancer, colorectal cancer, breast cancer, hepatocellular carcinoma, lung cancer, T-cell acute lymphoblastic leukemia, bladder cancer metabolic defects, DiGeorge or Velocardiofacial syndrome, cardiovascular defects, Weaver syndrome
ASH1 sinonasal neuroendocrine tumors, lung cancer, prostate cancer impaired ovule and anther development

H3K9 methyltransferases Related cancer Related developmental defects

SUV39h1 lung cancer, breast cancer, bladder cancer, acute myeloid leukemia, glioma, prostate cancer, hepatocellular carcinoma fetal lung defects, defects in terminal differentiation of the intestine, exocrine pancreas and retina
SUV39h2 glioma, hepatocellular carcinoma, prostate cancer, lung cancer fetal lung defects
G9a leukemia, breast cancer, head and neck squamous cell carcinoma, oral squamous cell carcinoma, endometrial cancer, lung cancer, hepatocellular carcinoma, glioma congenital and metabolic defects, growth retardation of embryos, calvaria defects, postnatal lethality, atrioventricular septal defects, retina defects
EZH2 prostate cancer, colorectal cancer, breast cancer, hepatocellular carcinoma, lung cancer, T-cell acute lymphoblastic leukemia, bladder cancer metabolic defects, DiGeorge or Velocardiofacial syndrome, cardiovascular defects, Weaver syndrome
GLP breast cancer, hepatocellular carcinoma, pancreatic ductal adenocarcinoma, leukemia, prostate cancer growth retardation of embryos, ossification defects of calvaria, postnatal lethality, atrioventricular septal defects
RIZ lung cancer, breast cancer, neuroblastoma, prostate cancer, leukemia, gastric and colorectal cancer N/A
SETDB1 lung cancer, hepatocellular carcinoma, breast cancer, prostate cancer, glioma congenital and metabolic defects, brain defects and early lethality
ASH1 sinonasal neuroendocrine tumors, lung cancer, prostate cancer impaired ovule and anther development

H3K27 methyltransferases Related cancer Related developmental defects

EZH1 leukemia, myeloproliferative neoplasms, breast cancer impaired differentiation of skeletal muscle cells
EZH2 prostate cancer, colorectal cancer, breast cancer, hepatocellular carcinoma, lung cancer, T-cell acute lymphoblastic leukemia, bladder cancer metabolic defects, DiGeorge or Velocardiofacial syndrome, cardiovascular defects, Weaver syndrome
NSD2 acute lymphoblastic leukemia, malignant lymphoproliferative diseases, prostate cancer Wolf-Hirschhorn syndrome
G9a leukemia, breast cancer, head and neck squamous cell carcinoma, oral squamous cell carcinoma, endometrial cancer, lung cancer, hepatocellular carcinoma, glioma congenital and metabolic defects, growth retardation of embryos, calvaria defects, postnatal lethality, atrioventricular septal defects, retina defects

H3K36 methyltransferases Related cancer Related developmental defects

NSD1 breast cancer, lung cancer, prostate cancers, acute myeloid leukemia, neuroblastoma and glioma Sotos or Weaver syndromes
NSD2 acute lymphoblastic leukemia, malignant lymphoproliferative diseases, prostate cancer Wolf-Hirschhorn syndrome
NSD3 breast cancer and acute myeloid leukemia N/A
ASH1 sinonasal neuroendocrine tumors, lung cancer, prostate cancer impaired ovule and anther development
SetD2 renal clear cell carcinoma, lymphoblastic leukemia, breast cancer, prostate cancer, lung cancer, glioma, thymic carcinoma, acute myeloid leukemia severe vascular defects, embryonic lethality, Sotos or Weaver syndromes

H3K79 methyltransferases Related cancer Related developmental defects

DOT1L acute myeloid leukemia, breast cancer, gastric cancer, colorectal cancer, prostate cancer impaired oocytes meiosis, growth impairment, angiogenesis defects in the yolk sac, cardiac dilation

H4K20 methyltransferases Related cancer Related developmental defects

SUV4-20H1 lung cancer N/A
SUV4-20H2 lung cancer, breast cancer, hepatocarcinogenesis N/A
NSD1 breast cancer, lung cancer, prostate cancers, acute myeloid leukemia, neuroblastoma and glioma Sotos or Weaver syndromes
ASH1 sinonasal neuroendocrine tumors, lung cancer, prostate cancer impaired ovule and anther development
SET9 multiple myeloma, prostate cancer N/A
SET8/PR-SET7 lung cancer, hepatocellular carcinoma, non-Hodgkin’s lymphomas, pancreatic cancer, cervical cancer, prostate cancer, breast cancer early embryonic lethality

H3K36-specific methyltransferases

Recent studies have found that methylation of H3K36 plays important roles in expression and regulation of genetic information, including gene transcription and regulation, alternative splicing, dosage compensation, DNA replication, repair and methylation, and gene inheritance and expression [64-68]. Therefore, this article is to lay special emphasis on types of H3K36-specific methyltransferases, patterns of H3K36 methylation, roles of H3K36 methylation in gene transcription and regulation, and the correlation between H3K36 methylation and human diseases.

Types and modifications of H3K36-specific methyltransferases

HMTs add methyl groups on the certain lysine or arginine residue through active adenosylmethionine in order to regulate the patterns of histone methylation [69]. To date, in vitro and in vivo studies have demonstrated that there are eight types of HMTs regulating H3K36 methylation levels in mammals, including nuclear receptor SET domain-containing 1 (NSD1), nuclear receptor SET domain-containing 2 (NSD2), SET domain containing 2 (SetD2), and other methyltransferases (i.e., nuclear receptor SET domain-containing 3 (NSD3), mesoderm-expressed 4 (MES-4), absent small and homeotic disks protein 1-like protein (ASH1L), SET domain and mariner transposase fusion (SETMAR), SET and MYND domain-containing 2 (SMYD2) and SET domain containing 3 (SETD3)[70]. All H3K36-specific methyltransferases contain highly conserved SET domains, but vary with patterns of H3K36 methylation, which mainly include mono-, di- and tri-methylation. In yeast, however, H3K36 can be mono-, di- and tri-methylated by Set2 simultaneously; in higher eukaryotic species, these patterns of methylation require coordination and division of mono-, di-methyltransferases and SET2-related tri-methyltransferases [71]. H3K36-specific methyltransferases possess a variety of domains which interacted with chromatin, i.e., PWWP domain interacted with methylated H3K36, and PHD fingers interacted with other methylated histone residues [72]. Novelly, for most species there is a domain interacted with RNA polymerase II (RNAPII) at the C-terminus of Set2, termed RNAP II subunit B1 (RPB1) [73].

Nuclear receptor SET domain-containing 1 (NSD1)

The initial discovery of nuclear receptor SET domain-containing 1 (NSD1), also known as androgen receptor activator protein (KMT3B), is because it can bind to nuclear steroid receptors, which, together with NSD2 and NSD3, belongs to the NSD protein family. After that, it is found that NSD1 can methylate H3K6 and H4K20 by its own SET domain. However, besides histones, it can methylate non-histones as well, such as p65 subunit of nuclear factor kappa B (NF-κB) [74]. To date, it has been found that the substrate of NSD1 is mainly unmethylated H3K36 or H3K36me1, which mainly methylate H3K36 in a mono- or di-methylated manner to produce H3K36me1 or H3K36me2 [75].

Nuclear receptor SET domain-containing 2 (NSD2)

Nuclear Receptor SET Domain-Containing 2 (NSD2) is also known as Wolf-Hirschhorn syndrome candidate 1/Multiple Myeloma SET domain (WHSC1/MMSET). Currently, the methylated site of NSD2 is still controversial. There are two major viewpoints. Some researchers believed that NSD2 could trimethylate H3K4, H3K27, H3K36 and H4K20 to produce H3K4me3, H3K27me3, H3K36me3 and H4K20me3 [76,77]; while some believed that NSD2 could dimethylate H4K20 and H3K36 to produce H4K20me2 and H3K36me2 [78]. However, a study of the protein spatial structure by Kuo et al. found that neither full-length NSD2 nor NSD-SET domain methylated H4K20, while research on protein functions found that NSD2 could mono- and di-methylate H3K36 to produce H3K3me and H3K36me2 [79]. It was also found that methylation of NSD2 depends on the substrate specificity. For example, NSD2 dimethylate H3K36 and H4K20 if the substrate is nucleosome; whereas, NSD2 methylate H4K44 if the substrate is histone octamer [80]. Hence, different sites and patterns of methylation of NSD2 determine the diversity of NSD2 regulating the gene transcription. That is to say,methylation of lysine residues can either activate or repress the gene transcription, showing that NSD2 represses the gene transcription by methylation of H3K27, H3K36 and H4K20 while methylation of H3K4 and H3K36 promotes the gene transcription [81]. To sum up, further research will be required to find out the biological functions of NSD2 and its exact regulation mechanism.

SET domain containing 2 (SetD2)

Human-derived histone methyltransferase SetD2, also known as huntingtin-interacting protein B (HYPB), is closely related to Huntington’s disease (HD). Located at 3p21.31, it is a protein with a molecular weight of approximately ~230 kDa. Its open reading frame (ORF) is ~6,186 bp, encoding 2,061 amino acid residues [82]. SetD2 also contains a vital active methyltransferase domain, the SET domain; beside it is the pre- and post-SET domain, respectively. The three domains are termed the associate with SET (AWS)-SET-Post SET domain, which is cysteine-rich and specifically methylates H3K36. Behind this domain is a cysteine- and proline-rich domain, termed the low charged region (LCR) domain, which activates the gene transcription and is highly conserved in vertebrates. There is a WW domain binding to the LCR domain, with two conservative tryptophan residues, which is currently accepted that its functions may be related to the regulation of the protein-protein interaction and that it exerts its effects by binding to the proline-rich region or SH3-binding motif. There is a 142 amino acid-long Set2-Rbp1 interacting (SRI) domain close to the C-terminus of SetD2, which interacts with phosphorylated RNA polymerase II (RNAPII) but not non-phosphorylated RNAPII [83-85] (Figure 1). Studies have found that SetD2 is related to the heterogeneous nuclear ribonucleo protein L (hnRNPL) and that partial knockout of hnRNPL results in a marked decrement in H3K36me3 level while levels of H3K36me1 and H3K36me2 are not affected, suggesting that SetD2 can specifically catalyze H3K36 and modify the H3K36me3 level [86].

Figure 1.

Figure 1

Schematic representation of SetD2 structure.

Specificity of H3K36 methylation

Structural specificity of H3K36-specific methyltransferase also determines its specific patterns of methylation, including mono-, di- and tri-methylation. However, so far, the specificity or difference of modification patterns of H3K36-specific methyltransferase is related to many factors in studies themselves, including assays of substrate specificity (peptide fragments, histones and nucleosomes), enzyme sources (the full-length and SET domain of the enzyme), and condition differences of the assay itself (antibody specificity and mass spectrometry). To date, 8 substrates of H3K36-specific methyltransferase and their patterns of methylation have been confirmed by means of experiments for loss-of-function and/or nucleosome-related substrates [70,87,88]. However, as for SETD3, its action characteristics are demonstrated by peptide fragments and core histones, their specific active sites and patterns of methylation should be further verified [89].

Roles of H3K36 methylation in regulation of gene expression

H3K36 methylation and transcriptional regulation

So far, many studies have confirmed in many ways that H3K36 methylation is inseparable from transcriptional activation [90-93]. In general, H3K36 methylation commonly occurs at the promoter and 3-terminus of the gene, from mono- to tri-methylation [81]. It has been found that H3K36 methylation is mostly present in the coding region of the transcriptional activator, which mainly affects the elongation of transcription. During transcription elongation, when phosphorylation occurs at Serine 2 of RNAPII C-terminal domain, the Paf1 elongation complex promotes Set2 to 5-ORF so that H3K36 is methylated. Two subunits of the Rpd3S histone deacetylase complex, Rco1 and Eaf3, can specifically recognize H3K36me. Once identified, Rpd3S can deacetylate the RNAPII-affected histone so as to prevent the latent transcription [94-96]. Moreover, regulation of H3K36 methylation levels by NSD1 and NSD3 is related to the transcription initiation. NSD1 binds to the upstream of the BMP4 promoter through RNAPII recruitment, regulating levels of H3K36me, H3K36me2 and H3K36me3 in organism [97]; NSD3 binds to LSD2 and BRD4 complexes or locate at the promoters or in interior regions of above complexes in order to facilitate H3K36 methylation and thus regulate the transcription initiation and elongation of genes [98].

H3K36 methylation and DNA replication/mismatch repair

DNA replication principally occurs during the S-phase. Initiation is defined by the origin recognition complex (ORC), which can recruit many factors before promoting DNA polymerase, such as CDC6, minichromosome maintenance proteins (MCMs) and CDC45. Studies of yeast conclude that H3K36 methylation plays an initiating role in DNA replication and that deletion of Set2 leads to delayed CDC45 recruitment [99]. In addition, H3K36 methylation is also closely associated with the checkpoint of DNA replication [100]. However, the exact molecular mechanisms by which H3K36 methylation regulates DNA replication during the S-phase have not been established.DNA mismatch repair can ensure high-fidelityreplication by correcting the mismatch generated during DNA replication. A study by Li F et al. reported that recruitment of in vivo mismatch recognition protein hMutSα (hMSH2-hMSH6) to chromatin needed H3K36me3, which could directly bind to the PWWP domain of hMSH6 to recruit hMutSα to chromatin [101]. Thus, a massive gathering of H3K36me3 during the G1 and S phases is to ensure hMutSα recruitment to chromatin before the mismatch occurs during DNA replication. The incidence of mutations will rise sharply when SetD2 is deleted. It follows that normal H3K36 methylationis of importance in DNA replication and mismatch repair.

Correlation between H3K36 methylation and human diseases

At present, research on H3K36 methylation has found that abnormities of patterns or levels of H3K36 methylation in higher eukaryotes may result in tumor genesis and progression, developmental defects, cerebral gigantism syndrome, ovarian fibroma, Wolf-Hirschhorn syndrome, etc [102-107] (Figure 2).

Figure 2.

Figure 2

Schematic representation of H3K36 methyltransfearses and related human diseases.

H3K36 methylation and tumor genesis/progression

To date, many studies have found that the members of the NSD family function as tumor suppressors of many cancers. For example, NSD1 is closely related to breast, lung and prostate cancers, acute myeloid leukemia and refractory anemia [108-111]; NSD2 is closely related to acute lymphoblastic leukemia, malignant lymphoproliferative diseases and prostate cancer [112,113]; whereas NSD3 is closely related to breast cancer and acute myeloid leukemia [114,115]. Recently, it was reported in Blood, a peer-reviewed medical journal, that co-expression of NUP98/NSD1 and FLT3/ITD was closely related to poor prognosis in acute myeloid leukemia [116]. Berdasco M et al. found inactivation of NSD1 in neuroblastoma and glioma that leads to abnormal H3K36 methylation levels and thus overgrowth and proliferation of tumor cells, which might be one of the prognostic markers for treating these two diseases [117]. Zhao Q et al. identified genomic alterations in the breast cancer cell line HCC1954 using high-throughput transcriptome sequencing and concluded that rearrangements and mutations occurred in the gene of NSD1, suggesting that NSD1 mutations are likely to play a key role in genesis of breast cancer [118]. Yang P et al. found that NSD2 was critical for cytokine-induced recruitment of NF-κB and acetyltransferase p300 and histone acetylation, and more importantly, they also found that NSD2 was overexpressed in prostate cancer tissues and its overexpression correlated with the activation of NF-κB signal pathway. Furthermore, they also indicated that NSD2 expression was induced by tumor necrosis factor alpha (TNF-α) and interleukin-6 (IL-60), playing a crucial role in tumor growth [119]. These results suggested that NSD2 played critical roles in cancer cell proliferation and tumor growth and progression.

Furthermore, current studies have found that H3K36 trimethyltransferase SetD2 mutates and is expressed aberrantly in renal clear cell carcinoma, lymphoblastic leukemia, breast and prostate cancers, gliomas and thymic carcinoma. To date, studies have suggested that SetD2 is a tumor suppressor gene [120-126]. SetD2 was initially found to mutate in patients with renal clear cell carcinoma and show a marked decrease in mRNA level so as to be identified as a novel tumor suppressor [120]. A most recent study found detrimental mutations in SetD2 in tumor tissues of patients with non-small cell lung cancer (NSCLC), speculating that SetD2 is promising to be one of the clinical diagnostic or prognostic markers for NSCLC in the future [127]. Moreover, it was reported that frameshift or nonsense mutation in many epigenetic regulators, including SetD2, was noted in patients with acute myeloid leukemia, with an incidence of 12%, suggesting epigenetic regulators will be one of the important targets for treating this disease [128].

H3K36 methylation and developmental defects

Apart from tumor genesis and progression, it has been widely reported that abnormity of H3K36 methylation is also related to developmental defects. Research found that: SetD2-deficient embryos failed to undergo normal implantation at the blastocyst stage while some lineage specific factors, including Eomes, Elf5 and Sox2, were distinctly reduced in SetD2-deficient embryos and that several imprinted genes (Mest, Peg3, Snrpn and Meg3) were aberrantly expressed, speculating that H3K36 trimethylation regulated by histone methyltransferase SetD2 plays an important role in maintaining normal embryo implantation [129]. In addition, Chen Zhu (an academician of the Institute of Health Sciences, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, China) and colleagues found that the H3K36me3 level was reduced in fibroblasts of SetD2-deficient embryos but there was no distinct change in levels of H3K36me and H3K36me2, and that all homozygous embryos of SetD2-knockout mice were lethal at E11-E11.5. Postmortem and pathological staining showed that embryonic growth retardation, incomplete chorio-allantoic fusion and araphia were noted in SetD2-deficient mice. Most notably, vascular developmental defects were noted in SetD2-deficient embryos during vascular development, including abnormal vasodilatation, morphological abnormalities of mesodermal epithelial cells, and increased spacing between epithelial cells. Microarray assay for SetD2-deficient embryos indicated that 10 angiogenesis-related secreted factors (Ang, Angptl3, Angptl6, CTGF, Cyr61, Igf1, Pdgfc, Plg, Serpinf1 and VEGFb) and 8 membrane proteins (Cav1, Flt1, Gja4, Lama1, Lama4, Rhob, Sema3c and Serpine3) were significantly expressed aberrantly [130]. It can be concluded that the mechanism by which SetD2 deficiency results in embryonic lethality may be related to anomalous blood supply due to embryonic abnormal angiogenesis.

H3K36 methylation and other diseases

Apart from tumor genesis and developmental defects, more and more studies using gene sequencing have found that NSD1 mutates in patients with cerebral gigantism syndrome and ovarian fibroma, which may be a promising new diagnostic marker in the future [131]. Furthermore, it has been found that NSD2-deficient mice may develop Wolf-Hirschhorn syndrome manifesting as growth retardation and congenital vascular malformation. In addition, Wolf-Hirschhorn syndrome is significantly more severe in NSD2-and-Nkx2.5-knockout heterozygous mice than in NSD2-knockout mice, speculating that Wolf-Hirschhorn syndrome is regulated by the interaction between NSD2 and transcription factor Nkx2.5 [132]. Nevertheless, SetD2, also known as huntingtin-interacting protein B, was initially discovered in Huntington’s disease, a neurodegenerative disorder. SetD2 is identified as a member of the huntingtin-interacting protein family because there is a WW motif in it.

Conclusions

Histone methylation is one of the most important fields in epigenetics. Therefore, further investigation of the regulatory effects and mechanisms of histone methylation on physiological functions (gene transcription and regulation, biological inheritance, etc.), and perfection of roles of histone methylation in human tumorigenesis, growth/development and immunoregulation by bioinformatics techniques will provide important clues and targets for diagnosis, prevention and treatment of human diseases. For example, it is very prospective for developing new drugs to investigate active sites or substrates of HMTs, their characteristics of domains, and specific recognition of their proteins or polypeptides.

Acknowledgements

We are grateful to all of the members of the Department of Cardiology and the Cardiovascular Research Institute of Renmin Hospital of Wuhan University for their expert advice. This work was supported by grants from the National Natural Science Foundation of China (No. 81170307) and Specialized Research Fund for the Doctoral Program of Higher Education (No. 20120141110013).

Disclosure of conflict of interest

The authors have declared that no conflict of interest exists.

References

  • 1.Tessarz P, Kouzarides T. Histone core modifications regulating nucleosome structure and dynamics. Nat Rev Mol Cell Biol. 2014;15:703–708. doi: 10.1038/nrm3890. [DOI] [PubMed] [Google Scholar]
  • 2.Lavelle C, Foray N. Chromatin structure and radiation-induced DNA damage: from structural biology to radiobiology. Int J Biochem Cell Biol. 2014;49:84–97. doi: 10.1016/j.biocel.2014.01.012. [DOI] [PubMed] [Google Scholar]
  • 3.Fullgrabe J, Hajji N, Joseph B. Cracking the death code: apoptosis-related histone modifications. Cell Death Differ. 2010;17:1238–1243. doi: 10.1038/cdd.2010.58. [DOI] [PubMed] [Google Scholar]
  • 4.Nikitina T, Wang D, Gomberg M, Grigoryev SA, Zhurkin VB. Combined micrococcal nuclease and exonuclease III digestion reveals precise positions of the nucleosome core/linker junctions: implications for high-resolution nucleosome mapping. J Mol Biol. 2013;425:1946–1960. doi: 10.1016/j.jmb.2013.02.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Caterino TL, Fang H, Hayes JJ. Nucleosome linker DNA contacts and induces specific folding of the intrinsically disordered H1 carboxylterminal domain. Mol Cell Biol. 2011;31:2341–2348. doi: 10.1128/MCB.05145-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Chen HP, Zhao YT, Zhao TC. Histone deacetylases and mechanisms of regulation of gene expression. Crit Rev Oncog. 2015;20:35–47. doi: 10.1615/critrevoncog.2015012997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Geranton SM, Tochiki KK. Regulation of gene expression and pain states by epigenetic mechanisms. Prog Mol Biol Transl Sci. 2015;131:147–183. doi: 10.1016/bs.pmbts.2014.11.012. [DOI] [PubMed] [Google Scholar]
  • 8.Lewis BA, Hanover JA. O-GlcNAc and the epigenetic regulation of gene expression. J Biol Chem. 2014;289:34440–34448. doi: 10.1074/jbc.R114.595439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Song ZT, Sun L, Lu SJ, Tian Y, Ding Y, Liu JX. Transcription factor interaction with COMPASSlike complex regulates histone H3K4 trimethylation for specific gene expression in plants. Proc Natl Acad Sci U S A. 2015;112:2900–2905. doi: 10.1073/pnas.1419703112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Leung D, Jung I, Rajagopal N, Schmitt A, Selvaraj S, Lee AY, Yen CA, Lin S, Lin Y, Qiu Y, Xie W, Yue F, Hariharan M, Ray P, Kuan S, Edsall L, Yang H, Chi NC, Zhang MQ, Ecker JR, Ren B. Integrative analysis of haplotype-resolved epigenomes across human tissues. Nature. 2015;518:350–354. doi: 10.1038/nature14217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Gatti M, Pinato S, Maiolica A, Rocchio F, Prato MG, Aebersold R, Penengo L. RNF168 promotes noncanonical K27 ubiquitination to signal DNA damage. Cell Rep. 2015;10:226–238. doi: 10.1016/j.celrep.2014.12.021. [DOI] [PubMed] [Google Scholar]
  • 12.Sakamoto A, Hino S, Nagaoka K, Anan K, Takase R, Matsumori H, Ojima H, Kanai Y, Arita K, Nakao M. Lysine Demethylase LSD1 Coordinates Glycolytic and Mitochondrial Metabolism in Hepatocellular Carcinoma Cells. Cancer Res. 2015;75:1445–1456. doi: 10.1158/0008-5472.CAN-14-1560. [DOI] [PubMed] [Google Scholar]
  • 13.Van Rechem C, Whetstine JR. Examining the impact of gene variants on histone lysine methylation. Biochim Biophys Acta. 2014;1839:1463–1476. doi: 10.1016/j.bbagrm.2014.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Di Lorenzo A, Bedford MT. Histone arginine methylation. FEBS Lett. 2011;585:2024–2031. doi: 10.1016/j.febslet.2010.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Thakur JK, Malik MR, Bhatt V, Reddy MK, Sopory SK, Tyagi AK, Khurana JP. A POLYCOMB group gene of rice (Oryza sativa L. subspecies indica), OsiEZ1, codes for a nuclearlocalized protein expressed preferentially in young seedlings and during reproductive development. Gene. 2003;314:1–13. doi: 10.1016/s0378-1119(03)00723-6. [DOI] [PubMed] [Google Scholar]
  • 16.Min J, Zhang X, Cheng X, Grewal SI, Xu RM. Structure of the SET domain histone lysine methyltransferase Clr4. Nat Struct Biol. 2002;9:828–832. doi: 10.1038/nsb860. [DOI] [PubMed] [Google Scholar]
  • 17.Qian Y, Xi Y, Cheng B, Zhu S, Kan X. Identification and characterization of the SET domain gene family in maize. Mol Biol Rep. 2014;41:1341–1354. doi: 10.1007/s11033-013-2980-x. [DOI] [PubMed] [Google Scholar]
  • 18.Binda O. On your histone mark, SET, methylate! Epigenetics. 2013;8:457–463. doi: 10.4161/epi.24451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Qian C, Zhou MM. SET domain protein lysine methyltransferases: Structure, specificity and catalysis. Cell Mol Life Sci. 2006;63:2755–2763. doi: 10.1007/s00018-006-6274-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Del Rizzo PA, Trievel RC. Substrate and product specificities of SET domain methyltransferases. Epigenetics. 2011;6:1059–1067. doi: 10.4161/epi.6.9.16069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Dillon SC, Zhang X, Trievel RC, Cheng X. The SET-domain protein superfamily: protein lysine methyltransferases. Genome Biol. 2005;6:227. doi: 10.1186/gb-2005-6-8-227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Yeates TO. Structures of SET domain proteins: protein lysine methyltransferases make their mark. Cell. 2002;111:5–7. doi: 10.1016/s0092-8674(02)01010-3. [DOI] [PubMed] [Google Scholar]
  • 23.Zhang X, Bruice TC. Enzymatic mechanism and product specificity of SET-domain protein lysine methyltransferases. Proc Natl Acad Sci U S A. 2008;105:5728–5732. doi: 10.1073/pnas.0801788105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Valencia-Morales Mdel P, Camas-Reyes JA, Cabrera-Ponce JL, Alvarez-Venegas R. The Arabidopsis thaliana SET-domain-containing protein ASHH1/SDG26 interacts with itself and with distinct histone lysine methyltransferases. J Plant Res. 2012;125:679–692. doi: 10.1007/s10265-012-0485-7. [DOI] [PubMed] [Google Scholar]
  • 25.Hu P, Wang S, Zhang Y. How do SET-domain protein lysine methyltransferases achieve the methylation state specificity? Revisited by Ab initio QM/MM molecular dynamics simulations. J Am Chem Soc. 2008;130:3806–3813. doi: 10.1021/ja075896n. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Wilson JR, Jing C, Walker PA, Martin SR, Howell SA, Blackburn GM, Gamblin SJ, Xiao B. Crystal structure and functional analysis of the histone methyltransferase SET7/9. Cell. 2002;111:105–115. doi: 10.1016/s0092-8674(02)00964-9. [DOI] [PubMed] [Google Scholar]
  • 27.Zhang X, Tamaru H, Khan SI, Horton JR, Keefe LJ, Selker EU, Cheng X. Structure of the Neurospora SET domain protein DIM-5, a histone H3 lysine methyltransferase. Cell. 2002;111:117–127. doi: 10.1016/s0092-8674(02)00999-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Zhang X, Yang Z, Khan SI, Horton JR, Tamaru H, Selker EU, Cheng X. Structural basis for the product specificity of histone lysine methyltransferases. Mol Cell. 2003;12:177–185. doi: 10.1016/s1097-2765(03)00224-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.An S, Yeo KJ, Jeon YH, Song JJ. Crystal structure of the human histone methyltransferase ASH1L catalytic domain and its implications for the regulatory mechanism. J Biol Chem. 2011;286:8369–8374. doi: 10.1074/jbc.M110.203380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Huang S, Shao G, Liu L. The PR domain of the Rb-binding zinc finger protein RIZ1 is a protein binding interface and is related to the SET domain functioning in chromatin-mediated gene expression. J Biol Chem. 1998;273:15933–15939. doi: 10.1074/jbc.273.26.15933. [DOI] [PubMed] [Google Scholar]
  • 31.Yakubov E, Dinerman P, Kuperstein F, Saban S, Yavin E. Improved representation of gene markers on microarray by PCR-Select subtracted cDNA targets. Brain Res Mol Brain Res. 2005;137:110–118. doi: 10.1016/j.molbrainres.2005.02.019. [DOI] [PubMed] [Google Scholar]
  • 32.Xu Q, Dunbrack RL Jr. Assignment of protein sequences to existing domain and family classification systems: Pfam and the PDB. Bioinformatics. 2012;28:2763–2772. doi: 10.1093/bioinformatics/bts533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Spellmon N, Holcomb J, Trescott L, Sirinupong N, Yang Z. Structure and function of SET and MYND domain-containing proteins. Int J Mol Sci. 2015;16:1406–1428. doi: 10.3390/ijms16011406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Berr A, Shafiq S, Pinon V, Dong A, Shen WH. The trxG family histone methyltransferase SET DOMAIN GROUP 26 promotes flowering via a distinctive genetic pathway. Plant J. 2015;81:316–328. doi: 10.1111/tpj.12729. [DOI] [PubMed] [Google Scholar]
  • 35.Alvarez-Venegas R. Bacterial SET domain proteins and their role in eukaryotic chromatin modification. Front Genet. 2014;5:65. doi: 10.3389/fgene.2014.00065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Congdon LM, Sims JK, Tuzon CT, Rice JC. The PR-Set7 binding domain of Riz1 is required for the H4K20me1-H3K9me1 trans-tail ‘histone code’ and Riz1 tumor suppressor function. Nucleic Acids Res. 2014;42:3580–3589. doi: 10.1093/nar/gkt1377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Wu H, Siarheyeva A, Zeng H, Lam R, Dong A, Wu XH, Li Y, Schapira M, Vedadi M, Min J. Crystal structures of the human histone H4K20 methyltransferases SUV420H1 and SUV420H2. FEBS Lett. 2013;587:3859–3868. doi: 10.1016/j.febslet.2013.10.020. [DOI] [PubMed] [Google Scholar]
  • 38.Schultz DC, Ayyanathan K, Negorev D, Maul GG, Rauscher FJ 3rd. SETDB1: a novel KAP-1-associated histone H3, lysine 9-specific methyltransferase that contributes to HP1-mediated silencing of euchromatic genes by KRAB zinc-finger proteins. Genes Dev. 2002;16:919–932. doi: 10.1101/gad.973302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Du TT, Xu PF, Dong ZW, Fan HB, Jin Y, Dong M, Chen Y, Pan WJ, Ren RB, Liu TX, Deng M, Huang QH. Setdb2 controls convergence and extension movements during zebrafish gastrulation by transcriptional regulation of dvr1. Dev Biol. 2014;392:233–244. doi: 10.1016/j.ydbio.2014.05.022. [DOI] [PubMed] [Google Scholar]
  • 40.Rothbart SB, Dickson BM, Strahl BD. From histones to ribosomes: a chromatin regulator tangoes with translation. Cancer Discov. 2015;5:228–230. doi: 10.1158/2159-8290.CD-15-0073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Wozniak GG, Strahl BD. Hitting the ‘mark’: interpreting lysine methylation in the context of active transcription. Biochim Biophys Acta. 2014;1839:1353–1361. doi: 10.1016/j.bbagrm.2014.03.002. [DOI] [PubMed] [Google Scholar]
  • 42.Lan F, Shi Y. Epigenetic regulation: methylation of histone and non-histone proteins. Sci China C Life Sci. 2009;52:311–322. doi: 10.1007/s11427-009-0054-z. [DOI] [PubMed] [Google Scholar]
  • 43.Herz HM, Garruss A, Shilatifard A. SET for life: biochemical activities and biological functions of SET domain-containing proteins. Trends Biochem Sci. 2013;38:621–639. doi: 10.1016/j.tibs.2013.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Niwa H, Handa N, Tomabechi Y, Honda K, Toyama M, Ohsawa N, Shirouzu M, Kagechika H, Hirano T, Umehara T, Yokoyama S. Structures of histone methyltransferase SET7/9 in complexes with adenosylmethionine derivatives. Acta Crystallogr D Biol Crystallogr. 2013;69:595–602. doi: 10.1107/S0907444912052092. [DOI] [PubMed] [Google Scholar]
  • 45.Liu N, Zhang Z, Wu H, Jiang Y, Meng L, Xiong J, Zhao Z, Zhou X, Li J, Li H, Zheng Y, Chen S, Cai T, Gao S, Zhu B. Recognition of H3K9 methylation by GLP is required for efficient establishment of H3K9 methylation, rapid target gene repression, and mouse viability. Genes Dev. 2015;29:379–393. doi: 10.1101/gad.254425.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Inoue S, Honma K, Mochizuki K, Goda T. Induction of histone H3K4 methylation at the promoter, enhancer, and transcribed regions of the Si and Sglt1 genes in rat jejunum in response to a high-starch/low-fat diet. Nutrition. 2015;31:366–372. doi: 10.1016/j.nut.2014.07.017. [DOI] [PubMed] [Google Scholar]
  • 47.Lillycrop KA, Burdge GC. Environmental challenge, epigenetic plasticity and the induction of altered phenotypes in mammals. Epigenomics. 2014;6:623–636. doi: 10.2217/epi.14.51. [DOI] [PubMed] [Google Scholar]
  • 48.Sadakierska-Chudy A, Filip M. A comprehensive view of the epigenetic landscape. Part II: Histone post-translational modification, nucleosome level, and chromatin regulation by ncRNAs. Neurotox Res. 2015;27:172–197. doi: 10.1007/s12640-014-9508-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Kim EJ, Ma X, Cerutti H. Gene silencing in microalgae: mechanisms and biological roles. Bioresour Technol. 2015;184:23–32. doi: 10.1016/j.biortech.2014.10.119. [DOI] [PubMed] [Google Scholar]
  • 50.Hu J, Chen S, Kong X, Zhu K, Cheng S, Zheng M, Jiang H, Luo C. Interaction between DNA/histone methyltransferases and their inhibitors. Curr Med Chem. 2015;22:360–372. doi: 10.2174/0929867321666141106114538. [DOI] [PubMed] [Google Scholar]
  • 51.Wakeman TP, Wang Q, Feng J, Wang XF. Bat3 facilitates H3K79 dimethylation by DOT1L and promotes DNA damage-induced 53BP1 foci at G1/G2 cell-cycle phases. EMBO J. 2012;31:2169–2181. doi: 10.1038/emboj.2012.50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Garcia-Cao M, O’Sullivan R, Peters AH, Jenuwein T, Blasco MA. Epigenetic regulation of telomere length in mammalian cells by the Suv39h1 and Suv39h2 histone methyltransferases. Nat Genet. 2004;36:94–99. doi: 10.1038/ng1278. [DOI] [PubMed] [Google Scholar]
  • 53.Muramatsu D, Singh PB, Kimura H, Tachibana M, Shinkai Y. Pericentric heterochromatin generated by HP1 protein interaction-defective histone methyltransferase Suv39h1. J Biol Chem. 2013;288:25285–25296. doi: 10.1074/jbc.M113.470724. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 54.Kalish JM, Jiang C, Bartolomei MS. Epigenetics and imprinting in human disease. Int J Dev Biol. 2014;58:291–298. doi: 10.1387/ijdb.140077mb. [DOI] [PubMed] [Google Scholar]
  • 55.Weaver JR, Bartolomei MS. Chromatin regulators of genomic imprinting. Biochim Biophys Acta. 2014;1839:169–177. doi: 10.1016/j.bbagrm.2013.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Chaumeil J, Waters PD, Koina E, Gilbert C, Robinson TJ, Graves JA. Evolution from XIST-independent to XIST-controlled X-chromosome inactivation: epigenetic modifications in distantly related mammals. PLoS One. 2011;6:e19040. doi: 10.1371/journal.pone.0019040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.He L, Yu JX, Liu L, Buyse IM, Wang MS, Yang QC, Nakagawara A, Brodeur GM, Shi YE, Huang S. RIZ1, but not the alternative RIZ2 product of the same gene, is underexpressed in breast cancer, and forced RIZ1 expression causes G2-M cell cycle arrest and/or apoptosis. Cancer Res. 1998;58:4238–4244. [PubMed] [Google Scholar]
  • 58.Canote R, Du Y, Carling T, Tian F, Peng Z, Huang S. The tumor suppressor gene RIZ in cancer gene therapy (review) Oncol Rep. 2002;9:57–60. [PubMed] [Google Scholar]
  • 59.Deng Q, Huang S. PRDM5 is silenced in human cancers and has growth suppressive activities. Oncogene. 2004;23:4903–4910. doi: 10.1038/sj.onc.1207615. [DOI] [PubMed] [Google Scholar]
  • 60.Czvitkovich S, Sauer S, Peters AH, Deiner E, Wolf A, Laible G, Opravil S, Beug H, Jenuwein T. Over-expression of the SUV39H1 histone methyltransferase induces altered proliferation and differentiation in transgenic mice. Mech Dev. 2001;107:141–153. doi: 10.1016/s0925-4773(01)00464-6. [DOI] [PubMed] [Google Scholar]
  • 61.Yokoyama A, Wang Z, Wysocka J, Sanyal M, Aufiero DJ, Kitabayashi I, Herr W, Cleary ML. Leukemia proto-oncoprotein MLL forms a SET1-like histone methyltransferase complex with menin to regulate Hox gene expression. Mol Cell Biol. 2004;24:5639–5649. doi: 10.1128/MCB.24.13.5639-5649.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Paneni F, Costantino S, Battista R, Castello L, Capretti G, Chiandotto S, Scavone G, Villano A, Pitocco D, Lanza G, Volpe M, Luscher TF, Cosentino F. Adverse epigenetic signatures by histone methyltransferase Set7 contribute to vascular dysfunction in patients with type 2 diabetes mellitus. Circ Cardiovasc Genet. 2015;8:150–158. doi: 10.1161/CIRCGENETICS.114.000671. [DOI] [PubMed] [Google Scholar]
  • 63.Goldsworthy M, Absalom NL, Schroter D, Matthews HC, Bogani D, Moir L, Long A, Church C, Hugill A, Anstee QM, Goldin R, Thursz M, Hollfelder F, Cox RD. Mutations in Mll2, an H3K4 methyltransferase, result in insulin resistance and impaired glucose tolerance in mice. PLoS One. 2013;8:e61870. doi: 10.1371/journal.pone.0061870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Jha DK, Strahl BD. An RNA polymerase IIcoupled function for histone H3K36 methylation in checkpoint activation and DSB repair. Nat Commun. 2014;5:3965. doi: 10.1038/ncomms4965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Kimura H. Histone modifications for human epigenome analysis. J Hum Genet. 2013;58:439–445. doi: 10.1038/jhg.2013.66. [DOI] [PubMed] [Google Scholar]
  • 66.Hossain MA, Chung C, Pradhan SK, Johnson TL. The yeast cap binding complex modulates transcription factor recruitment and establishes proper histone H3K36 trimethylation during active transcription. Mol Cell Biol. 2013;33:785–799. doi: 10.1128/MCB.00947-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Venkatesh S, Smolle M, Li H, Gogol MM, Saint M, Kumar S, Natarajan K, Workman JL. Set2 methylation of histone H3 lysine 36 suppresses histone exchange on transcribed genes. Nature. 2012;489:452–455. doi: 10.1038/nature11326. [DOI] [PubMed] [Google Scholar]
  • 68.Maltby VE, Martin BJ, Schulze JM, Johnson I, Hentrich T, Sharma A, Kobor MS, Howe L. Histone H3 lysine 36 methylation targets the Isw1b remodeling complex to chromatin. Mol Cell Biol. 2012;32:3479–3485. doi: 10.1128/MCB.00389-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Del Rizzo PA, Trievel RC. Molecular basis for substrate recognition by lysine methyltransferases and demethylases. Biochim Biophys Acta. 2014;1839:1404–1415. doi: 10.1016/j.bbagrm.2014.06.008. [DOI] [PubMed] [Google Scholar]
  • 70.Wagner EJ, Carpenter PB. Understanding the language of Lys36 methylation at histone H3. Nat Rev Mol Cell Biol. 2012;13:115–126. doi: 10.1038/nrm3274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Xu G, Liu G, Xiong S, Liu H, Chen X, Zheng B. The histone methyltransferase Smyd2 is a negative regulator of macrophage activation by suppressing interleukin 6 (IL-6) and tumor necrosis factor alpha (TNF-alpha) production. J Biol Chem. 2015;290:5414–5423. doi: 10.1074/jbc.M114.610345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.van Nuland R, van Schaik FM, Simonis M, van Heesch S, Cuppen E, Boelens R, Timmers HM, van Ingen H. Nucleosomal DNA binding drives the recognition of H3K36-methylated nucleosomes by the PSIP1-PWWP domain. Epigenetics Chromatin. 2013;6:12. doi: 10.1186/1756-8935-6-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Endo H, Nakabayashi Y, Kawashima S, Enomoto T, Seki M, Horikoshi M. Nucleosome surface containing nucleosomal DNA entry/exit site regulates H3-K36me3 via association with RNA polymerase II and Set2. Genes Cells. 2012;17:65–81. doi: 10.1111/j.1365-2443.2011.01573.x. [DOI] [PubMed] [Google Scholar]
  • 74.Sakhon OS, Victor KA, Choy A, Tsuchiya T, Eulgem T, Pedra JH. NSD1 mitigates caspase-1 activation by listeriolysin O in macrophages. PLoS One. 2013;8:e75911. doi: 10.1371/journal.pone.0075911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Wang GG, Cai L, Pasillas MP, Kamps MP. NUP98-NSD1 links H3K36 methylation to Hox-A gene activation and leukaemogenesis. Nat Cell Biol. 2007;9:804–812. doi: 10.1038/ncb1608. [DOI] [PubMed] [Google Scholar]
  • 76.Popovic R, Martinez-Garcia E, Giannopoulou EG, Zhang Q, Zhang Q, Ezponda T, Shah MY, Zheng Y, Will CM, Small EC, Hua Y, Bulic M, Jiang Y, Carrara M, Calogero RA, Kath WL, Kelleher NL, Wang JP, Elemento O, Licht JD. Histone methyltransferase MMSET/NSD2 alters EZH2 binding and reprograms the myeloma epigenome through global and focal changes in H3K36 and H3K27 methylation. PLoS Genet. 2014;10:e1004566. doi: 10.1371/journal.pgen.1004566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Zheng Y, Sweet SM, Popovic R, Martinez-Garcia E, Tipton JD, Thomas PM, Licht JD, Kelleher NL. Total kinetic analysis reveals how combinatorial methylation patterns are established on lysines 27 and 36 of histone H3. Proc Natl Acad Sci U S A. 2012;109:13549–13554. doi: 10.1073/pnas.1205707109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Morishita M, Mevius D, di Luccio E. In vitro histone lysine methylation by NSD1, NSD2/MMSET/WHSC1 and NSD3/WHSC1L. BMC Struct Biol. 2014;14:25. doi: 10.1186/s12900-014-0025-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Kuo AJ, Cheung P, Chen K, Zee BM, Kioi M, Lauring J, Xi Y, Park BH, Shi X, Garcia BA, Li W, Gozani O. NSD2 links dimethylation of histone H3 at lysine 36 to oncogenic programming. Mol Cell. 2011;44:609–620. doi: 10.1016/j.molcel.2011.08.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Morishita M, di Luccio E. Cancers and the NSD family of histone lysine methyltransferases. Biochim Biophys Acta. 2011;1816:158–163. doi: 10.1016/j.bbcan.2011.05.004. [DOI] [PubMed] [Google Scholar]
  • 81.Bannister AJ, Schneider R, Myers FA, Thorne AW, Crane-Robinson C, Kouzarides T. Spatial distribution of di- and tri-methyl lysine 36 of histone H3 at active genes. J Biol Chem. 2005;280:17732–17736. doi: 10.1074/jbc.M500796200. [DOI] [PubMed] [Google Scholar]
  • 82.Edmunds JW, Mahadevan LC, Clayton AL. Dynamic histone H3 methylation during gene induction: HYPB/Setd2 mediates all H3K36 trimethylation. EMBO J. 2008;27:406–420. doi: 10.1038/sj.emboj.7601967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Pfister SX, Ahrabi S, Zalmas LP, Sarkar S, Aymard F, Bachrati CZ, Helleday T, Legube G, La Thangue NB, Porter AC, Humphrey TC. SETD2-dependent histone H3K36 trimethylation is required for homologous recombination repair and genome stability. Cell Rep. 2014;7:2006–2018. doi: 10.1016/j.celrep.2014.05.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Gao YG, Yang H, Zhao J, Jiang YJ, Hu HY. Autoinhibitory structure of the WW domain of HYPB/SETD2 regulates its interaction with the proline-rich region of huntingtin. Structure. 2014;22:378–386. doi: 10.1016/j.str.2013.12.005. [DOI] [PubMed] [Google Scholar]
  • 85.Sun XJ, Wei J, Wu XY, Hu M, Wang L, Wang HH, Zhang QH, Chen SJ, Huang QH, Chen Z. Identification and characterization of a novel human histone H3 lysine 36-specific methyltransferase. J Biol Chem. 2005;280:35261–35271. doi: 10.1074/jbc.M504012200. [DOI] [PubMed] [Google Scholar]
  • 86.Yuan W, Xie J, Long C, Erdjument-Bromage H, Ding X, Zheng Y, Tempst P, Chen S, Zhu B, Reinberg D. Heterogeneous nuclear ribonucleoprotein L Is a subunit of human KMT3a/Set2 complex required for H3 Lys-36 trimethylation activity in vivo. J Biol Chem. 2009;284:15701–15707. doi: 10.1074/jbc.M808431200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Park D, Shivram H, Iyer VR. Chd1 co-localizes with early transcription elongation factors independently of H3K36 methylation and releases stalled RNA polymerase II at introns. Epigenetics Chromatin. 2014;7:32. doi: 10.1186/1756-8935-7-32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Eram MS, Bustos SP, Lima-Fernandes E, Siarheyeva A, Senisterra G, Hajian T, Chau I, Duan S, Wu H, Dombrovski L, Schapira M, Arrowsmith CH, Vedadi M. Trimethylation of histone H3 lysine 36 by human methyltransferase PRDM9 protein. J Biol Chem. 2014;289:12177–12188. doi: 10.1074/jbc.M113.523183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Kim DW, Kim KB, Kim JY, Seo SB. Characterization of a novel histone H3K36 methyltransferase setd3 in zebrafish. Biosci Biotechnol Biochem. 2011;75:289–294. doi: 10.1271/bbb.100648. [DOI] [PubMed] [Google Scholar]
  • 90.Ginsburg DS, Anlembom TE, Wang J, Patel SR, Li B, Hinnebusch AG. NuA4 links methylation of histone H3 lysines 4 and 36 to acetylation of histones H4 and H3. J Biol Chem. 2014;289:32656–32670. doi: 10.1074/jbc.M114.585588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Lee YF, Nimura K, Lo WN, Saga K, Kaneda Y. Histone H3 lysine 36 methyltransferase Whsc1 promotes the association of Runx2 and p300 in the activation of bone-related genes. PLoS One. 2014;9:e106661. doi: 10.1371/journal.pone.0106661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Diao YF, Oqani RK, Li XX, Lin T, Kang JW, Jin DI. Changes in histone H3 lysine 36 methylation in porcine oocytes and preimplantation embryos. PLoS One. 2014;9:e100205. doi: 10.1371/journal.pone.0100205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Wen H, Li Y, Xi Y, Jiang S, Stratton S, Peng D, Tanaka K, Ren Y, Xia Z, Wu J, Li B, Barton MC, Li W, Li H, Shi X. ZMYND11 links histone H3.3K36me3 to transcription elongation and tumour suppression. Nature. 2014;508:263–268. doi: 10.1038/nature13045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Li B, Gogol M, Carey M, Lee D, Seidel C, Workman JL. Combined action of PHD and chromo domains directs the Rpd3S HDAC to transcribed chromatin. Science. 2007;316:1050–1054. doi: 10.1126/science.1139004. [DOI] [PubMed] [Google Scholar]
  • 95.Joshi AA, Struhl K. Eaf3 chromodomain interaction with methylated H3-K36 links histone deacetylation to Pol II elongation. Mol Cell. 2005;20:971–978. doi: 10.1016/j.molcel.2005.11.021. [DOI] [PubMed] [Google Scholar]
  • 96.Carrozza MJ, Li B, Florens L, Suganuma T, Swanson SK, Lee KK, Shia WJ, Anderson S, Yates J, Washburn MP, Workman JL. Histone H3 methylation by Set2 directs deacetylation of coding regions by Rpd3S to suppress spurious intragenic transcription. Cell. 2005;123:581–592. doi: 10.1016/j.cell.2005.10.023. [DOI] [PubMed] [Google Scholar]
  • 97.Lucio-Eterovic AK, Singh MM, Gardner JE, Veerappan CS, Rice JC, Carpenter PB. Role for the nuclear receptor-binding SET domain protein 1 (NSD1) methyltransferase in coordinating lysine 36 methylation at histone 3 with RNA polymerase II function. Proc Natl Acad Sci U S A. 2010;107:16952–16957. doi: 10.1073/pnas.1002653107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Fang R, Barbera AJ, Xu Y, Rutenberg M, Leonor T, Bi Q, Lan F, Mei P, Yuan GC, Lian C, Peng J, Cheng D, Sui G, Kaiser UB, Shi Y, Shi YG. Human LSD2/KDM1b/AOF1 regulates gene transcription by modulating intragenic H3K4me2 methylation. Mol Cell. 2010;39:222–233. doi: 10.1016/j.molcel.2010.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Pryde F, Jain D, Kerr A, Curley R, Mariotti FR, Vogelauer M. H3 k36 methylation helps determine the timing of cdc45 association with replication origins. PLoS One. 2009;4:e5882. doi: 10.1371/journal.pone.0005882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Kim HS, Rhee DK, Jang YK. Methylations of histone H3 lysine 9 and lysine 36 are functionally linked to DNA replication checkpoint control in fission yeast. Biochem Biophys Res Commun. 2008;368:419–425. doi: 10.1016/j.bbrc.2008.01.104. [DOI] [PubMed] [Google Scholar]
  • 101.Li F, Mao G, Tong D, Huang J, Gu L, Yang W, Li GM. The histone mark H3K36me3 regulates human DNA mismatch repair through its interaction with MutSalpha. Cell. 2013;153:590–600. doi: 10.1016/j.cell.2013.03.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Akiki S, Dyer SA, Grimwade D, Ivey A, Abou-Zeid N, Borrow J, Jeffries S, Caddick J, Newell H, Begum S, Tawana K, Mason J, Velangi M, Griffiths M. NUP98-NSD1 fusion in association with FLT3-ITD mutation identifies a prognostically relevant subgroup of pediatric acute myeloid leukemia patients suitable for monitoring by real time quantitative PCR. Genes Chromosomes Cancer. 2013;52:1053–1064. doi: 10.1002/gcc.22100. [DOI] [PubMed] [Google Scholar]
  • 103.Tatton-Brown K, Rahman N. The NSD1 and EZH2 overgrowth genes, similarities and differences. Am J Med Genet C Semin Med Genet. 2013;163C:86–91. doi: 10.1002/ajmg.c.31359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Cross NC. Histone modification defects in developmental disorders and cancer. Oncotarget. 2012;3:3–4. doi: 10.18632/oncotarget.436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Keats JJ, Maxwell CA, Taylor BJ, Hendzel MJ, Chesi M, Bergsagel PL, Larratt LM, Mant MJ, Reiman T, Belch AR, Pilarski LM. Overexpression of transcripts originating from the MMSET locus characterizes all t(4;14) (p16;q32)-positive multiple myeloma patients. Blood. 2005;105:4060–4069. doi: 10.1182/blood-2004-09-3704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Douglas J, Coleman K, Tatton-Brown K, Hughes HE, Temple IK, Cole TR, Rahman N Childhood Overgrowth Collaboration. Evaluation of NSD2 and NSD3 in overgrowth syndromes. Eur J Hum Genet. 2005;13:150–153. doi: 10.1038/sj.ejhg.5201298. [DOI] [PubMed] [Google Scholar]
  • 107.Svedlund J, Koskinen Edblom S, Marquez VE, Akerstrom G, Bjorklund P, Westin G. Hypermethylated in cancer 1 (HIC1), a tumor suppressor gene epigenetically deregulated in hyperparathyroid tumors by histone H3 lysine modification. J Clin Endocrinol Metab. 2012;97:E1307–1315. doi: 10.1210/jc.2011-3136. [DOI] [PubMed] [Google Scholar]
  • 108.Deshpande AJ, Deshpande A, Sinha AU, Chen L, Chang J, Cihan A, Fazio M, Chen CW, Zhu N, Koche R, Dzhekieva L, Ibanez G, Dias S, Banka D, Krivtsov A, Luo M, Roeder RG, Bradner JE, Bernt KM, Armstrong SA. AF10 regulates progressive H3K79 methylation and HOX gene expression in diverse AML subtypes. Cancer Cell. 2014;26:896–908. doi: 10.1016/j.ccell.2014.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Drake KM, Watson VG, Kisielewski A, Glynn R, Napper AD. A sensitive luminescent assay for the histone methyltransferase NSD1 and other SAM-dependent enzymes. Assay Drug Dev Technol. 2014;12:258–271. doi: 10.1089/adt.2014.583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Shiba N, Ichikawa H, Taki T, Park MJ, Jo A, Mitani S, Kobayashi T, Shimada A, Sotomatsu M, Arakawa H, Adachi S, Tawa A, Horibe K, Tsuchida M, Hanada R, Tsukimoto I, Hayashi Y. NUP98-NSD1 gene fusion and its related gene expression signature are strongly associated with a poor prognosis in pediatric acute myeloid leukemia. Genes Chromosomes Cancer. 2013;52:683–693. doi: 10.1002/gcc.22064. [DOI] [PubMed] [Google Scholar]
  • 111.Migdalska AM, van der Weyden L, Ismail O, Sanger Mouse Genetics P, Rust AG, Rashid M, White JK, Sanchez-Andrade G, Lupski JR, Logan DW, Arends MJ, Adams DJ. Generation of the Sotos syndrome deletion in mice. Mamm Genome. 2012;23:749–757. doi: 10.1007/s00335-012-9416-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Jaffe JD, Wang Y, Chan HM, Zhang J, Huether R, Kryukov GV, Bhang HE, Taylor JE, Hu M, Englund NP, Yan F, Wang Z, Robert McDonald E 3rd, Wei L, Ma J, Easton J, Yu Z, deBeaumount R, Gibaja V, Venkatesan K, Schlegel R, Sellers WR, Keen N, Liu J, Caponigro G, Barretina J, Cooke VG, Mullighan C, Carr SA, Downing JR, Garraway LA, Stegmeier F. Global chromatin profiling reveals NSD2 mutations in pediatric acute lymphoblastic leukemia. Nat Genet. 2013;45:1386–1391. doi: 10.1038/ng.2777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Oyer JA, Huang X, Zheng Y, Shim J, Ezponda T, Carpenter Z, Allegretta M, Okot-Kotber CI, Patel JP, Melnick A, Levine RL, Ferrando A, Mackerell AD Jr, Kelleher NL, Licht JD, Popovic R. Point mutation E1099K in MMSET/NSD2 enhances its methyltranferase activity and leads to altered global chromatin methylation in lymphoid malignancies. Leukemia. 2014;28:198–201. doi: 10.1038/leu.2013.204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Angrand PO, Apiou F, Stewart AF, Dutrillaux B, Losson R, Chambon P. NSD3, a new SET domain-containing gene, maps to 8p12 and is amplified in human breast cancer cell lines. Genomics. 2001;74:79–88. doi: 10.1006/geno.2001.6524. [DOI] [PubMed] [Google Scholar]
  • 115.Rosati R, La Starza R, Veronese A, Aventin A, Schwienbacher C, Vallespi T, Negrini M, Martelli MF, Mecucci C. NUP98 is fused to the NSD3 gene in acute myeloid leukemia associated with t(8;11)(p11.2;p15) Blood. 2002;99:3857–3860. doi: 10.1182/blood.v99.10.3857. [DOI] [PubMed] [Google Scholar]
  • 116.Ostronoff F, Othus M, Gerbing RB, Loken MR, Raimondi SC, Hirsch BA, Lange BJ, Petersdorf S, Radich J, Appelbaum FR, Gamis AS, Alonzo TA, Meshinchi S. NUP98/NSD1 and FLT3/ITD coexpression is more prevalent in younger AML patients and leads to induction failure: a COG and SWOG report. Blood. 2014;124:2400–2407. doi: 10.1182/blood-2014-04-570929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Berdasco M, Ropero S, Setien F, Fraga MF, Lapunzina P, Losson R, Alaminos M, Cheung NK, Rahman N, Esteller M. Epigenetic inactivation of the Sotos overgrowth syndrome gene histone methyltransferase NSD1 in human neuroblastoma and glioma. Proc Natl Acad Sci U S A. 2009;106:21830–21835. doi: 10.1073/pnas.0906831106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Zhao Q, Caballero OL, Levy S, Stevenson BJ, Iseli C, de Souza SJ, Galante PA, Busam D, Leversha MA, Chadalavada K, Rogers YH, Venter JC, Simpson AJ, Strausberg RL. Transcriptome-guided characterization of genomic rearrangements in a breast cancer cell line. Proc Natl Acad Sci U S A. 2009;106:1886–1891. doi: 10.1073/pnas.0812945106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Yang P, Guo L, Duan ZJ, Tepper CG, Xue L, Chen X, Kung HJ, Gao AC, Zou JX, Chen HW. Histone methyltransferase NSD2/MMSET mediates constitutive NF-kappaB signaling for cancer cell proliferation, survival, and tumor growth via a feed-forward loop. Mol Cell Biol. 2012;32:3121–3131. doi: 10.1128/MCB.00204-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Sanidas I, Polytarchou C, Hatziapostolou M, Ezell SA, Kottakis F, Hu L, Guo A, Xie J, Comb MJ, Iliopoulos D, Tsichlis PN. Phosphoproteomics screen reveals akt isoformspecific signals linking RNA processing to lung cancer. Mol Cell. 2014;53:577–590. doi: 10.1016/j.molcel.2013.12.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Gossage L, Murtaza M, Slatter AF, Lichtenstein CP, Warren A, Haynes B, Marass F, Roberts I, Shanahan SJ, Claas A, Dunham A, May AP, Rosenfeld N, Forshew T, Eisen T. Clinical and pathological impact of VHL, PBRM1, BAP1, SETD2, KDM6A, and JARID1c in clear cell renal cell carcinoma. Genes Chromosomes Cancer. 2014;53:38–51. doi: 10.1002/gcc.22116. [DOI] [PubMed] [Google Scholar]
  • 122.Fontebasso AM, Schwartzentruber J, Khuong-Quang DA, Liu XY, Sturm D, Korshunov A, Jones DT, Witt H, Kool M, Albrecht S, Fleming A, Hadjadj D, Busche S, Lepage P, Montpetit A, Staffa A, Gerges N, Zakrzewska M, Zakrzewski K, Liberski PP, Hauser P, Garami M, Klekner A, Bognar L, Zadeh G, Faury D, Pfister SM, Jabado N, Majewski J. Mutations in SETD2 and genes affecting histone H3K36 methylation target hemispheric high-grade gliomas. Acta Neuropathol. 2013;125:659–669. doi: 10.1007/s00401-013-1095-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Patani N, Jiang WG, Newbold RF, Mokbel K. Histone-modifier gene expression profiles are associated with pathological and clinical outcomes in human breast cancer. Anticancer Res. 2011;31:4115–4125. [PubMed] [Google Scholar]
  • 124.Wang Y, Thomas A, Lau C, Rajan A, Zhu Y, Killian JK, Petrini I, Pham T, Morrow B, Zhong X, Meltzer PS, Giaccone G. Mutations of epigenetic regulatory genes are common in thymic carcinomas. Sci Rep. 2014;4:7336. doi: 10.1038/srep07336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Sankin A, Hakimi AA, Mikkilineni N, Ostrovnaya I, Silk MT, Liang Y, Mano R, Chevinsky M, Motzer RJ, Solomon SB, Cheng EH, Durack JC, Coleman JA, Russo P, Hsieh JJ. The impact of genetic heterogeneity on biomarker development in kidney cancer assessed by multiregional sampling. Cancer Med. 2014;3:1485–1492. doi: 10.1002/cam4.293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Kudithipudi S, Jeltsch A. Role of somatic cancer mutations in human protein lysine methyltransferases. Biochim Biophys Acta. 2014;1846:366–379. doi: 10.1016/j.bbcan.2014.08.002. [DOI] [PubMed] [Google Scholar]
  • 127.Hao C, Wang L, Peng S, Cao M, Li H, Hu J, Huang X, Liu W, Zhang H, Wu S, Pataer A, Heymach JV, Eterovic AK, Zhang Q, Shaw KR, Chen K, Futreal A, Wang M, Hofstetter W, Mehran R, Rice D, Roth JA, Sepesi B, Swisher SG, Vaporciyan A, Walsh GL, Johnson FM, Fang B. Gene mutations in primary tumors and corresponding patient-derived xenografts derived from non-small cell lung cancer. Cancer Lett. 2015;357:179–185. doi: 10.1016/j.canlet.2014.11.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Mar BG, Bullinger LB, McLean KM, Grauman PV, Harris MH, Stevenson K, Neuberg DS, Sinha AU, Sallan SE, Silverman LB, Kung AL, Lo Nigro L, Ebert BL, Armstrong SA. Mutations in epigenetic regulators including SETD2 are gained during relapse in paediatric acute lymphoblastic leukaemia. Nat Commun. 2014;5:3469. doi: 10.1038/ncomms4469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Zhang K, Haversat JM, Mager J. CTR9/PAF1c regulates molecular lineage identity, histone H3K36 trimethylation and genomic imprinting during preimplantation development. Dev Biol. 2013;383:15–27. doi: 10.1016/j.ydbio.2013.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Hu M, Sun XJ, Zhang YL, Kuang Y, Hu CQ, Wu WL, Shen SH, Du TT, Li H, He F, Xiao HS, Wang ZG, Liu TX, Lu H, Huang QH, Chen SJ, Chen Z. Histone H3 lysine 36 methyltransferase Hypb/Setd2 is required for embryonic vascular remodeling. Proc Natl Acad Sci U S A. 2010;107:2956–2961. doi: 10.1073/pnas.0915033107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Martinez HR, Belmont JW, Craigen WJ, Taylor MD, Jefferies JL. Left ventricular noncompaction in Sotos syndrome. Am J Med Genet A. 2011;155A:1115–1118. doi: 10.1002/ajmg.a.33838. [DOI] [PubMed] [Google Scholar]
  • 132.Nimura K, Ura K, Shiratori H, Ikawa M, Okabe M, Schwartz RJ, Kaneda Y. A histone H3 lysine 36 trimethyltransferase links Nkx2-5 to Wolf-Hirschhorn syndrome. Nature. 2009;460:287–291. doi: 10.1038/nature08086. [DOI] [PubMed] [Google Scholar]

Articles from American Journal of Translational Research are provided here courtesy of e-Century Publishing Corporation

RESOURCES