KMT2D

Last updated
KMT2D
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases KMT2D , ALR, KABUK1, MLL2, MLL4, lysine methyltransferase 2D, histone-lysine methyltransferase 2D, TNRC21, AAD10, KMS, CAGL114
External IDs OMIM: 602113 MGI: 2682319 HomoloGene: 86893 GeneCards: KMT2D
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_003482

NM_001033276
NM_001033388

RefSeq (protein)

NP_003473

NP_001028448

Location (UCSC) Chr 12: 49.02 – 49.06 Mb Chr 15: 98.73 – 98.77 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

Histone-lysine N-methyltransferase 2D (KMT2D), also known as MLL4 and sometimes MLL2 in humans and Mll4 in mice, is a major mammalian histone H3 lysine 4 (H3K4) mono-methyltransferase. [5] It is part of a family of six Set1-like H3K4 methyltransferases that also contains KMT2A (or MLL1), KMT2B (or MLL2), KMT2C (or MLL3), KMT2F (or SET1A), and KMT2G (or SET1B).

Contents

KMT2D is a large protein over 5,500 amino acids in size and is widely expressed in adult tissues. [6] The protein co-localizes with lineage determining transcription factors on transcriptional enhancers and is essential for cell differentiation and embryonic development. [5] It also plays critical roles in regulating cell fate transition, [5] [7] [8] [9] metabolism, [10] [11] and tumor suppression. [12] [13] [14] [15]

Mutations in KMT2D cause human genetic conditions including Kabuki syndrome, [16] another distinct congenital malformations disorder [17] and various forms of cancer. [18]

Structure

Gene

In mice, KMT2D is coded by the Kmt2d gene located on chromosome 15F1. Its transcript is 19,823 base pairs long and contains 55 exons and 54 introns. [19]

In humans, KMT2D is coded by the KMT2D gene located on chromosome 12q13.12. It's transcript is 19,419 base pairs long and contains 54 exons and 53 introns. [20]

Protein

KMT2D is homologous to Trithorax-related (Trr), which is a Trithorax-group protein. [21] The mouse and human KMT2D proteins are 5,588 and 5,537 amino acids in length, respectively. Both species of the protein weigh about 600 kDa. [19] [20] KMT2D contains an enzymatically active C-terminal SET domain that is responsible for its methyltransferase activity and maintaining protein stability in cells. [22] Near the SET domain are a plant homeotic domain (PHD) and FY-rich N/C-terminal (FYRN and FYRC) domains. The protein also contains six N-terminal PHDs, a high mobility group (HMG-I), and nine nuclear receptor interacting motifs (LXXLLs). [18] It was shown that amino acids Y5426 and Y5512 are critical for the enzymatic activity of human KMT2D in vitro. [23] In addition, mutation of Y5477 in mouse KMT2D, which corresponds to Y5426 in human KMT2D, resulted in the inactivation of KMT2D's enzymatic activity in embryonic stem cells. [24] Depletion of cellular H3K4 methylation reduces KMT2D levels, indicating that the protein's stability could be regulated by cellular H3K4 methylation. [23]

Protein complex

Several components of the KMT2D complex were first purified in 2003, [25] and then the entire complex was identified in 2007. [26] [27] [28] [29] Along with KMT2D, the complex also contains ASH2L, RbBP5, WDR5, DPY30, NCOA6, UTX (also known as KDM6A), PA1, and PTIP. WDR5, RbBP5, ASH2L, and DPY30 form the four-subunit sub-complex WRAD, which is critical for H3K4 methyltransferase activity in all mammalian Set1-like histone methyltransferase complexes. [30] WDR5 binds directly with FYRN/FYRC domains of C-terminal SET domain-containing fragments of human KMT2C and KMT2D. [26] UTX, the complex’s H3K27 demethylase, PTIP, and PA1 are subunits that are unique to KMT2C and KMT2D. [26] [31] [32] KMT2D acts as a scaffold protein within the complex; absence of KMT2D results in destabilization of UTX and collapse of the complex in cells. [5] [23]

Enhancer regulation

KMT2D is a major enhancer mono-methyltransferase and has partial functional redundancy with KMT2C. [5] [7] The protein selectively binds enhancer regions based on type of cell and stage of differentiation. During differentiation, lineage determining transcription factors recruit KMT2D to establish cell-type specific enhancers. For example, CCAAT/enhancer-binding protein β (C/EBPβ), an early adipogenic transcription factor, recruits and requires KMT2D to establish a subset of adipogenic enhancers during adipogenesis. Depletion of KMT2D prior to differentiation prevents the accumulation of H3K4 mono-methylation (H3K4me1), H3K27 acetylation, the transcriptional coactivator Mediator, and RNA polymerase II on enhancers, resulting in severe defects in gene expression and cell differentiation. [5] KMT2C and KMT2D also identify super-enhancers and are required for formation of super-enhancers during cell differentiation. [33] Mechanistically, KMT2C and KMT2D are required for the binding of H3K27 acetyltransferases CREB-binding protein (CBP) and/or p300 on enhancers, enhancer activation, and enhancer-promotor looping prior to gene transcription. [5] [33] The KMT2C and KMT2D proteins, rather than the KMT2C and KMT2D-mediated H3K4me1, control p300 recruitment to enhancers, enhancer activation, and transcription from promoters in embryonic stem cells. [7]

Functions

Development

Whole-body knockout of Kmt2d in mice results in early embryonic lethality. [5] Targeted knockout of Kmt2d in precursors cells of brown adipocytes and myocytes results in decreases in brown adipose tissue and muscle mass in mice, indicating that KMT2D is required for adipose and muscle tissue development. [5] In the hearts of mice, a single copy of the Kmt2d gene is sufficient for normal heart development. [34] Complete loss of Kmt2d in cardiac precursors and myocardium leads to severe cardiac defects and early embryonic lethality. KMT2D mediated mono- and di-methylation is required for maintaining necessary gene expression programs during heart development. Knockout studies in mice also show that KMT2D is required for proper B-cell development. [12]

Cell fate transition

KMT2D is partially functionally redundant with KMT2C and is required for cell differentiation in culture. [5] [7] KMT2D regulates the induction of adipogenic and myogenic genes and is required for cell-type specific gene expression during differentiation. KMT2C and KMT2D are essential for adipogenesis and myogenesis. [5] Similar functions are seen in neuronal and osteoblast differentiation. [8] [9] KMT2D facilitates cell fate transition by priming enhancers (through H3K4me1) for p300-mediated activation. For p300 to bind the enhancer, the physical presence of KMT2D, and not just the KMT2D-mediated H3K4me1, is required. However, KMT2D is dispensable for maintaining embryonic stem cell and somatic cell identity. [7]

Metabolism

KMT2D is partially functionally redundant with KMT2C in the liver as well. Heterozygous Kmt2d+/- mice exhibit enhanced glucose tolerance and insulin sensitivity and increased serum bile acid. [10] KMT2C and KMT2D are significant epigenetic regulators of the hepatic circadian clock and are co-activators of the circadian transcription factors retinoid-related orphan receptor (ROR)-α and -γ. [10] In mice, KMT2D also acts as a coactivator of PPARγ within the liver to direct over-nutrition induced steatosis. Heterozygous Kmt2d+/- mice exhibit resistance to over-nutrition induced hepatic steatosis. [11]

Tumor suppression

KMT2C and KMT2D along with NCOA6 act as coactivators of p53, a well-established tumor suppressor and transcription factor, and are necessary for endogenous expression of p53 in response to doxorubicin, a DNA damaging agent. [13] KMT2C and KMT2D have also been implicated with tumor suppressor roles in acute myeloid leukemia, follicular lymphoma, and diffuse large B cell lymphoma. [12] [14] [15] Knockout of Kmt2d in mice negatively affects the expression of tumor suppressor genes TNFAIP3, SOCS3, and TNFRSF14. [15]

Conversely, KMT2D deficiency in several breast and colon cancer cell lines leads to reduced proliferation. [35] [36] [37] Increased KMT2D was shown to facilitate chromatin opening and recruitment of transcription factors, including estrogen receptor (ER), in ER-positive breast cancer cells. [38] Thus, KMT2D may have diverse effects on tumor suppression in different cell types.

Clinical significance

Germline heterozygous loss of function mutations in KMT2D, also known as MLL2 in humans, cause Kabuki syndrome type 1 [16] with mutational occurrence rates between 56% and 75%. [39] [40] [41] [42] Mosaic mutations and intragenic deletions and duplications have also been described in this condition. [43] Type 1 Kabuki syndrome is characterised by developmental delay, intellectual disability, postnatal dwarfism, recognizable facial dysmorphism (reminiscent of the make-up of actors of Kabuki theatre), a broad and depressed nasal tip, large prominent earlobes, a cleft or high-arched palate, scoliosis, short fifth finger, persistence of fingerpads, radiographic abnormalities of the vertebrae, hands, and hip joints, and recurrent otitis media in infancy. [44] Note that variants in a functionally related gene, KDM6A, cause Kabuki syndrome type 2 that is an X-linked condition, shares several clinical features with Kabuki syndrome type 1 but phenotypically is a much more variable condition. [45]

Germline heterozygous missense variants in exon 38 or 39 of the KMT2D gene cause another rare distinct multiple malformation disorder characterized by choanal atresia, athelia or hypoplastic nipples, branchial sinus abnormalities, neck pits, lacrimal duct anomalies, hearing loss, external ear malformations, and thyroid abnormalities. [17]

Congenital heart disease has been associated with an excess of mutations in genes that regulate H3K4 methylation, including KMT2D. [46]

Somatic frameshift and nonsense mutations in the SET and PHD domains affect 37% and 60%, respectively, of the total KMT2D mutations in cancers. [18] Cancers with somatic mutations in KMT2D occur most commonly in the brain, lymph nodes, blood, lungs, large intestine, and endometrium. [18] These cancers include medulloblastoma, [47] [48] [49] pheochromocytoma, [50] non-Hodgkin lymphomas, [51] cutaneous T-cell lymphoma, Sézary syndrome, [52] bladder, lung, and endometrial carcinomas, [53] esophageal squamous cell carcinoma, [54] [55] [56] pancreatic cancer, [57] and prostate cancer. [58]

Note

Related Research Articles

<span class="mw-page-title-main">Histone methyltransferase</span> Histone-modifying enzymes

Histone methyltransferases (HMT) are histone-modifying enzymes, that catalyze the transfer of one, two, or three methyl groups to lysine and arginine residues of histone proteins. The attachment of methyl groups occurs predominantly at specific lysine or arginine residues on histones H3 and H4. Two major types of histone methyltranferases exist, lysine-specific and arginine-specific. In both types of histone methyltransferases, S-Adenosyl methionine (SAM) serves as a cofactor and methyl donor group.
The genomic DNA of eukaryotes associates with histones to form chromatin. The level of chromatin compaction depends heavily on histone methylation and other post-translational modifications of histones. Histone methylation is a principal epigenetic modification of chromatin that determines gene expression, genomic stability, stem cell maturation, cell lineage development, genetic imprinting, DNA methylation, and cell mitosis.

Histone methylation is a process by which methyl groups are transferred to amino acids of histone proteins that make up nucleosomes, which the DNA double helix wraps around to form chromosomes. Methylation of histones can either increase or decrease transcription of genes, depending on which amino acids in the histones are methylated, and how many methyl groups are attached. Methylation events that weaken chemical attractions between histone tails and DNA increase transcription because they enable the DNA to uncoil from nucleosomes so that transcription factor proteins and RNA polymerase can access the DNA. This process is critical for the regulation of gene expression that allows different cells to express different genes.

The histone code is a hypothesis that the transcription of genetic information encoded in DNA is in part regulated by chemical modifications to histone proteins, primarily on their unstructured ends. Together with similar modifications such as DNA methylation it is part of the epigenetic code. Histones associate with DNA to form nucleosomes, which themselves bundle to form chromatin fibers, which in turn make up the more familiar chromosome. Histones are globular proteins with a flexible N-terminus that protrudes from the nucleosome. Many of the histone tail modifications correlate very well to chromatin structure and both histone modification state and chromatin structure correlate well to gene expression levels. The critical concept of the histone code hypothesis is that the histone modifications serve to recruit other proteins by specific recognition of the modified histone via protein domains specialized for such purposes, rather than through simply stabilizing or destabilizing the interaction between histone and the underlying DNA. These recruited proteins then act to alter chromatin structure actively or to promote transcription. For details of gene expression regulation by histone modifications see table below.

Demethylases are enzymes that remove methyl (CH3) groups from nucleic acids, proteins (particularly histones), and other molecules. Demethylases are important epigenetic proteins, as they are responsible for transcriptional regulation of the genome by controlling the methylation of DNA and histones, and by extension, the chromatin state at specific gene loci.

<span class="mw-page-title-main">KDM1A</span> Protein-coding gene in the species Homo sapiens

Lysine-specific histone demethylase 1A (LSD1) also known as lysine (K)-specific demethylase 1A (KDM1A) is a protein that in humans is encoded by the KDM1A gene. LSD1 is a flavin-dependent monoamine oxidase, which can demethylate mono- and di-methylated lysines, specifically histone 3, lysine 4 (H3K4). Other reported methylated lysine substrates such as histone H3K9 and TP53 have not been biochemically validated. This enzyme plays a critical role in oocyte growth, embryogenesis, hematopoiesis and tissue-specific differentiation. LSD1 was the first histone demethylase to be discovered though more than 30 have since been described.

<span class="mw-page-title-main">DNMT3B</span> Protein-coding gene in the species Homo sapiens

DNA (cytosine-5)-methyltransferase 3 beta, is an enzyme that in humans in encoded by the DNMT3B gene. Mutation in this gene are associated with immunodeficiency, centromere instability and facial anomalies syndrome.

<span class="mw-page-title-main">EZH2</span> Protein-coding gene in the species Homo sapiens

Enhancer of zeste homolog 2 (EZH2) is a histone-lysine N-methyltransferase enzyme encoded by EZH2 gene, that participates in histone methylation and, ultimately, transcriptional repression. EZH2 catalyzes the addition of methyl groups to histone H3 at lysine 27, by using the cofactor S-adenosyl-L-methionine. Methylation activity of EZH2 facilitates heterochromatin formation thereby silences gene function. Remodeling of chromosomal heterochromatin by EZH2 is also required during cell mitosis.

<span class="mw-page-title-main">H3F3A</span> Gene for histone H3.3 protein

Histone H3.3 is a protein that in humans is encoded by the H3F3A and H3F3B genes. It plays an essential role in maintaining genome integrity during mammalian development.

<span class="mw-page-title-main">DNA (cytosine-5)-methyltransferase 3A</span> Protein-coding gene in the species Homo sapiens

DNA (cytosine-5)-methyltransferase 3A (DNMT3A) is an enzyme that catalyzes the transfer of methyl groups to specific CpG structures in DNA, a process called DNA methylation. The enzyme is encoded in humans by the DNMT3A gene.

<span class="mw-page-title-main">KMT2A</span> Protein-coding gene in the species Homo sapiens

Histone-lysine N-methyltransferase 2A, also known as acute lymphoblastic leukemia 1 (ALL-1), myeloid/lymphoid or mixed-lineage leukemia1 (MLL1), or zinc finger protein HRX (HRX), is an enzyme that in humans is encoded by the KMT2A gene.

<span class="mw-page-title-main">SETDB1</span> Enzyme-coding gene in humans

Histone-lysine N-methyltransferase SETDB1 is an enzyme that in humans is encoded by the SETDB1 gene. SETDB1 is also known as KMT1E or H3K9 methyltransferase ESET.

<span class="mw-page-title-main">KMT2C</span> Protein-coding gene in the species Homo sapiens

Lysine N-methyltransferase 2C (KMT2C) also known as myeloid/lymphoid or mixed-lineage leukemia protein 3 (MLL3) is an enzyme that in humans is encoded by the KMT2C gene.

<span class="mw-page-title-main">SETMAR</span> Protein-coding gene in the species Homo sapiens

Histone-lysine N-methyltransferase SETMAR is an enzyme that in humans is encoded by the SETMAR gene.

<span class="mw-page-title-main">SMYD3</span> Protein-coding gene in the species Homo sapiens

SET and MYND (myeloid-Nervy-DEAF-1) domain-containing protein 3 is a protein that in humans is encoded by the SMYD3 gene.

<span class="mw-page-title-main">Cancer epigenetics</span> Field of study in cancer research

Cancer epigenetics is the study of epigenetic modifications to the DNA of cancer cells that do not involve a change in the nucleotide sequence, but instead involve a change in the way the genetic code is expressed. Epigenetic mechanisms are necessary to maintain normal sequences of tissue specific gene expression and are crucial for normal development. They may be just as important, if not even more important, than genetic mutations in a cell's transformation to cancer. The disturbance of epigenetic processes in cancers, can lead to a loss of expression of genes that occurs about 10 times more frequently by transcription silencing than by mutations. As Vogelstein et al. points out, in a colorectal cancer there are usually about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations. However, in colon tumors compared to adjacent normal-appearing colonic mucosa, there are about 600 to 800 heavily methylated CpG islands in the promoters of genes in the tumors while these CpG islands are not methylated in the adjacent mucosa. Manipulation of epigenetic alterations holds great promise for cancer prevention, detection, and therapy. In different types of cancer, a variety of epigenetic mechanisms can be perturbed, such as the silencing of tumor suppressor genes and activation of oncogenes by altered CpG island methylation patterns, histone modifications, and dysregulation of DNA binding proteins. There are several medications which have epigenetic impact, that are now used in a number of these diseases.

H3K4me3 is an epigenetic modification to the DNA packaging protein Histone H3 that indicates tri-methylation at the 4th lysine residue of the histone H3 protein and is often involved in the regulation of gene expression. The name denotes the addition of three methyl groups (trimethylation) to the lysine 4 on the histone H3 protein.

H3K4me1 is an epigenetic modification to the DNA packaging protein Histone H3. It is a mark that indicates the mono-methylation at the 4th lysine residue of the histone H3 protein and often associated with gene enhancers.

H3K36me3 is an epigenetic modification to the DNA packaging protein Histone H3. It is a mark that indicates the tri-methylation at the 36th lysine residue of the histone H3 protein and often associated with gene bodies.

H4K20me is an epigenetic modification to the DNA packaging protein Histone H4. It is a mark that indicates the mono-methylation at the 20th lysine residue of the histone H4 protein. This mark can be di- and tri-methylated. It is critical for genome integrity including DNA damage repair, DNA replication and chromatin compaction.

H3K36me is an epigenetic modification to the DNA packaging protein Histone H3, specifically, the mono-methylation at the 36th lysine residue of the histone H3 protein.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000167548 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000048154 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. 1 2 3 4 5 6 7 8 9 10 11 Lee JE, Wang C, Xu S, Cho YW, Wang L, Feng X, et al. (December 2013). "H3K4 mono- and di-methyltransferase MLL4 is required for enhancer activation during cell differentiation". eLife. 2: e01503. arXiv: 1311.7328 . doi: 10.7554/eLife.01503 . PMC   3869375 . PMID   24368734.
  6. Prasad R, Zhadanov AB, Sedkov Y, Bullrich F, Druck T, Rallapalli R, et al. (July 1997). "Structure and expression pattern of human ALR, a novel gene with strong homology to ALL-1 involved in acute leukemia and to Drosophila trithorax". Oncogene. 15 (5): 549–60. doi:10.1038/sj.onc.1201211. PMID   9247308. S2CID   31178216.
  7. 1 2 3 4 5 Wang C, Lee JE, Lai B, Macfarlan TS, Xu S, Zhuang L, Liu C, Peng W, Ge K (October 2016). "Enhancer priming by H3K4 methyltransferase MLL4 controls cell fate transition". Proceedings of the National Academy of Sciences of the United States of America. 113 (42): 11871–11876. Bibcode:2016PNAS..11311871W. doi: 10.1073/pnas.1606857113 . PMC   5081576 . PMID   27698142.
  8. 1 2 Dhar SS, Lee SH, Kan PY, Voigt P, Ma L, Shi X, Reinberg D, Lee MG (December 2012). "Trans-tail regulation of MLL4-catalyzed H3K4 methylation by H4R3 symmetric dimethylation is mediated by a tandem PHD of MLL4". Genes & Development. 26 (24): 2749–62. doi:10.1101/gad.203356.112. PMC   3533079 . PMID   23249737.
  9. 1 2 Munehira Y, Yang Z, Gozani O (October 2016). "Systematic Analysis of Known and Candidate Lysine Demethylases in the Regulation of Myoblast Differentiation". Journal of Molecular Biology. 429 (13): 2055–2065. doi:10.1016/j.jmb.2016.10.004. PMC   5388604 . PMID   27732873.
  10. 1 2 3 Kim DH, Rhee JC, Yeo S, Shen R, Lee SK, Lee JW, Lee S (March 2015). "Crucial roles of mixed-lineage leukemia 3 and 4 as epigenetic switches of the hepatic circadian clock controlling bile acid homeostasis in mice". Hepatology. 61 (3): 1012–23. doi:10.1002/hep.27578. PMC   4474368 . PMID   25346535.
  11. 1 2 Kim DH, Kim J, Kwon JS, Sandhu J, Tontonoz P, Lee SK, Lee S, Lee JW (November 2016). "Critical Roles of the Histone Methyltransferase MLL4/KMT2D in Murine Hepatic Steatosis Directed by ABL1 and PPARγ2". Cell Reports. 17 (6): 1671–1682. doi: 10.1016/j.celrep.2016.10.023 . PMID   27806304.
  12. 1 2 3 Zhang J, Dominguez-Sola D, Hussein S, Lee JE, Holmes AB, Bansal M, et al. (October 2015). "Disruption of KMT2D perturbs germinal center B cell development and promotes lymphomagenesis". Nature Medicine. 21 (10): 1190–8. doi:10.1038/nm.3940. PMC   5145002 . PMID   26366712.
  13. 1 2 Lee J, Kim DH, Lee S, Yang QH, Lee DK, Lee SK, Roeder RG, Lee JW (May 2009). "A tumor suppressive coactivator complex of p53 containing ASC-2 and histone H3-lysine-4 methyltransferase MLL3 or its paralogue MLL4". Proceedings of the National Academy of Sciences of the United States of America. 106 (21): 8513–8. Bibcode:2009PNAS..106.8513L. doi: 10.1073/pnas.0902873106 . PMC   2689008 . PMID   19433796.
  14. 1 2 Chen C, Liu Y, Rappaport AR, Kitzing T, Schultz N, Zhao Z, Shroff AS, Dickins RA, Vakoc CR, Bradner JE, Stock W, LeBeau MM, Shannon KM, Kogan S, Zuber J, Lowe SW (May 2014). "MLL3 is a haploinsufficient 7q tumor suppressor in acute myeloid leukemia". Cancer Cell. 25 (5): 652–65. doi:10.1016/j.ccr.2014.03.016. PMC   4206212 . PMID   24794707.
  15. 1 2 3 Ortega-Molina A, Boss IW, Canela A, Pan H, Jiang Y, Zhao C, et al. (October 2015). "The histone lysine methyltransferase KMT2D sustains a gene expression program that represses B cell lymphoma development". Nature Medicine. 21 (10): 1199–208. doi:10.1038/nm.3943. PMC   4676270 . PMID   26366710.
  16. 1 2 Ng SB, Bigham AW, Buckingham KJ, Hannibal MC, McMillin MJ, Gildersleeve HI, et al. (September 2010). "Exome sequencing identifies MLL2 mutations as a cause of Kabuki syndrome". Nature Genetics. 42 (9): 790–793. doi:10.1038/ng.646. PMC   2930028 . PMID   20711175.
  17. 1 2 Cuvertino S, Hartill V, Colyer A, Garner T, Nair N, Al-Gazali L, et al. (May 2020). "A restricted spectrum of missense KMT2D variants cause a multiple malformations disorder distinct from Kabuki syndrome". Genetics in Medicine. 22 (5): 867–877. doi:10.1038/s41436-019-0743-3. PMC   7200597 . PMID   31949313.
  18. 1 2 3 4 Rao RC, Dou Y (June 2015). "Hijacked in cancer: the KMT2 (MLL) family of methyltransferases". Nature Reviews. Cancer. 15 (6): 334–346. doi:10.1038/nrc3929. PMC   4493861 . PMID   25998713.
  19. 1 2 "Transcript: Kmt2d-001 (ENSMUST00000023741.15) - Summary - Mus musculus - Ensembl genome browser 88". www.ensembl.org.
  20. 1 2 "Transcript: KMT2D-001 (ENST00000301067.11) - Summary - Homo sapiens - Ensembl genome browser 88". www.ensembl.org.
  21. Mohan M, Herz HM, Smith ER, Zhang Y, Jackson J, Washburn MP, Florens L, Eissenberg JC, Shilatifard A (November 2011). "The COMPASS family of H3K4 methylases in Drosophila". Molecular and Cellular Biology. 31 (21): 4310–8. doi:10.1128/MCB.06092-11. PMC   3209330 . PMID   21875999.
  22. Ruthenburg AJ, Allis CD, Wysocka J (January 2007). "Methylation of lysine 4 on histone H3: intricacy of writing and reading a single epigenetic mark". Molecular Cell. 25 (1): 15–30. doi: 10.1016/j.molcel.2006.12.014 . PMID   17218268.
  23. 1 2 3 Jang Y, Wang C, Zhuang L, Liu C, Ge K (December 2016). "H3K4 Methyltransferase Activity Is Required for MLL4 Protein Stability". Journal of Molecular Biology. 429 (13): 2046–2054. doi:10.1016/j.jmb.2016.12.016. PMC   5474351 . PMID   28013028.
  24. Dorighi KM, Swigut T, Henriques T, Bhanu NV, Scruggs BS, Nady N, Still CD, Garcia BA, Adelman K, Wysocka J (May 2017). "Mll3 and Mll4 Facilitate Enhancer RNA Synthesis and Transcription from Promoters Independently of H3K4 Monomethylation". Molecular Cell. 66 (4): 568–576.e4. doi:10.1016/j.molcel.2017.04.018. PMC   5662137 . PMID   28483418.
  25. Goo YH, Sohn YC, Kim DH, Kim SW, Kang MJ, Jung DJ, Kwak E, Barlev NA, Berger SL, Chow VT, Roeder RG, Azorsa DO, Meltzer PS, Suh PG, Song EJ, Lee KJ, Lee YC, Lee JW (January 2003). "Activating signal cointegrator 2 belongs to a novel steady-state complex that contains a subset of trithorax group proteins". Molecular and Cellular Biology. 23 (1): 140–9. doi:10.1128/mcb.23.1.140-149.2003. PMC   140670 . PMID   12482968.
  26. 1 2 3 Cho YW, Hong T, Hong S, Guo H, Yu H, Kim D, Guszczynski T, Dressler GR, Copeland TD, Kalkum M, Ge K (July 2007). "PTIP associates with MLL3- and MLL4-containing histone H3 lysine 4 methyltransferase complex". The Journal of Biological Chemistry. 282 (28): 20395–406. doi: 10.1074/jbc.M701574200 . PMC   2729684 . PMID   17500065.
  27. Issaeva I, Zonis Y, Rozovskaia T, Orlovsky K, Croce CM, Nakamura T, Mazo A, Eisenbach L, Canaani E (March 2007). "Knockdown of ALR (MLL2) reveals ALR target genes and leads to alterations in cell adhesion and growth". Molecular and Cellular Biology. 27 (5): 1889–903. doi:10.1128/MCB.01506-06. PMC   1820476 . PMID   17178841.
  28. Lee MG, Villa R, Trojer P, Norman J, Yan KP, Reinberg D, Di Croce L, Shiekhattar R (October 2007). "Demethylation of H3K27 regulates polycomb recruitment and H2A ubiquitination". Science. 318 (5849): 447–50. Bibcode:2007Sci...318..447L. doi: 10.1126/science.1149042 . PMID   17761849. S2CID   23883131.
  29. Patel A, Vought VE, Dharmarajan V, Cosgrove MS (November 2008). "A conserved arginine-containing motif crucial for the assembly and enzymatic activity of the mixed lineage leukemia protein-1 core complex". The Journal of Biological Chemistry. 283 (47): 32162–75. doi: 10.1074/jbc.M806317200 . PMID   18829457.
  30. Ernst P, Vakoc CR (May 2012). "WRAD: enabler of the SET1-family of H3K4 methyltransferases". Briefings in Functional Genomics. 11 (3): 217–26. doi:10.1093/bfgp/els017. PMC   3388306 . PMID   22652693.
  31. Cho YW, Hong S, Ge K (2012). "Affi nity Purifi cation of MLL3/MLL4 Histone H3K4 Methyltransferase Complex". Transcriptional Regulation. Methods in Molecular Biology. Vol. 809. pp. 465–72. doi:10.1007/978-1-61779-376-9_30. ISBN   978-1-61779-375-2. PMC   3467094 . PMID   22113294.
  32. Hong S, Cho YW, Yu LR, Yu H, Veenstra TD, Ge K (November 2007). "Identification of JmjC domain-containing UTX and JMJD3 as histone H3 lysine 27 demethylases". Proceedings of the National Academy of Sciences of the United States of America. 104 (47): 18439–44. Bibcode:2007PNAS..10418439H. doi: 10.1073/pnas.0707292104 . PMC   2141795 . PMID   18003914.
  33. 1 2 Lai B, Lee JE, Jang Y, Wang L, Peng W, Ge K (April 2017). "MLL3/MLL4 are required for CBP/p300 binding on enhancers and super-enhancer formation in brown adipogenesis". Nucleic Acids Research. 45 (11): 6388–6403. doi:10.1093/nar/gkx234. PMC   5499743 . PMID   28398509.
  34. Ang SY, Uebersohn A, Spencer CI, Huang Y, Lee JE, Ge K, Bruneau BG (March 2016). "KMT2D regulates specific programs in heart development via histone H3 lysine 4 di-methylation". Development. 143 (5): 810–21. doi:10.1242/dev.132688. PMC   4813342 . PMID   26932671.
  35. Guo C, Chen LH, Huang Y, Chang CC, Wang P, Pirozzi CJ, et al. (November 2013). "KMT2D maintains neoplastic cell proliferation and global histone H3 lysine 4 monomethylation". Oncotarget. 4 (11): 2144–53. doi:10.18632/oncotarget.1555. PMC   3875776 . PMID   24240169.
  36. Kim JH, Sharma A, Dhar SS, Lee SH, Gu B, Chan CH, Lin HK, Lee MG (March 2014). "UTX and MLL4 coordinately regulate transcriptional programs for cell proliferation and invasiveness in breast cancer cells". Cancer Research. 74 (6): 1705–17. doi:10.1158/0008-5472.CAN-13-1896. PMC   3962500 . PMID   24491801.
  37. Mo R, Rao SM, Zhu YJ (June 2006). "Identification of the MLL2 complex as a coactivator for estrogen receptor alpha". The Journal of Biological Chemistry. 281 (23): 15714–20. doi: 10.1074/jbc.M513245200 . PMID   16603732.
  38. Toska E, Osmanbeyoglu HU, Castel P, Chan C, Hendrickson RC, Elkabets M, Dickler MN, Scaltriti M, Leslie CS, Armstrong SA, Baselga J (March 2017). "PI3K pathway regulates ER-dependent transcription in breast cancer through the epigenetic regulator KMT2D". Science. 355 (6331): 1324–1330. Bibcode:2017Sci...355.1324T. doi:10.1126/science.aah6893. PMC   5485411 . PMID   28336670.
  39. Banka S, Veeramachaneni R, Reardon W, Howard E, Bunstone S, Ragge N, et al. (April 2012). "How genetically heterogeneous is Kabuki syndrome?: MLL2 testing in 116 patients, review and analyses of mutation and phenotypic spectrum". European Journal of Human Genetics. 20 (4): 381–388. doi:10.1038/ejhg.2011.220. PMC   3306863 . PMID   22126750.
  40. Bögershausen N, Wollnik B (March 2013). "Unmasking Kabuki syndrome". Clinical Genetics. 83 (3): 201–211. doi:10.1111/cge.12051. PMID   23131014. S2CID   204999137.
  41. Li Y, Bögershausen N, Alanay Y, Simsek Kiper PO, Plume N, Keupp K, et al. (December 2011). "A mutation screen in patients with Kabuki syndrome". Human Genetics. 130 (6): 715–724. doi:10.1007/s00439-011-1004-y. PMID   21607748. S2CID   12327505.
  42. Paulussen AD, Stegmann AP, Blok MJ, Tserpelis D, Posma-Velter C, Detisch Y, et al. (February 2011). "MLL2 mutation spectrum in 45 patients with Kabuki syndrome". Human Mutation. 32 (2): E2018–E2025. doi: 10.1002/humu.21416 . PMID   21280141. S2CID   7380692.
  43. Banka S, Howard E, Bunstone S, Chandler KE, Kerr B, Lachlan K, et al. (May 2013). "MLL2 mosaic mutations and intragenic deletion-duplications in patients with Kabuki syndrome". Clinical Genetics. 83 (5): 467–471. doi:10.1111/j.1399-0004.2012.01955.x. PMID   22901312. S2CID   8316503.
  44. "OMIM Entry - # 147920 - KABUKI SYNDROME 1; KABUK1". omim.org. Retrieved 2022-01-03.
  45. Faundes V, Goh S, Akilapa R, Bezuidenhout H, Bjornsson HT, Bradley L, et al. (July 2021). "Clinical delineation, sex differences, and genotype-phenotype correlation in pathogenic KDM6A variants causing X-linked Kabuki syndrome type 2". Genetics in Medicine. 23 (7): 1202–1210. doi:10.1038/s41436-021-01119-8. PMC   8257478 . PMID   33674768.
  46. Zaidi S, Choi M, Wakimoto H, Ma L, Jiang J, Overton JD, et al. (June 2013). "De novo mutations in histone-modifying genes in congenital heart disease" (PDF). Nature. 498 (7453): 220–3. Bibcode:2013Natur.498..220Z. doi:10.1038/nature12141. PMC   3706629 . PMID   23665959.
  47. Pugh TJ, Weeraratne SD, Archer TC, Pomeranz Krummel DA, Auclair D, Bochicchio J, et al. (August 2012). "Medulloblastoma exome sequencing uncovers subtype-specific somatic mutations". Nature. 488 (7409): 106–10. Bibcode:2012Natur.488..106P. doi:10.1038/nature11329. PMC   3413789 . PMID   22820256.
  48. Parsons DW, Li M, Zhang X, Jones S, Leary RJ, Lin JC, et al. (January 2011). "The genetic landscape of the childhood cancer medulloblastoma". Science. 331 (6016): 435–9. Bibcode:2011Sci...331..435P. doi:10.1126/science.1198056. PMC   3110744 . PMID   21163964.
  49. Jones DT, Jäger N, Kool M, Zichner T, Hutter B, Sultan M, et al. (August 2012). "Dissecting the genomic complexity underlying medulloblastoma". Nature. 488 (7409): 100–5. Bibcode:2012Natur.488..100J. doi:10.1038/nature11284. PMC   3662966 . PMID   22832583.
  50. Juhlin CC, Stenman A, Haglund F, Clark VE, Brown TC, Baranoski J, et al. (September 2015). "Whole-exome sequencing defines the mutational landscape of pheochromocytoma and identifies KMT2D as a recurrently mutated gene". Genes, Chromosomes & Cancer. 54 (9): 542–54. doi:10.1002/gcc.22267. PMC   4755142 . PMID   26032282.
  51. Morin RD, Mendez-Lago M, Mungall AJ, Goya R, Mungall KL, Corbett RD, et al. (July 2011). "Frequent mutation of histone-modifying genes in non-Hodgkin lymphoma". Nature. 476 (7360): 298–303. Bibcode:2011Natur.476..298M. doi:10.1038/nature10351. PMC   3210554 . PMID   21796119.
  52. da Silva Almeida AC, Abate F, Khiabanian H, Martinez-Escala E, Guitart J, Tensen CP, Vermeer MH, Rabadan R, Ferrando A, Palomero T (December 2015). "The mutational landscape of cutaneous T cell lymphoma and Sézary syndrome". Nature Genetics. 47 (12): 1465–70. doi:10.1038/ng.3442. PMC   4878831 . PMID   26551667.
  53. Kandoth C, McLellan MD, Vandin F, Ye K, Niu B, Lu C, et al. (October 2013). "Mutational landscape and significance across 12 major cancer types". Nature. 502 (7471): 333–9. Bibcode:2013Natur.502..333K. doi:10.1038/nature12634. PMC   3927368 . PMID   24132290.
  54. Gao YB, Chen ZL, Li JG, Hu XD, Shi XJ, Sun ZM, et al. (October 2014). "Genetic landscape of esophageal squamous cell carcinoma". Nature Genetics. 46 (10): 1097–102. doi:10.1038/ng.3076. PMID   25151357. S2CID   32172173.
  55. Lin DC, Hao JJ, Nagata Y, Xu L, Shang L, Meng X, Sato Y, Okuno Y, Varela AM, Ding LW, Garg M, Liu LZ, Yang H, Yin D, Shi ZZ, Jiang YY, Gu WY, Gong T, Zhang Y, Xu X, Kalid O, Shacham S, Ogawa S, Wang MR, Koeffler HP (May 2014). "Genomic and molecular characterization of esophageal squamous cell carcinoma". Nature Genetics. 46 (5): 467–73. doi:10.1038/ng.2935. PMC   4070589 . PMID   24686850.
  56. Song Y, Li L, Ou Y, Gao Z, Li E, Li X, et al. (May 2014). "Identification of genomic alterations in oesophageal squamous cell cancer". Nature. 509 (7498): 91–5. Bibcode:2014Natur.509...91S. doi:10.1038/nature13176. PMID   24670651. S2CID   4467061.
  57. Sausen M, Phallen J, Adleff V, Jones S, Leary RJ, Barrett MT, et al. (July 2015). "Clinical implications of genomic alterations in the tumour and circulation of pancreatic cancer patients". Nature Communications. 6: 7686. Bibcode:2015NatCo...6.7686S. doi:10.1038/ncomms8686. PMC   4634573 . PMID   26154128.
  58. Grasso CS, Wu YM, Robinson DR, Cao X, Dhanasekaran SM, Khan AP, et al. (July 2012). "The mutational landscape of lethal castration-resistant prostate cancer". Nature. 487 (7406): 239–43. Bibcode:2012Natur.487..239G. doi:10.1038/nature11125. PMC   3396711 . PMID   22722839.

This article incorporates text from the United States National Library of Medicine, which is in the public domain.