ChapterPDF AvailableLiterature Review

Genetic Studies on Mammalian DNA Methyltransferases

Authors:

Abstract and Figures

Cytosine methylation at the C5-position, generating 5-methylcytosine (5mC), is a DNA modification found in many eukaryotic organisms, including fungi, plants, invertebrates, and vertebrates, albeit its levels vary greatly in different organisms. In mammals, cytosine methylation occurs predominantly in the context of CpG dinucleotides, with the majority (60–80 %) of CpG sites in their genomes being methylated. DNA methylation plays crucial roles in the regulation of chromatin structure and gene expression and is essential for mammalian development. Aberrant changes in DNA methylation levels and patterns are associated with various human diseases, including cancer and developmental disorders. DNA methylation is mediated by three active DNA methyltransferases (Dnmts), namely, Dnmt1, Dnmt3a, and Dnmt3b, in mammals. Over the last two decades, genetic manipulations of these enzymes, as well as their regulators, in mice have greatly contributed to our understanding of the biological functions of DNA methylation in mammals. In this chapter, we discuss genetic studies on mammalian Dnmts, focusing on their roles in embryogenesis, cellular differentiation, genomic imprinting, and X-chromosome inactivation.
Content may be subject to copyright.
123
© Springer International Publishing Switzerland 2016
A. Jeltsch, R.Z. Jurkowska (eds.), DNA Methyltransferases - Role and Function,
Advances in Experimental Medicine and Biology 945,
DOI 10.1007/978-3-319-43624-1_6
J. Dan
Department of Epigenetics and Molecular Carcinogenesis , The University of Texas MD
Anderson Cancer Center , 1808 Park Road 1C , Smithville , TX 78957 , USA
Center for Cancer Epigenetics , The University of Texas MD Anderson Cancer Center ,
1808 Park Road 1C , Smithville , TX 78957 , USA
T. Chen (*)
Department of Epigenetics and Molecular Carcinogenesis , The University of Texas MD
Anderson Cancer Center , 1808 Park Road 1C , Smithville , TX 78957 , USA
Center for Cancer Epigenetics , The University of Texas MD Anderson Cancer Center ,
1808 Park Road 1C , Smithville , TX 78957 , USA
Graduate School of Biomedical Sciences at Houston , Houston , TX 77030 , USA
e-mail: tchen2@mdanderson.org
Genetic Studies on Mammalian DNA
Methyltransferases
Jiameng Dan and Taiping Chen
Abstract
Cytosine methylation at the C5-position, generating 5-methylcytosine (5mC), is
a DNA modifi cation found in many eukaryotic organisms, including fungi,
plants, invertebrates, and vertebrates, albeit its levels vary greatly in different
organisms. In mammals, cytosine methylation occurs predominantly in the con-
text of CpG dinucleotides, with the majority (60–80 %) of CpG sites in their
genomes being methylated. DNA methylation plays crucial roles in the regula-
tion of chromatin structure and gene expression and is essential for mammalian
development. Aberrant changes in DNA methylation levels and patterns are asso-
ciated with various human diseases, including cancer and developmental disor-
ders. DNA methylation is mediated by three active DNA methyltransferases
(Dnmts), namely, Dnmt1, Dnmt3a, and Dnmt3b, in mammals. Over the last two
decades, genetic manipulations of these enzymes, as well as their regulators, in
mice have greatly contributed to our understanding of the biological functions of
DNA methylation in mammals. In this chapter, we discuss genetic studies on
124
mammalian Dnmts, focusing on their roles in embryogenesis, cellular differen-
tiation, genomic imprinting, and X-chromosome inactivation.
Abbreviations
5caC 5-Carboxylcytosine
5fC 5-Formylcytosine
5hmC 5-Hydroxymethylcytosine
5mC 5-Methylcytosine
ADCA-DN Autosomal dominant cerebellar ataxia deafness and narcolepsy
ADD ATRX-Dnmt3-Dnmt3L
AML Acute myeloid leukemia
BAH Bromo-adjacent homology
DKO Double knockout
DMR Differentially methylated region
DNMT DNA methyltransferase
ES Embryonic stem
EST Expressed sequence tag
HP1 Heterochromatin protein 1
HSAN IE Hereditary sensory and autonomic neuropathy with dementia and
hearing loss type IE
ICF Immunodefi ciency centromeric instability and facial anomalies
ICM Inner cell mass
ICR Imprinting control region
KAP1 KRAB-associated protein 1
KRAB Krüppel -associated box
lncRNA Long non-coding RNA
MBD3 Methyl CpG-binding domain protein-3
MEF Mouse embryonic broblast
MTA2 Metastasis tumor antigen 2
NLS Nuclear localization signal
NuRD Nuclear remodeling and histone deacetylation
PBD PCNA-binding domain
PCNA Proliferating cell nuclear antigen
PGC Primordial germ cell
PHD Plant homeodomain
PRC2 Polycomb repressive complex 2
PWWP Proline-tryptophan-tryptophan-proline
RFTS Replication foci-targeting sequence
RING Really Interesting New Gene
SRA SET- and RING-associated
TDG Thymine DNA glycosylase
TTD Tandem tudor domain
UBL Ubiquitin-like
J. Dan and T. Chen
125
Uhrf1 Ubiquitin-like with PHD and RING nger domains 1
Xa Active X chromosome
XCI X-chromosome inactivation
Xi Inactive X chromosome
Xic X-inactivation center
Xist X-inactive-specifi c transcript
Xm Maternal X chromosome
Xp Paternal X chromosome
1 Distinct Roles of Dnmt1 and Dnmt3 Families
in DNA Methylation
In 1975, long before the identifi cation of any mammalian DNA methyltransferase,
Holliday and Pugh and Riggs independently proposed a theory that DNA methyla-
tion could serve as a heritable epigenetic mark for cellular memory. Recognizing that
the CpG dinucleotide is self-complementary, they postulated that methylated and
unmethylated CpG sites could be copied when cells divide so that DNA methylation
patterns would be replicated semiconservatively like the base sequence of DNA itself
(Holliday and Pugh 1975 ; Riggs 1975 ). A prediction of the theory was the existence
of at least two DNA methyltransferase activities: de novo methyltransferase(s) would
methylate unmodifi ed DNA and establish DNA methylation patterns, and mainte-
nance methyltransferase(s) would recognize hemimethylated sites and “copy” the
methylation patterns from the parental strand onto the daughter strand at each round
of DNA replication.
1.1 Dnmt1: The Major Maintenance Methyltransferase
The rst mammalian DNA methyltransferase gene, Dnmt1 , was cloned from murine
cells (Bestor et al. 1988 ). The Dnmt1 locus has several transcription start sites and
produces two major protein products (Mertineit et al. 1998 ; Rouleau et al. 1992 ).
Transcription initiation within a somatic cell-specifi c exon (exon 1 s) results in the
Dnmt1s isoform (generally referred to as Dnmt1) which consists of 1620 amino
acids. Transcription initiation within an oocyte-specifi c exon (exon 1o) produces a
transcript that utilizes a downstream AUG as the translation initiation codon. As a
result, the oocyte-specifi c isoform, Dnmt1o, lacks the N-terminal 118 amino acids
of Dnmt1s (Mertineit et al. 1998 ). Although Dnmt1o is more stable than Dnmt1s,
genetic evidence suggests no functional difference between these isoforms (Ding
and Chaillet 2002 ). Human DNMT1, consisting of 1616 amino acids, is nearly 80 %
identical to the mouse Dnmt1 at the amino acid level.
Dnmt1 contains a C-terminal catalytic domain containing characteristic amino
acid sequence motifs that are homologous to bacterial DNA methyltransferases and
an N-terminal regulatory region that is not present in bacterial enzymes (Bestor
Genetic Studies on Mammalian DNA Methyltransferases
126
et al. 1988 ). The N-terminal regulatory region contains several functional domains,
including a proliferating cell nuclear antigen (PCNA)-binding domain (PBD)
responsible for the interaction with the DNA replication machinery, a nuclear local-
ization signal (NLS), a replication foci-targeting sequence (RFTS) that mediates the
association with late replicating heterochromatin, a zinc-fi nger CXXC domain that
recognizes unmethylated CpG-containing DNA, and a pair of bromo-adjacent
homology (BAH) domains (Fig. 1a ). Recent structural data revealed that the RFTS
domain binds to the catalytic domain and blocks the catalytic center, suggesting an
autoinhibitory role in the regulation of enzymatic activity (Takeshita et al. 2011 ).
In vitro biochemical assays revealed that, although Dnmt1 is capable of methyl-
ating both unmethylated and hemimethylated CpG dinucleotides, its activity toward
hemimethylated substrates is far more effi cient (Pradhan et al. 1999 ). Dnmt1 is
ubiquitously expressed through development, with high levels in proliferating cells.
Dnmt1 associates with the DNA replication machinery at S phase and with hetero-
chromatin at late S and G2 phases (Chuang et al. 1997 ; Easwaran et al. 2004 ;
Leonhardt et al. 1992 ; Schneider et al. 2013 ), suggesting that Dnmt1-mediated
methylation is coupled to DNA replication. These fi ndings supported the notion that
Dnmt1 functions as a maintenance enzyme (Fig. 1b ). However, because Dnmt1, the
only known DNA methyltransferase at the time, also has de novo methylation activ-
ity in vitro , it was initially debated whether de novo methylation and maintenance
methylation are carried out by Dnmt1 alone or by two or more distinct enzymes
(Bestor 1992 ).
Genetic studies in mouse models and murine cells helped settling the debate.
Several Dnmt1 mutant alleles were generated by gene targeting. The Dnmt1 n
allele
(n stands for N-terminal disruption) was reported in 1992 (Li et al. 1992 ). This
allele, in which a genomic region coding 60 amino acids near the N-terminal end
was replaced by a neomycin resistance cassette, is a partial loss-of-function (hypo-
morphic) mutation. Dnmt1 n / n
embryos have a ~70 % reduction in global DNA meth-
ylation and show mid-gestation lethality (Li et al. 1992 ). Subsequently, the Dnmt1 s
allele (s stands for Sal I site) was reported, which had a neomycin resistance cassette
inserted into a Sal I site in exon 17, disrupting the RFTS (Li et al. 1993 ). The Dnmt1 s
allele is functionally more severe than the Dnmt1 n
allele, as Dnmt1 s / s
embryos show
lower levels of DNA methylation and earlier lethality (Lei et al. 1996 ). However, it
was unclear whether the Dnmt1 s
allele was a null mutation, because the C-terminal
catalytic domain was intact. To completely inactivate Dnmt1 , Lei et al. generated
the Dnmt1 c
allele (c stands for C-terminal disruption) by disrupting the catalytic
domain, including the highly conserved PC and ENV motifs that are essential for
enzymatic activity (Lei et al. 1996 ). The development of Dnmt1 c / c
embryos is
arrested prior to the 8-somite stage, signifi cantly earlier than the developmental
phenotype of Dnmt1 n / n
embryos, while the viability and proliferation of Dnmt1 null
embryonic stem (ES) cells are not affected (Lei et al. 1996 ). Inactivation of Dnmt1
by mutating the cysteine (C1229) residue at the catalytic center (PC motif) results
in similar developmental defects (Takebayashi et al. 2007 ), suggesting that the phe-
notype is largely due to the loss of catalytic activity. DNA methylation analyses
revealed that Dnmt1 null embryos and ES cells contain low but stable levels of
J. Dan and T. Chen
127
F i g . 1 DNA methyltransferases and major regulatory proteins involved in DNA methylation. ( a )
Schematic diagrams of Dnmt1, Dnmt3a, Dnmt3b, Dnmt3L, and Uhrf1. The C-terminal catalytic
domains of the Dnmt1 and Dnmt3 families are conserved (the highly conserved signature motifs I,
IV, VI, IX, and X are shown), but their N-terminal regulatory regions are distinct. Functional
domains of the proteins are indicated. PBD PCNA-binding domain, NLS nuclear localization sig-
nal, RFTS replication foci-targeting sequence, CXXC a cysteine-rich domain implicated in binding
CpG-containing DNA sequences, BAH bromo-adjacent homology domain, PWWP proline-
tryptophan- tryptophan-proline domain, ADD ATRX-Dnmt3-Dnmt3L domain, and UBL ubiquitin-
like domain; TTD tandem tudor domain, PHD plant homeodomain, SRA SET- and RING-associated
domain, and RING Really Interesting New Gene domain. ( b ) De novo and maintenance methyl-
transferase activities. The de novo methyltransferases Dnmt3a and Dnmt3b, in complex with their
accessory factor Dnmt3L, methylate unmodifi ed DNA and establish methylation patterns. At each
round of DNA replication, the maintenance methyltransferase Dnmt1, aided by its accessory factor
Uhrf1, “copies” the methylation pattern from the parental strand onto the daughter strand. Open
circles represent unmethylated CpG dinucleotides, and fi lled circles represent methylated CpG
dinucleotides
Genetic Studies on Mammalian DNA Methyltransferases
128
5-methylcytosine (5mC) and methyltransferase activity. Moreover, the de novo
methylation activity is not impaired by Dnmt1 loss, as integrated provirus DNA in
MoMuLV-infected Dnmt1 null ES cells becomes methylated at a similar rate as in
wild-type ES cells (Lei et al. 1996 ). Taken together, these studies provided compel-
ling evidence for the existence of one or more DNA methyltransferases that are
important for de novo methylation.
1.2 Dnmt2/Trdmt1: A tRNA Methyltransferase
Results from genetic studies of Dnmt1 prompted the search for more DNA meth-
yltransferase genes. In 1998, several groups reported the identifi cation of a second
putative DNA methyltransferase gene, named Dnmt2 , which encodes a protein of
391 amino acids in human or 415 amino acids in mouse (Okano et al. 1998b ; Van
den Wyngaert et al. 1998 ; Yoder and Bestor 1998 ). Despite the presence of all the
conserved motifs shared by known prokaryotic and eukaryotic DNA cytosine
methyltransferases, Dnmt2 has no detectable DNA methyltransferase activity in
standard in vitro assays. Furthermore, inactivation of Dnmt2 in mouse ES cells by
gene targeting has no effect on preexisting genomic methylation patterns or on the
ability to methylate newly integrated retrovirus DNA de novo (Okano et al.
1998b ). Indeed, a subsequent study demonstrated that Dnmt2 is a tRNA methyl-
transferase, specifi c for cytosine 38 in the anticodon loop of aspartic acid tRNA,
and has been renamed tRNA aspartic acid (D) methyltransferase 1 (Trdmt1) (Goll
et al. 2006 ).
1.3 Dnmt3a and Dnmt3b: The De Novo Methyltransferases
By searching an expressed sequence tag (EST) database using full-length bacterial
type II cytosine-C5 methyltransferase sequences as queries, Okano et al . identifi ed
two additional homologous genes, Dnmt3a and Dnmt3b , in both mouse and human.
Their protein products contain the highly conserved DNA methyltransferase motifs
in their C-terminal regions, but their N-terminal regulatory regions are unrelated to
that of Dnmt1 (Okano et al. 1998a ). The N-terminal regions of Dnmt3a and Dnmt3b
contain a variable region and two conserved domains, the proline-tryptophan-
tryptophan- proline (PWWP) domain and the ATRX-Dnmt3-Dnmt3L (ADD)
domain (Fig. 1a ). Both domains are implicated in chromatin binding. The PWWP
domain is required for heterochromatin localization and mediates H3K36me3 bind-
ing (Baubec et al. 2015 ; Chen et al. 2004 ; Dhayalan et al. 2010 ). The ADD domain
interacts with the N-terminal tail of histone H3, and the interaction is disrupted by
various posttranslational modifi cations of H3, including di- and trimethylation of
K4, acetylation of K4, and phosphorylation of T3, S10, or T11 (Otani et al. 2009 ;
Noh et al. 2015 ; Zhang et al. 2010 ).
Dnmt3a produces two major isoforms, Dnmt3a and Dnmt3a2, driven by differ-
ent promoters (Chen et al. 2002 ). The full-length Dnmt3a protein, consisting of
J. Dan and T. Chen
129
908 amino acids in mouse and 912 amino acids in human, is expressed ubiqui-
tously at relatively low levels. The Dnmt3a2 transcript is initiated in intron 6 of the
Dnmt3a gene and encodes a protein that lacks the N-terminal 219 (in mouse) or
223 (in human) amino acids of Dnmt3a. Dnmt3a2, which is catalytically active, is
the predominant form in mouse ES cells, early embryos, and developing germ
cells, as well as human embryonal carcinoma cells, and is also detectable in spleen
and thymus (Chen et al. 2002 ). The Dnmt3b gene produces multiple alternatively
spliced isoforms, some of which encode catalytically inactive protein products.
The longest isoform, Dnmt3b1, consists of 859 amino acids in mouse and 853
amino acids in human, respectively. Both active and inactive Dnmt3b isoforms
appear to co-express in most, if not all, cell types. For example, Dnmt3b1, an active
form, and Dnmt3b6, an inactive form, are the predominant forms in mouse ES
cells, whereas Dnmt3b2, an active form, and Dnmt3b3, an inactive form, are
expressed at low levels in many somatic cells (Chen et al. 2002 ). There is evidence
that catalytically inactive Dnmt3b protein products may play regulatory roles in
DNA methylation. For example, overexpression of human DNMT3B7, a truncated
isoform frequently found in cancer cells, leads to higher levels of total genomic
methylation and altered gene expression in both transgenic mice and human cancer
cells (Ostler et al. 2012 ; Shah et al. 2010 ).
Several lines of evidence suggest the involvement of Dnmt3a and Dnmt3b in de
novo DNA methylation (Fig. 1b ). First, Dnmt3a and Dnmt3b are highly expressed
in early embryos (and ES cells) and developing germ cells, where an active de novo
methylation takes place, but are downregulated in somatic tissues and when ES
cells are induced to differentiate (Okano et al. 1998a ). Second, recombinant
Dnmt3a and Dnmt3b proteins methylate unmethylated and hemimethylated DNA
with equal effi ciency (Okano et al. 1998a ). Genetic studies provided defi nitive
evidence that Dnmt3a and Dnmt3b were the long-sought de novo methyltransfer-
ases. Inactivation of both Dnmt3a and Dnmt3b by gene targeting blocks de novo
methylation in ES cells and early embryos but has no effect on maintenance of
imprinted methylation patterns (Okano et al. 1999 ). Dnmt3a defi ciency also leads
to failure to establish DNA methylation imprints in developing germ cells (Kaneda
et al. 2004 ).
It is worth noting that the de novo DNA methyltransferase activity of Dnmt3a
and Dnmt3b is not only essential for the establishment of new DNA methylation
patterns but also important for the faithful maintenance of these patterns. In cul-
ture, Dnmt3a / 3b double knockout (DKO) ES cells exhibit gradual loss of global
DNA methylation and, after multiple passages, show severe hypomethylation
(Chen et al. 2003 ), suggesting that Dnmt1 and Dnmt3 enzymes have distinct and
nonredundant functions but act cooperatively in the maintenance of global DNA
methylation. Based on the kinetics of DNA methylation loss, it was proposed that
Dnmt1 is the major maintenance methyltransferase that, upon DNA replication,
methylates hemimethylated CpG sites with high effi ciency but not absolute fi del-
ity, whereas Dnmt3a and Dnmt3b, as de novo methyltransferases, act as “proof-
reading” enzymes that methylate the hemimethylated CpG sites missed by Dnmt1
(Chen et al.
2003 ).
Genetic Studies on Mammalian DNA Methyltransferases
130
1.4 Dnmt3L: A Regulator of De Novo Methylation
A third member of the Dnmt3 family, Dnmt3 -like ( Dnmt3L ), was originally isolated
by database analysis of the human genome sequence (Aapola et al. 2000 ). Its murine
homolog was subsequently identifi ed (Aapola et al. 2001 ; Hata et al. 2002 ). The
human and mouse Dnmt3L proteins consist of 387 and 421 amino acids, respec-
tively. Dnmt3L contains an ADD domain, but lacks a PWWP domain, in the
N-terminal region. Its C-terminal region is highly related to the catalytic domains of
Dnmt3a and Dnmt3b, but lacks some motifs essential for enzymatic activity, includ-
ing the PC dipeptide at the active site and the sequence motif involved in binding of
the methyl donor S -adenosyl-L-methionine (Aapola et al. 2000 , 2001 ; Hata et al.
2002 ) (Fig. 1a ). Therefore, Dnmt3L has no methyltransferase activity. However,
Dnmt3L has been shown to interact with Dnmt3a and Dnmt3b, stimulate their enzy-
matic activities, and target them to chromatin (Hata et al. 2002 ; Jia et al. 2007 ; Ooi
et al. 2007 ; Suetake et al. 2004 ). The expression pattern of Dnmt3L during develop-
ment is also strikingly similar to that of Dnmt3a and Dnmt3b, including high expres-
sion in developing germ cells, early embryos, and ES cells (Hata et al. 2002 ). These
ndings indicate that Dnmt3L may regulate Dnmt3a/3b functions (Fig. 1b ). Genetic
studies indeed demonstrate that Dnmt3L is an essential accessory factor of Dnmt3a
in the germ line. Dnmt3L homozygous null mice are viable and grossly normal, but
both male and female mice are infertile (Bourc’his et al. 2001 ; Hata et al. 2002 ).
Male mice show activation of retrotransposons in spermatogonia and spermato-
cytes, due to failure to establish methylation at these elements, and are azoospermic
(Bourc’his and Bestor 2004 ). Female mice fail to establish maternal methylation
imprints in oocytes, and, as a result, embryos derived from these oocytes cannot
survive beyond mid-gestation (Bourc’his et al. 2001 ; Hata et al. 2002 ). The pheno-
type is indistinguishable from that of mice with conditional Dnmt3a deletion in
germ cells (Kaneda et al. 2004 ). Recently, Dnmt3L was shown to antagonize DNA
methylation at H3K4me3/K27me3 bivalent promoters, which are often associated
with developmental genes, and favor DNA methylation at gene bodies in ES cells.
It was suggested that Dnmt3L, via its ADD domain, interacts with Polycomb repres-
sive complex 2 (PRC2) in competition with Dnmt3a and Dnmt3b to maintain low
methylation levels at regions with H3K27me3, thus maintaining hypomethylation at
promoters of bivalent developmental genes (Neri et al. 2013 ). The physiological
relevance of this fi nding remains to be determined, given that zygotic Dnmt3L is not
required for embryonic development and postnatal survival (Bourc’his et al. 2001 ;
Hata et al. 2002 ).
1.5 Uhrf1: A Regulator of Maintenance Methylation
Besides Dnmts, a number of DNA methylation regulators have been identifi ed,
including the multi-domain protein Uhrf1 (ubiquitin-like with PHD and RING fi n-
ger domains 1), also known as NP95 (mouse) and ICBP90 (human) (Fig. 1a ).
Genetic studies demonstrated an essential role for Uhrf1 in maintaining DNA
J. Dan and T. Chen
131
methylation (Fig. 1b ). Uhrf1 defi ciency leads to embryonic lethality and global
DNA hypomethylation (Bostick et al. 2007 ; Muto et al. 2002 ; Sharif et al. 2007 ),
resembling the phenotype of Dnmt1 defi ciency. Cellular and biochemical evidence
suggested functional interactions between Uhrf1 and Dnmt1. Uhrf1 co-localizes
with Dnmt1 at DNA replication foci and heterochromatin, and Dnmt1 fails to enrich
at these regions in the absence of Uhrf1 (Bostick et al. 2007 ; Liu et al. 2013 ; Sharif
et al. 2007 ). These fi ndings suggest that Uhrf1 is a key accessory factor for directing
Dnmt1 to hemimethylated CpG sites. However, it remains somewhat controversial
as to whether Uhrf1 directly recruits Dnmt1 or indirectly controls Dnmt1 localiza-
tion by affecting chromatin structure. Uhrf1 harbors fi ve known functional domains:
a ubiquitin-like domain (UBL) at the N-terminus, followed by a tandem tudor
domain (TTD), a plant homeodomain (PHD), a SET- and RING-associated (SRA)
domain, and a Really Interesting New Gene (RING) domain (Fig. 1a ). All the
domains, with the exception of UBL, have been shown to be important for Dnmt1
subnuclear localization and maintenance of DNA methylation. Biochemical and
structural evidence revealed that the SRA domain preferentially binds hemimethyl-
ated DNA and is thought to play an important role in loading Dnmt1 onto newly
synthesized DNA substrates (Arita et al. 2008 ; Avvakumov et al. 2008 ; Bostick
et al. 2007 ; Hashimoto et al. 2008 ; Sharif et al. 2007 ). The association of Uhrf1 with
heterochromatin is also mediated by TTD, which contains an aromatic cage for
binding of the heterochromatic H3K9me3 mark. The PHD acts in combination with
TTD to read the H3K9me3 mark and, additionally, interacts with histone H3 tails
with unmethylated R2 (H3R2me0) (Cheng et al. 2013 ; Liu et al. 2013 ; Rothbart
et al. 2012 , 2013 ; Rottach et al. 2010 ). Recent studies suggested that Uhrf1, via the
E3 ligase activity of its RING domain, mediates ubiquitylation of H3K23 and
H3K18, creating binding sites for Dnmt1 (Nishiyama et al. 2013 ; Qin et al. 2015 ).
It is worth noting that Uhrf1 has also been shown to control Dnmt1 ubiquitylation
and stability (Du et al. 2010 ; Qin et al. 2011 ). Indeed, a recent study revealed that
Uhrf1 overexpression results in DNA hypomethylation, due to destabilization and
delocalization of Dnmt1, which led the authors to propose that Uhrf1 overexpres-
sion, which is frequently observed in cancer cells, is a mechanism underlying global
DNA hypomethylation in cancer (Mudbhary et al. 2014 ).
2 Dnmts in Embryonic Development and Cellular
Differentiation
2.1 Dynamic Changes of DNA Methylation During Early
Embryogenesis
DNA methylation is relatively stable in somatic tissues but exhibits dynamic
changes in early embryos. During preimplantation development, both the maternal
and paternal genomes undergo global DNA demethylation, albeit the mechanisms
involved are distinct. The maternal genome is demethylated mainly through DNA
replication-dependent passive dilution because of defi cient maintenance
Genetic Studies on Mammalian DNA Methyltransferases
132
methylation, presumably due to the exclusion of Dnmt1 from the nucleus (Hirasawa
et al. 2008 ; Howell et al. 2001 ). In contrast, demethylation of the paternal genome
involves both active and passive mechanisms. Shortly after fertilization and before
the fi rst cell division, the 5mC dioxygenase Tet3 converts the majority of 5mC in the
male pronucleus to 5-hydroxymethylcytosine (5hmC) (Gu et al. 2011 ; Wossidlo
et al. 2011 ). 5hmC can be further oxidized to 5-formylcytosine (5fC) and
5- carboxylcytosine (5caC), which can be excised by thymine DNA glycosylase
(TDG) and replaced by unmodifi ed cytosine (He et al. 2011 ; Ito et al. 2011 ). 5hmC,
5fC, and 5caC can also be passively diluted during cleavage divisions (Inoue et al.
2011 ; Inoue and Zhang 2011 ). As a result of these processes, DNA methylation
marks inherited from gametes are largely erased by the blastocyst stage, with the
exception of imprinting control regions (ICRs) and some retroelements, which
resist this wave of global demethylation (Borgel et al. 2010 ; Smith et al. 2012 ).
Around the time of implantation, de novo methylation takes place when the inner
cell mass (ICM) starts to differentiate to form the embryonic ectoderm. Lineage-
specifi c DNA methylation patterns are then stably maintained.
2.2 Embryonic and Adult Phenotypes of Dnmt Mutant Mice
Most of our knowledge about the signifi cance of DNA methylation in mammalian
development came from genetic manipulations of Dnmt genes in mice. Results from
characterization of Dnmt mutant mice demonstrated that the establishment of
embryonic methylation patterns requires both de novo and maintenance Dnmts and
that maintaining genomic methylation above a threshold level is essential for embry-
onic development (Lei et al. 1996 ; Li et al. 1992 ; Okano et al. 1999 ). Complete
inactivation of Dnmt1 results in the arrest of embryonic development between pre-
somite and 8-somite stage around E9.5 (Lei et al. 1996 ). DNA methylation analysis
showed that embryos defi cient for Dnmt1 undergo dramatic decreases in global
DNA methylation (Lei et al. 1996 ; Li et al. 1992 ), in agreement with its role as the
major maintenance Dnmt. Disruption of Dnmt3b also leads to embryonic lethality
after E12.5, with multiple defects, including growth impairment and rostral neural
tube defects (Okano et al. 1999 ). In contrast, Dnmt3a -defi cient mice develop to term
and appear to be normal at birth but become runted and die at about 4 weeks (Okano
et al. 1999 ). Consistent with the developmental phenotypes, DNA methylation anal-
ysis of E9.5 embryos revealed that germ line-specifi c genes, pluripotency genes,
hematopoietic genes, and eye genes are severely hypomethylated in the absence of
Dnmt3b but not of Dnmt3a (Borgel et al. 2010 ). This suggested that Dnmt3b is the
main enzyme responsible for de novo methylation during embryogenesis. Dnmt3b
shows a dynamic expression change during pre- and early postimplantation devel-
opment, with preferential expression in the trophectoderm at the mid-blastocyst
stage and subsequent transition of expression in the embryonic lineage (Hirasawa
and Sasaki 2009 ). Notably, DNA methylation at certain genes such as Brdt , Dpep3 ,
Cytip , and Crygd is only partially reduced in Dnmt3b -/-
embryos (Borgel et al. 2010 ),
suggesting that Dnmt3a cooperates with Dnmt3b to methylate some loci. Indeed,
J. Dan and T. Chen
133
Dnmt3a / 3b DKO embryos exhibit more severe defects than Dnmt3b -/-
embryos.
Specifi cally, DKO embryos show smaller size, lack somites, do not undergo embry-
onic turning, and die before E11.5, indicating that their growth and morphogenesis
are arrested shortly after gastrulation (Okano et al. 1999 ).
Conditional knockout (KO) studies have also demonstrated that Dnmts and DNA
methylation are essential in various organs and tissues. For example, disruption of
both Dnmt1 and Dnmt3a in forebrain excitatory neurons leads to abnormal synaptic
plasticity and defi cits in learning and memory (Feng et al. 2010 ). Conditional dele-
tion of Dnmt1 at sequential stages of T cell development has also revealed a critical
role for DNA methylation in T cell development, function, and survival. Specifi cally,
deletion of Dnmt1 in early double-negative thymocytes leads to an impaired sur-
vival of TCRαβ(+) cells and the generation of atypical CD8(+) TCRγδ(+) cells and
deletion of Dnmt1 in double-positive thymocytes impairs activation-induced prolif-
eration but differentially enhanced cytokine mRNA expression by naive peripheral
T cells (Lee et al. 2001 ).
2.3 Cellular Defects of Dnmt Mutations
The mechanisms underlying the developmental defects observed in Dnmt mutant
mice are not fully understood. Dnmt1, Dnmt3a, and Dnmt3b are all highly expressed
in pluripotent ES cells, but disruption of these genes individually, both Dnmt3a and
Dnmt3b , or even all three Dnmts , has no deleterious effects on mouse ES cells in the
undifferentiated state (Lei et al. 1996 ; Li et al. 1992 ; Okano et al. 1999 ; Tsumura
et al. 2006 ). However, Dnmt1 -/-
and Dnmt3a / 3b DKO ES cells die upon induction of
differentiation (Chen et al. 2003 ; Lei et al. 1996 ; Tucker et al. 1996 ). Interestingly,
a recent study showed that, in contrast to mouse ES cells, human ES cells require
DNMT1 , but not DNMT3A and DNMT3B , for survival (Liao et al. 2015 ). It is well
established that mouse and human ES cells represent different pluripotent states,
with human ES cells resembling the more mature epiblast state (Tesar et al. 2007 ),
which may explain the sensitivity of human ES cells to loss of DNA methylation.
During development, the effects of DNA methylation defi ciency become apparent
during or after gastrulation, when the embryo differentiates to form the three germ
layers (Lei et al. 1996 ; Li et al. 1992 ; Okano et al. 1999 ). Conditional inactivation
of Dnmt1 in mouse embryonic fi broblasts (MEFs) leads to severe hypomethylation
and cell death, and Dnmt3b -defi cient MEFs show modest hypomethylation, chro-
mosomal instability, and abnormal cell proliferation (Dodge et al. 2005 ; Jackson-
Grusby et al. 2001 ). Furthermore, although a hypomorphic mutation affecting the
N-terminal region of human DNMT1 has no effect on the survival and proliferation
of the colon cancer cell line HCT116 (Rhee et al. 2000 ), disruption of the DNMT1
catalytic domain in HCT116 leads to mitotic catastrophe and cell death (Chen et al.
2007 ). Taken together, these results suggest crucial roles for DNA methylation in
cellular differentiation and in the viability and proper functioning of differentiated
cells. Deregulation of gene expression likely plays a major role in the developmen-
tal and cellular defects associated with Dnmt mutations.
Genetic Studies on Mammalian DNA Methyltransferases
134
3 Dnmts in Genomic Imprinting
In early 1980s, elegant nuclear transplantation experiments using pronuclear stage
embryos showed that mouse embryos constructed to contain only maternal or pater-
nal diploid genome complements fail to develop beyond mid-gestation. This sug-
gested that the parental genomes are functionally nonequivalent and marked or
“imprinted” differently during male and female gametogenesis (Barton et al. 1984 ;
McGrath and Solter 1984 ; Surani et al. 1984 ). Separate experiments using chromo-
some translocations in mice showed that specifi c chromosome segments function
differently depending on the parental origin (Cattanach and Kirk 1985 ). In early
1990s, the fi rst murine imprinted genes, Igf2r , Igf2 , and H19 , were discovered,
which are expressed only from one parental allele (Barlow et al. 1991 ; Bartolomei
et al. 1991 ; DeChiara et al. 1991 ). To date, approximately 150 imprinted genes,
which exhibit monoallelic expression strictly according to the parental origin,
have been identifi ed in mouse ( http://www.mousebook.org/mousebook-catalogs/
imprinting- resource ), and many of them are also imprinted in human. Imprinted
genes are involved in diverse biological processes, including embryonic develop-
ment, placental formation, fetal and postnatal growth, and adult behavior (Frontera
et al. 2008 ; Reik and Walter 2001 ). In human, altered expression of imprinted genes,
due to genetic and epigenetic changes, has been linked to infertility, molar preg-
nancy, and various congenital disorders such as Prader-Willi syndrome, Angelman
syndrome, Beckwith-Wiedemann syndrome, and Silver-Russell syndrome (Butler
2009 ; Tomizawa and Sasaki 2012 ; Walter and Paulsen 2003 ). Loss of imprinting
(biallelic expression or silencing of imprinted genes) is also frequently observed in
cancer (Jelinic and Shaw 2007 ).
The majority of imprinted genes are arranged in chromosomal clusters, which
usually span hundreds to thousands of kilobases. Each of the imprinting clusters is
controlled by an ICR, an essential regulatory sequence that contains one or more
differentially methylated regions (DMRs) between the two alleles. Thus, allele-
specifi c DNA methylation is believed to be the primary epigenetic mark that con-
trols the monoallelic expression of imprinted genes.
The life cycle of DNA methylation imprints consists of three major steps: estab-
lishment, maintenance, and erasure (Fig. 2 ).
3.1 Establishment of Methylation Imprints
DNA methylation imprints are acquired in the germ line, with the majority being
established during oogenesis (maternally imprinted) and only four known loci ( H19 ,
Dlk1 - Gtl2 , Rasgrf1 , and Zdbf2 ) being established during spermatogenesis (pater-
nally imprinted). Conditional deletion of Dnmt3a in primordial germ cells (PGCs)
disrupts both maternal and paternal imprinting. Embryos from crosses between con-
ditional Dnmt3a mutant females and wild-type males die around E10.5, and condi-
tional Dnmt3a mutant males are sterile due to impaired spermatogenesis (Kaneda
et al. 2004 ). Dnmt3L KO mice show an identical phenotype, with the exception of
J. Dan and T. Chen
135
one paternally methylated locus, Dlk1 - Gtl2 , which is methylated in Dnmt3L KO but
not in Dnmt3a mutant spermatogonia (Bourc’his et al. 2001 ; Hata et al. 2002 ;
Kaneda et al. 2004 ). In contrast, conditional deletion of Dnmt3b in PGCs shows no
apparent phenotype (Kaneda et al. 2004 ). These results provide compelling genetic
evidence that Dnmt3a is responsible for the establishment of germ line imprints,
and Dnmt3L is an essential cofactor for Dnmt3a in this regard.
How Dnmt3L facilitates Dnmt3a function in the germ line is not fully under-
stood. Dnmt3L, via its C-terminal domain, forms a tetrameric complex with Dnmt3a
and, via its ADD domain, interacts with the N-terminal tail of histone H3 (Hata
et al. 2002 ; Jia et al. 2007 ; Ooi et al. 2007 ; Suetake et al. 2004 ). These fi ndings led
to the hypothesis that Dnmt3L plays a critical role in targeting Dnmt3a to specifi c
chromatin regions, including imprinted loci. A recent study showed that, similar to
Dnmt3L null mutant mice, mice homozygous for an engineered point mutation
(D124A) in the Dnmt3L ADD domain exhibit DNA methylation and spermatogen-
esis defects (Vlachogiannis et al. 2015 ), supporting a critical role of the ADD
domain in Dnmt3L function in the male germ line. It would be interesting to
F i g . 2 Life cycle of DNA methylation imprints. The paternal ( blue ) and maternal ( red ) methyla-
tion imprints are established during gametogenesis and transmitted to the offspring through fertil-
ization. These marks are maintained and control monoallelic expression of imprinted genes during
embryogenesis and in somatic cells throughout adult life. However, they are erased in primordial
germ cells (PGCs) before sex-specifi c methylation imprints are reestablished in later stages of
germ cell development
Genetic Studies on Mammalian DNA Methyltransferases
136
determine whether female mice homozygous for the D1124A mutation show defects
in the establishment of maternal imprints. However, the Dnmt3a ADD domain also
binds H3K4-unmethylated histone H3 (Otani et al. 2009 ; Zhang et al. 2010 ), which
raises the question of the specifi c role of the Dnmt3L ADD domain. It is possible
that Dnmt3L interacts with one or more proteins or histone marks that are not rec-
ognized by Dnmt3a.
The observation that H3K4 modifi cations disrupts the interaction between
Dnmt3 proteins and histone H3 (Ooi et al. 2007 ; Otani et al. 2009 ; Zhang et al.
2010 ) suggests that chromatin organization may be an important determinant of the
sites of de novo DNA methylation in the germ line. Indeed, genetic evidence indi-
cated that the H3K4 demethylase KDM1B (also known as LSD2 and AOF1) is
essential for the establishment of a subset of maternal imprints (Ciccone et al. 2009 ).
KDM1B is highly expressed in growing oocytes, where maternal imprints are
acquired, but shows little expression in most somatic tissues. Kdm1b KO mice are
viable and show no defects in spermatogenesis and oogenesis, and male mice are
fertile. However, oocytes from KDM1B-defi cient females exhibit global accumula-
tion of H3K4me2 and fail to establish DNA methylation imprints at a subset of
imprinted loci. Consequently, embryos derived from these oocytes die around mid-
gestation (Ciccone et al. 2009 ), similar to embryos derived from Dnmt3L- or
Dnmt3a-defi cient female mice (Bourc’his et al. 2001 ; Hata et al. 2002 ; Kaneda
et al. 2004 ). These results strongly suggested that removal of H3K4me2 is a prereq-
uisite for de novo DNA methylation. There is also evidence that transcription is an
additional requirement in specifying DNA methylation, at least at some maternally
imprinted loci. In the mouse Gnas locus, transcription initiated at the promoter of
Nesp55 , a gene upstream of the DMRs of the Gnas locus, occurs in growing oocytes,
placing a large genomic region, including the DMRs, within an active transcription
unit. Deletion of the Nesp55 promoter or insertion of a transcription termination
cassette downstream of Nesp55 to ablate transcription results in failure to establish
DNA methylation at the ICR of the Gnas locus (Chotalia et al. 2009 ; Frohlich et al.
2010 ; Williamson et al. 2011 ). Methylation of the DMR at the Snrpn locus has also
been shown to depend upon transcription (Smith et al. 2011 ).
3.2 Maintenance of Methylation Imprints
The paternal and maternal imprints are transmitted to the zygote through fertiliza-
tion, and despite extensive demethylation during preimplantation development (as
described above), parental allele-specifi c DNA methylation imprints are faithfully
maintained through development and adult life. Notably, recent genome-wide DNA
methylation analyses revealed far more differentially methylated loci in oocytes and
sperm than the number of imprinted genes (Kobayashi et al. 2012 ; Smallwood et al.
2011 ; Smith et al. 2012 ). Thus, the previous notion that imprinted loci are deter-
mined by distinct methylation patterns in gametes has been revised to the current
view that genomic imprinting results from selective maintenance of germ line-
derived allele-specifi c methylation. Genetic studies using conditional KO mice
J. Dan and T. Chen
137
demonstrated that Dnmt1, but not Dnmt3a or Dnmt3b, is responsible for maintain-
ing methylation marks at imprinted loci during preimplantation development
(Hirasawa et al. 2008 ). The oocyte-specifi c variant, Dnmt1o, is the predominant
Dnmt1 isoform in preimplantation embryos (Hirasawa et al. 2008 ; Kurihara et al.
2008 ). However, offspring of females lacking Dnmt1o exhibit only a ~50 %
reduction of methylation at certain imprinted loci (Howell et al. 2001 ). While initial
evidence suggested that the somatic form, Dnmt1s, does not express until the blas-
tocyst stage (Ratnam et al. 2002 ), subsequent work showed that Dnmt1s is present
at very low levels in the nucleus of oocytes and preimplantation embryos (Hirasawa
et al. 2008 ; Kurihara et al. 2008 ). Conditional deletion of Dnmt1 (both Dnmt1o and
Dnmt1s) in growing oocytes leads to a partial loss of methylation imprints in the
offspring (Hirasawa et al. 2008 ), resembling the effect of Dnmt1o loss (Howell
et al. 2001 ). However, ablation of both maternal and zygotic Dnmt1 results in a
complete loss of methylation at paternally and maternally methylated DMRs in
embryos (Hirasawa et al. 2008 ). Therefore, both maternal and zygotic Dnmt1 pro-
teins are necessary for the maintenance of methylation imprints in preimplantation
embryos. Dnmt1 is also responsible for the maintenance of methylation imprints in
postimplantation embryos (Li et al. 1993 ) and likely in adult somatic cells as well.
It is not well understood what confers the specifi city of Dnmt1, such that meth-
ylation is maintained at imprinted genes but not at other sequences in preimplanta-
tion embryos. Genetic and epigenetic features may distinguish imprinted loci from
other regions. Several other factors have been shown to be essential for the main-
tenance of DNA methylation imprints. PGC7 (also known as Stella and Dppa3), a
DNA-binding protein, is highly expressed in oocytes and persists in preimplanta-
tion embryos. Genetic evidence suggested that, in early embryos, maternal PGC7
plays a crucial role in protecting the maternal genome against DNA demethylation.
PGC7 also protects the paternally imprinted H19 and Rasgrf1 against demethyl-
ation (Nakamura et al. 2007 ). While the mechanisms involved remain to be deter-
mined, PGC7 has been shown to play a role in chromatin condensation during
oogenesis and to protect the maternal genome against Tet3-mediated conversion of
5mC to 5hmC in early embryos (Bian and Yu 2014 ; Liu et al. 2012 ; Nakamura
et al. 2012 ). Gene targeting experiments in mice have also implicated the involve-
ment of the Krüppel -associated box (KRAB)-containing zinc-fi nger protein
ZFP57 in the maintenance of genomic imprints (Li et al. 2008 ), and human ZFP57
mutations are associated with hypomethylation at multiple imprinted loci in
patients with transient neonatal diabetes (Mackay et al. 2008 ). ZFP57 specifi cally
binds the methylated allele of ICRs, recognizing a hexanucleotide sequence
(TGCCGC) shared by all murine ICRs and some human ICRs (Quenneville et al.
2011 ). ZFP57 interacts with KRAB-associated protein 1 (KAP1, also known as
TRIM28), which acts as a scaffold protein for various heterochromatin proteins,
including heterochromatin protein 1 (HP1), the histone H3K9 methyltransferase
Setdb1 (also known as ESET and KMT1E), the nuclear remodeling and histone
deacetylation (NuRD) complex, and Dnmt proteins and Uhrf1 (Nielsen et al. 1999 ;
Quenneville et al.
2011 ; Ryan et al. 1999 ; Schultz et al. 2001 , 2002 ; Zuo et al.
2012 ). Ablation of either maternal or zygotic KAP1 causes partial loss of DNA
Genetic Studies on Mammalian DNA Methyltransferases
138
methylation imprints, and ablation of both maternal and zygotic KAP1 leads to a
complete loss of imprinting (Lorthongpanich et al. 2013 ; Messerschmidt et al.
2012 ; Quenneville et al. 2011 ). Depletion of the NuRD components methyl CpG-
binding domain protein-3 (MBD3) or metastasis tumor antigen 2 (MTA2) also
results in reduction of methylation at some imprinted loci in preimplantation
embryos (Ma et al. 2010 ; Reese et al. 2007 ). The current view is that the ZFP57/
KAP1 complex specifi cally recruits the DNA methylation machinery, as well as
other heterochromatin proteins, to the methylated allele of ICRs to maintain
genomic imprints and control monoallelic expression of imprinted genes.
3.3 Erasure of Methylation Imprints
The last step of the imprint life cycle is the erasure of methylation imprints in
PGCs, which ensures the establishment of sex-specifi c imprints in later stages of
germ cell development. In mice, PGCs are specifi ed around E7.25 in the epiblast
of the developing embryo. Shortly afterwards, PGCs begin migrating along the
embryonic- extraembryonic interface and eventually arrive at the genital ridge by
E12.5. Recent genome-wide DNA methylation analyses reveal that PGCs undergo
demethylation in two major phases (Guibert et al. 2012 ; Kobayashi et al. 2013 ;
Popp et al. 2010 ; Seisenberger et al. 2012 ). The fi rst phase takes place during PGC
expansion and migration from ~ E8.5, which leads to a global demethylation affect-
ing almost all genomic regions. Passive demethylation likely plays a major role in
this phase, as Dnmt3a, Dnmt3b, and Uhrf1 are repressed in PGCs (Kagiwada et al.
2013 ; Kurimoto et al. 2008 ). The second phase occurs from E9.5 to E13.5 and
affects specifi c loci including ICRs, germ line-specifi c genes, and CpG islands on
the X chromosome (Guibert et al. 2012 ; Hackett et al. 2013 ; Popp et al. 2010 ;
Seisenberger et al. 2012 ; Yamaguchi et al. 2013 ). Genetic studies suggested that
Tet1- and Tet2-mediated 5mC oxidation is important in the second phase of
demethylation (Zhao and Chen 2013 ).
4 Dnmts in X-Chromosome Inactivation
In mammals, sex determination is controlled by a pair of sex chromosomes, the X
and the Y. Whereas the Y chromosome harbors very few protein-coding genes com-
pared to other chromosomes, the X chromosome has a relatively high gene density.
The balance of X-linked gene dosage between XX females and XY males is
achieved through X-chromosome inactivation (XCI), whereby one of the two X
chromosomes present in female mammals is inactivated. Marsupial mammals show
imprinted XCI, with only the paternal X chromosome (Xp) being inactivated.
Eutherian mammals exhibit two forms of XCI: imprinted Xp inactivation in the
early embryo and extraembryonic tissues and random inactivation of either Xp or
the maternal X chromosome (Xm) in the embryonic (epiblast) lineage (Payer and
Lee
2008 ; Wutz 2011 ).
J. Dan and T. Chen
139
Our knowledge about XCI mostly came from studies in the mouse. In the female
zygote, both X chromosomes appear active. Soon thereafter, a series of events result
in the inactivation of Xp. This imprinted XCI occurs in a two-step manner, with Xp
repeat elements fi rst silenced at the two-cell stage followed by Xp genic silencing
emerging at the eight- to sixteen-cell stage. Imprinted XCI is complete by the blas-
tocyst stage. Whereas inactivation of Xp is maintained in extraembryonic tissues, it
is reversed in the ICM of the blastocyst, resulting in the biallelic expression of
X-linked genes. Shortly after implantation, epiblast cells undergo random XCI
(Payer and Lee 2008 ). Murine ES cells, derived from the ICM, represent a useful
model for the study of XCI. Undifferentiated murine ES cells, like ICM cells, have
two active X chromosomes, and random XCI can be recapitulated during in vitro
differentiation (Chaumeil et al. 2004 ).
The process of XCI, which converts an X chromosome from relatively open
euchromatin to highly condensed heterochromatin (known as the “Barr body”), can
be divided into three steps: initiation of X inactivation, spreading of heterochroma-
tin to the entire chromosome, and maintenance of the inactive state (Fig. 3 ). XCI is
controlled by the X-inactivation center ( Xic ), a complex locus on the X chromo-
some that determines how many (counting step) and which X chromosomes (choice
step) will be silenced. A critical gene in Xic encodes the “X-inactive-specifi c tran-
script” ( Xist ), a 17-kb long, non-coding RNA (lncRNA). Xist is expressed only from
the presumptive inactive X chromosome (Xi) and then coats the same chromosome
in cis (Clemson et al. 1996 ). This step is necessary and suffi cient for the initiation
of XCI, as targeted disruption of the Xist gene abrogates XCI (Marahrens et al.
1997 ; Penny et al. 1996 ), and Xist transgenes on autosomes can induce autosomal
gene inactivation (Jiang et al. 2013 ; Lee and Jaenisch 1997 ; Lee et al. 1996 ). The
Xic also harbors several other genes encoding proteins and non-coding RNAs,
including the Xist antisense RNA Tsix , the Jpx and Ftx RNAs, and the E3 ubiquitin
ligase RNF12, that act as part of a sophisticated regulatory network to modulate Xist
expression in cis and in trans (Gendrel and Heard 2014 ). Xist is also required for the
spreading of XCI from the Xic to the rest of the chromosome. Xist RNA is able to
recruit PRC2, a complex responsible for the deposition of H3K27me3, which con-
tributes to chromatin and transcriptional changes during the initiation and spreading
of XCI (Zhao et al. 2008 ; da Rocha et al. 2014 ; Cifuentes-Rojas et al. 2014 ). Once
established, the globally silent state and heterochromatin structure of Xi are trans-
mitted through somatic cell division and clonally inherited. Although Xist is not
required for the maintenance of gene silencing on the Xi (Brown and Willard 1994 ;
Csankovszki et al. 1999 ), it appears to be important for maintaining the heterochro-
matin structure of Xi, as deletion of Xist leads to refolding of the Xi into a structure
resembling the active X chromosome (Xa) (Splinter et al. 2011 ).
The link between DNA methylation and XCI has been well established. In
somatic tissues, the 5 end of the Xist gene is fully methylated on Xa and completely
unmethylated on Xi. Similarly, in tissues that undergo imprinted Xp inactivation,
the paternal Xist allele is unmethylated, and the maternal allele is fully methylated
(Norris et al.
1994 ). Studies of Dnmt1 -defi cient ES cells and embryos revealed that
XCI can occur in the absence of DNA methylation, but maintenance of Xist
Genetic Studies on Mammalian DNA Methyltransferases
140
promoter methylation is necessary for its stable repression in differentiated cells
(Beard et al. 1995 ; Panning and Jaenisch 1996 ; Sado et al. 2000 ). A recent study
showed that loss of Dnmt1o disrupts imprinted XCI and accentuates placental
defects in females (McGraw et al. 2013 ). De novo DNA methylation is also dispens-
able for the initiation and propagation of XCI, as Xist expression is appropriately
regulated and XCI occurs properly in female embryos defi cient for both Dnmt3a
and Dnmt3b (Sado et al.
2004 ). Interestingly, despite multiple mechanisms involved
Fig. 3 Major steps of
X-chromosome
inactivation. The process
of X-chromosome
inactivation can be divided
into three steps: (1)
initiation ( Xist RNA is
expressed from the
presumptive inactive X
(Xi), but not the active X
(Xa)), (2) spreading ( Xist
RNA coats the entire
Xi chromosome, which
recruits other factors
(e.g., PRC2 complex,
Dnmts) to induce
heterochromatinization),
and (3) maintenance (the
highly compacted
chromatin structure and
most of the genes on Xi are
stably maintained and
clonally transmitted
through somatic cell
divisions). DNA
methylation is required for
the stable maintenance of
Xi-linked gene silencing
J. Dan and T. Chen
141
in X-chromosome gene silencing, approximately 25 % of genes on Xi escape inac-
tivation to some extent and exhibit biallelic expression in females (Carrel and
Willard 2005 ; Yang et al. 2010 ). The promoter regions of these escapee genes are
unmethylated (Weber et al. 2007 ). Furthermore, treatment of cells with the demeth-
ylating agents 5-azacytidine and 5-azadeoxycytidine has been shown to reactivate
some genes on Xi (Haaf 1995 ). Collectively, these fi ndings indicate that DNA meth-
ylation is not required for the initiation and propagation of XCI but is an essential
component of the epigenetic mechanisms that stably maintain the silent state of
Xi-linked genes.
5 Concluding Remarks
Since the discovery of mammalian Dnmts (Bestor et al. 1988 ; Okano et al. 1998a ),
great progress has been made in understanding the biological functions of DNA meth-
ylation in mammals. Genetic studies using Dnmt mutant mice and murine cells have
provided important insights into the roles of DNA methylation in various develop-
mental and cellular processes (Table 1 ). It is generally believed that DNA methyla-
tion, a relatively stable epigenetic mark, acts in concert with other epigenetic
mechanisms such as histone modifi cations to stably maintain gene silencing and chro-
matin structure. It is well documented that aberrant DNA methylation patterns are
associated with various human diseases. Studies in recent years have also identifi ed
genetic alterations affecting major components of the DNA methylation machinery,
including DNA methylation “writers” (DNMTs), “erasers” (e.g., TETs), and “read-
ers” (e.g., MeCP2), in cancer and developmental disorders (Hamidi et al. 2015 ). For
example, DNMT1 mutations are reported in two related neurodegenerative diseases
(hereditary sensory and autonomic neuropathy with dementia and hearing loss type IE
(HSAN IE), autosomal dominant cerebellar ataxia, deafness, and narcolepsy
(ADCA-DN)), DNMT3A mutations are frequently found in acute myeloid leukemia
(AML) and other hematologic malignancies, and DNMT3B mutations cause the
immunodefi ciency, centromeric instability, and facial anomalies (ICF) syndrome
(Hamidi et al. 2015 ). The mechanisms by which these mutations contribute to the
disease phenotypes are generally not well understood. Besides their values in eluci-
dating the fundamental functions of DNA methylation, Dnmt mutant mice and cells
provide important research tools for investigating the effects of DNMT mutations
found in human patients. For instance, by expressing a Dnmt3a protein harboring a
point mutation equivalent to human DNMT3A:R882H (the most prevalent DNMT3A
mutation in AML) in Dnmt3a and Dnmt3b mutant murine ES cells, we recently dem-
onstrated that this mutation, which occurs on only one allele in AML patients, not
only leads to haploinsuffi ciency of DNMT3A enzymatic activity but also exhibits
dominant-negative effect by forming functionally defi cient complexes with wild-type
DNMT3A and DNMT3B (Kim et al. 2013 ). Most of the DNMT mutations identifi ed
in patients are not null alleles, making Dnmt KO mice less ideal for modeling human
diseases. With the development of new technologies such as CRISPR/Cas9-mediated
gene editing, it now becomes more feasible to create genetically engineered animal
Genetic Studies on Mammalian DNA Methyltransferases
142
and cellular models that better recapitulate the major features of human diseases asso-
ciated with DNMT mutations. Genomic, epigenomic, transcriptomic, and proteomic
analyses of these models will be powerful approaches for defi ning the molecular
mechanisms and pathways involved in pathogenesis. Ultimately, such studies will
likely lead to novel therapeutic and preventive strategies.
Acknowledgments Work in the Chen laboratory is supported by a Rising Star Award from
Cancer Prevention and Research Institute of Texas (CPRIT, R1108) and a grant from the National
Institutes of Health (NIH, 1R01DK106418-01).
Table 1 Developmental phenotypes of Dnmt knockout mice
Gene Mutations Major developmental phenotypes References
Dnmt1 Dnmt1 n /n ~70 % reduction in global DNA
methylation, embryonic lethality at
E12.5–15.5
Li et al. (
1992 )
Dnmt1 c / c
~90 % reduction in global DNA
methylation, developmental arrest
at E8.5, embryonic lethality before
8-somite stage at ~ E9.5. Unstable
random XCI
Lei et al. (
1996 );
Sado et al. (
2000 )
Dnmt1 s / s
~90 % reduction in global DNA
methylation, developmental arrest
at E8.5, embryonic lethality
at ~ E9.5. Loss of methylation at
Xist locus and abnormal Xist
expression in male embryos
Beard et al.
(
1995 ); Lei et al.
(
1996 )
Dnmt1o -/-
Maternal-effect phenotype: partial
loss of DNA methylation imprints,
defects in imprinted XCI,
embryonic lethality at
mid-gestation
Howell et al.
(
2001 ); McGraw
et al. (
2013 )
Maternal
and zygotic Dnmt3a
-/-
Complete loss of paternal and
maternal methylation imprints,
embryonic lethality at
mid-gestation
Hirasawa et al.
(
2008 )
Dnmt3a Dnmt3a -/-
Gut malfunction, spermatogenesis
defects, death at ~4 weeks of age
Okano et al.
(
1999 )
Dnmt3a -/-
in PGCs Failure to establish maternal and
paternal methylation imprints,
spermatogenesis defects
Kaneda et al.
(
2004 )
Dnmt3b Dnmt3b -/-
Hypomethylation of minor satellite
DNA, neural tube defects,
embryonic lethality at E14.5–18.5
Okano et al.
(
1999 )
Dnmt3a -/- ,
Dnmt3b -/-
Failure to initiate de novo
methylation after implantation,
developmental arrest at E8.5
Okano et al.
(
1999 )
Dnmt3L Dnmt3L -/-
Failure to establish maternal and
paternal methylation imprints,
spermatogenesis defects
Bourc’his et al.
(
2001 ); Hata et al.
(
2002 )
J. Dan and T. Chen
143
References
Aapola U, Kawasaki K, Scott HS, Ollila J, Vihinen M, Heino M, Shintani A, Minoshima S, Krohn
K, Antonarakis SE, et al. Isolation and initial characterization of a novel zinc fi nger gene,
DNMT3L, on 21q22.3, related to the cytosine-5-methyltransferase 3 gene family. Genomics.
2000;65:293–8.
Aapola U, Lyle R, Krohn K, Antonarakis SE, Peterson P. Isolation and initial characterization of
the mouse Dnmt3l gene. Cytogenet Cell Genet. 2001;92:122–6.
Arita K, Ariyoshi M, Tochio H, Nakamura Y, Shirakawa M. Recognition of hemi-methylated DNA
by the SRA protein UHRF1 by a base-fl ipping mechanism. Nature. 2008;455:818–21.
Avvakumov GV, Walker JR, Xue S, Li Y, Duan S, Bronner C, Arrowsmith CH, Dhe-Paganon
S. Structural basis for recognition of hemi-methylated DNA by the SRA domain of human
UHRF1. Nature. 2008;455:822–5.
Barlow DP, Stoger R, Herrmann BG, Saito K, Schweifer N. The mouse insulin-like growth factor
type-2 receptor is imprinted and closely linked to the Tme locus. Nature. 1991;349:84–7.
Bartolomei MS, Zemel S, Tilghman SM. Parental imprinting of the mouse H19 gene. Nature.
1991;351:153–5.
Barton SC, Surani MA, Norris ML. Role of paternal and maternal genomes in mouse development.
Nature. 1984;311:374–6.
Baubec T, Colombo DF, Wirbelauer C, Schmidt J, Burger L, Krebs AR, Akalin A, Schubeler
D. Genomic profi ling of DNA methyltransferases reveals a role for DNMT3B in genic meth-
ylation. Nature. 2015;520:243–7.
Beard C, Li E, Jaenisch R. Loss of methylation activates Xist in somatic but not in embryonic cells.
Genes Dev. 1995;9:2325–34.
Bestor T, Laudano A, Mattaliano R, Ingram V. Cloning and sequencing of a cDNA encoding DNA
methyltransferase of mouse cells. The carboxyl-terminal domain of the mammalian enzymes is
related to bacterial restriction methyltransferases. J Mol Biol. 1988;203:971–83.
Bestor TH. Activation of mammalian DNA methyltransferase by cleavage of a Zn binding regula-
tory domain. EMBO J. 1992;11:2611–7.
Bian C, Yu X. PGC7 suppresses TET3 for protecting DNA methylation. Nucleic Acids Res.
2014;42:2893–905.
Borgel J, Guibert S, Li Y, Chiba H, Schubeler D, Sasaki H, Forne T, Weber M. Targets and dynam-
ics of promoter DNA methylation during early mouse development. Nat Genet. 2010;42:
1093–100.
Bostick M, Kim JK, Esteve PO, Clark A, Pradhan S, Jacobsen SE. UHRF1 plays a role in main-
taining DNA methylation in mammalian cells. Science. 2007;317:1760–4.
Bourc’his D, Bestor TH. Meiotic catastrophe and retrotransposon reactivation in male germ cells
lacking Dnmt3L. Nature. 2004;431:96–9.
Bourc’his D, Xu GL, Lin CS, Bollman B, Bestor TH. Dnmt3L and the establishment of maternal
genomic imprints. Science. 2001;294:2536–9.
Brown CJ, Willard HF. The human X-inactivation centre is not required for maintenance of
X-chromosome inactivation. Nature. 1994;368:154–6.
Butler MG. Genomic imprinting disorders in humans: a mini-review. J Assist Reprod Genet.
2009;26:477–86.
Carrel L, Willard HF. X-inactivation profi le reveals extensive variability in X-linked gene expres-
sion in females. Nature. 2005;434:400–4.
Cattanach BM, Kirk M. Differential activity of maternally and paternally derived chromosome
regions in mice. Nature. 1985;315:496–8.
Chaumeil J, Okamoto I, Heard E. X-chromosome inactivation in mouse embryonic stem cells:
analysis of histone modifi cations and transcriptional activity using immunofl uorescence and
FISH. Methods Enzymol. 2004;376:405–19.
Chen T, Hevi S, Gay F, Tsujimoto N, He T, Zhang B, Ueda Y, Li E. Complete inactivation of
DNMT1 leads to mitotic catastrophe in human cancer cells. Nat Genet. 2007;39:391–6.
Genetic Studies on Mammalian DNA Methyltransferases
144
Chen T, Tsujimoto N, Li E. The PWWP domain of Dnmt3a and Dnmt3b is required for directing
DNA methylation to the major satellite repeats at pericentric heterochromatin. Mol Cell Biol.
2004;24:9048–58.
Chen T, Ueda Y, Dodge JE, Wang Z, Li E. Establishment and maintenance of genomic methyla-
tion patterns in mouse embryonic stem cells by Dnmt3a and Dnmt3b. Mol Cell Biol. 2003;23:
5594–605.
Chen T, Ueda Y, Xie S, Li E. A novel Dnmt3a isoform produced from an alternative promoter
localizes to euchromatin and its expression correlates with active de novo methylation. J Biol
Chem. 2002;277:38746–54.
Cheng J, Yang Y, Fang J, Xiao J, Zhu T, Chen F, Wang P, Li Z, Yang H, Xu Y. Structural insight
into coordinated recognition of trimethylated histone H3 lysine 9 (H3K9me3) by the plant
homeodomain (PHD) and tandem tudor domain (TTD) of UHRF1 (ubiquitin-like, containing
PHD and RING fi nger domains, 1) protein. J Biol Chem. 2013;288:1329–39.
Chotalia M, Smallwood SA, Ruf N, Dawson C, Lucifero D, Frontera M, James K, Dean W, Kelsey
G. Transcription is required for establishment of germline methylation marks at imprinted
genes. Genes Dev. 2009;23:105–17.
Chuang LS, Ian HI, Koh TW, Ng HH, Xu G, Li BF. Human DNA-(cytosine-5) methyltransferase-
PCNA complex as a target for p21WAF1. Science. 1997;277:1996–2000.
Ciccone DN, Su H, Hevi S, Gay F, Lei H, Bajko J, Xu G, Li E, Chen T. KDM1B is a histone H3K4
demethylase required to establish maternal genomic imprints. Nature. 2009;461:415–8.
Cifuentes-Rojas C, Hernandez AJ, Sarma K, Lee JT. Regulatory interactions between RNA and
polycomb repressive complex 2. Mol Cell. 2014;55:171–85.
Clemson CM, McNeil JA, Willard HF, Lawrence JB. XIST RNA paints the inactive X chromo-
some at interphase: evidence for a novel RNA involved in nuclear/chromosome structure.
J Biol Chem. 1996;132:259–75.
Csankovszki G, Panning B, Bates B, Pehrson JR, Jaenisch R. Conditional deletion of Xist dis-
rupts histone macroH2A localization but not maintenance of X inactivation. Nat Genet. 1999;
22:323–4.
da Rocha ST, Boeva V, Escamilla-Del-Arenal M, Ancelin K, Granier C, Matias NR, Sanulli S,
Chow J, Schulz E, Picard C, et al. Jarid2 is implicated in the initial xist-induced targeting of
PRC2 to the inactive X chromosome. Mol Cell. 2014;53:301–16.
DeChiara TM, Robertson EJ, Efstratiadis A. Parental imprinting of the mouse insulin-like growth
factor II gene. Cell. 1991;64:849–59.
Dhayalan A, Rajavelu A, Rathert P, Tamas R, Jurkowska RZ, Ragozin S, Jeltsch A. The Dnmt3a
PWWP domain reads histone 3 lysine 36 trimethylation and guides DNA methylation. J Biol
Chem. 2010;285:26114–20.
Ding F, Chaillet JR. In vivo stabilization of the Dnmt1 (cytosine-5)- methyltransferase protein.
Proc Natl Acad Sci U S A. 2002;99:14861–6.
Dodge JE, Okano M, Dick F, Tsujimoto N, Chen T, Wang S, Ueda Y, Dyson N, Li E. Inactivation
of Dnmt3b in mouse embryonic fi broblasts results in DNA hypomethylation, chromosomal
instability, and spontaneous immortalization. J Biol Chem. 2005;280:17986–91.
Du Z, Song J, Wang Y, Zhao Y, Guda K, Yang S, Kao HY, Xu Y, Willis J, Markowitz SD, et al.
DNMT1 stability is regulated by proteins coordinating deubiquitination and acetylation-driven
ubiquitination. Sci Signal. 2010;3:ra80.
Easwaran HP, Schermelleh L, Leonhardt H, Cardoso MC. Replication-independent chromatin
loading of Dnmt1 during G2 and M phases. EMBO Rep. 2004;5:1181–6.
Feng J, Zhou Y, Campbell SL, Le T, Li E, Sweatt JD, Silva AJ, Fan G. Dnmt1 and Dnmt3a main-
tain DNA methylation and regulate synaptic function in adult forebrain neurons. Nat Neurosci.
2010;13:423–30.
Frohlich LF, Mrakovcic M, Steinborn R, Chung UI, Bastepe M, Juppner H. Targeted deletion of
the Nesp55 DMR defi nes another Gnas imprinting control region and provides a mouse model
of autosomal dominant PHP-Ib. Proc Natl Acad Sci U S A. 2010;107:9275–80.
Frontera M, Dickins B, Plagge A, Kelsey G. Imprinted genes, postnatal adaptations and enduring
effects on energy homeostasis. Adv Exp Med Biol. 2008;626:41–61.
J. Dan and T. Chen
145
Gendrel AV, Heard E. Noncoding RNAs and epigenetic mechanisms during X-chromosome inac-
tivation. Annu Rev Cell Dev Biol. 2014;30:561–80.
Goll MG, Kirpekar F, Maggert KA, Yoder JA, Hsieh CL, Zhang X, Golic KG, Jacobsen SE,
Bestor TH. Methylation of tRNAAsp by the DNA methyltransferase homolog Dnmt2. Science.
2006;311:395–8.
Gu TP, Guo F, Yang H, Wu HP, Xu GF, Liu W, Xie ZG, Shi L, He X, Jin SG, et al. The role of
Tet3 DNA dioxygenase in epigenetic reprogramming by oocytes. Nature. 2011;477:606–10.
Guibert S, Forne T, Weber M. Global profi ling of DNA methylation erasure in mouse primordial
germ cells. Genome Res. 2012;22:633–41.
Haaf T. The effects of 5-azacytidine and 5-azadeoxycytidine on chromosome structure and function:
implications for methylation-associated cellular processes. Pharmacol Ther. 1995;65:19–46.
Hackett JA, Sengupta R, Zylicz JJ, Murakami K, Lee C, Down TA, Surani MA. Germline DNA
demethylation dynamics and imprint erasure through 5-hydroxymethylcytosine. Science.
2013;339:448–52.
Hamidi T, Singh AK, Chen T. Genetic alterations of DNA methylation machinery in human dis-
eases. Epigenomics. 2015;7:247–65.
Hashimoto H, Horton JR, Zhang X, Bostick M, Jacobsen SE, Cheng X. The SRA domain of
UHRF1 fl ips 5-methylcytosine out of the DNA helix. Nature. 2008;455:826–9.
Hata K, Okano M, Lei H, Li E. Dnmt3L cooperates with the Dnmt3 family of de novo DNA
methyltransferases to establish maternal imprints in mice. Development. 2002;129:1983–93.
He YF, Li BZ, Li Z, Liu P, Wang Y, Tang Q, Ding J, Jia Y, Chen Z, Li L, et al. Tet-mediated formation
of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science. 2011;333:1303–7.
Hirasawa R, Chiba H, Kaneda M, Tajima S, Li E, Jaenisch R, Sasaki H. Maternal and zygotic
Dnmt1 are necessary and suffi cient for the maintenance of DNA methylation imprints during
preimplantation development. Genes Dev. 2008;22:1607–16.
Hirasawa R, Sasaki H. Dynamic transition of Dnmt3b expression in mouse pre- and early post-
implantation embryos. Gene Expr Patterns. 2009;9:27–30.
Holliday R, Pugh JE. DNA modifi cation mechanisms and gene activity during development.
Science. 1975;187:226–32.
Howell CY, Bestor TH, Ding F, Latham KE, Mertineit C, Trasler JM, Chaillet JR. Genomic imprint-
ing disrupted by a maternal effect mutation in the Dnmt1 gene. Cell. 2001;104:829–38.
Inoue A, Shen L, Dai Q, He C, Zhang Y. Generation and replication-dependent dilution of 5fC and
5caC during mouse preimplantation development. Cell Res. 2011;21:1670–6.
Inoue A, Zhang Y. Replication-dependent loss of 5-hydroxymethylcytosine in mouse preimplanta-
tion embryos. Science. 2011;334:194.
Ito S, Shen L, Dai Q, Wu SC, Collins LB, Swenberg JA, He C, Zhang Y. Tet proteins can convert
5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science. 2011;333:1300–3.
Jackson-Grusby L, Beard C, Possemato R, Tudor M, Fambrough D, Csankovszki G, Dausman J,
Lee P, Wilson C, Lander E, et al. Loss of genomic methylation causes p53-dependent apoptosis
and epigenetic deregulation. Nat Genet. 2001;27:31–9.
Jelinic P, Shaw P. Loss of imprinting and cancer. J Pathol. 2007;211:261–8.
Jia D, Jurkowska RZ, Zhang X, Jeltsch A, Cheng X. Structure of Dnmt3a bound to Dnmt3L sug-
gests a model for de novo DNA methylation. Nature. 2007;449:248–51.
Jiang J, Jing Y, Cost GJ, Chiang JC, Kolpa HJ, Cotton AM, Carone DM, Carone BR, Shivak DA,
Guschin DY, et al. Translating dosage compensation to trisomy 21. Nature. 2013;500:296–300.
Kagiwada S, Kurimoto K, Hirota T, Yamaji M, Saitou M. Replication-coupled passive DNA
demethylation for the erasure of genome imprints in mice. EMBO J. 2013;32:340–53.
Kaneda M, Okano M, Hata K, Sado T, Tsujimoto N, Li E, Sasaki H. Essential role for de novo
DNA methyltransferase Dnmt3a in paternal and maternal imprinting. Nature. 2004;429:900–3.
Kim SJ, Zhao H, Hardikar S, Singh AK, Goodell MA, Chen T. A DNMT3A mutation common in
AML exhibits dominant-negative effects in murine ES cells. Blood. 2013;122:4086–9.
Kobayashi H, Sakurai T, Imai M, Takahashi N, Fukuda A, Yayoi O, Sato S, Nakabayashi K, Hata
K, Sotomaru Y, et al. Contribution of intragenic DNA methylation in mouse gametic DNA
methylomes to establish oocyte-specifi c heritable marks. PLoS Genet. 2012;8:e1002440.
Genetic Studies on Mammalian DNA Methyltransferases
146
Kobayashi H, Sakurai T, Miura F, Imai M, Mochiduki K, Yanagisawa E, Sakashita A, Wakai T,
Suzuki Y, Ito T, et al. High-resolution DNA methylome analysis of primordial germ cells iden-
tifi es gender-specifi c reprogramming in mice. Genome Res. 2013;23:616–27.
Kurihara Y, Kawamura Y, Uchijima Y, Amamo T, Kobayashi H, Asano T, Kurihara H. Maintenance
of genomic methylation patterns during preimplantation development requires the somatic
form of DNA methyltransferase 1. Dev Biol. 2008;313:335–46.
Kurimoto K, Yabuta Y, Ohinata Y, Shigeta M, Yamanaka K, Saitou M. Complex genome-wide
transcription dynamics orchestrated by Blimp1 for the specifi cation of the germ cell lineage in
mice. Genes Dev. 2008;22:1617–35.
Lee JT, Jaenisch R. Long-range cis effects of ectopic X-inactivation centres on a mouse autosome.
Nature. 1997;386:275–9.
Lee JT, Strauss WM, Dausman JA, Jaenisch R. A 450 kb transgene displays properties of the mam-
malian X-inactivation center. Cell. 1996;86:83–94.
Lee PP, Fitzpatrick DR, Beard C, Jessup HK, Lehar S, Makar KW, Perez-Melgosa M, Sweetser
MT, Schlissel MS, Nguyen S, et al. A critical role for Dnmt1 and DNA methylation in T cell
development, function, and survival. Immunity. 2001;15:763–74.
Lei H, Oh SP, Okano M, Juttermann R, Goss KA, Jaenisch R, Li E. De novo DNA cytosine meth-
yltransferase activities in mouse embryonic stem cells. Development. 1996;122:3195–205.
Leonhardt H, Page AW, Weier HU, Bestor TH. A targeting sequence directs DNA methyltransfer-
ase to sites of DNA replication in mammalian nuclei. Cell. 1992;71:865–73.
Li E, Beard C, Jaenisch R. Role for DNA methylation in genomic imprinting. Nature. 1993;
366:362–5.
Li E, Bestor TH, Jaenisch R. Targeted mutation of the DNA methyltransferase gene results in
embryonic lethality. Cell. 1992;69:915–26.
Li X, Ito M, Zhou F, Youngson N, Zuo X, Leder P, Ferguson-Smith AC. A maternal-zygotic effect
gene, Zfp57, maintains both maternal and paternal imprints. Dev Cell. 2008;15:547–57.
Liao J, Karnik R, Gu H, Ziller MJ, Clement K, Tsankov AM, Akopian V, Gifford CA, Donaghey J,
Galonska C, et al. Targeted disruption of DNMT1, DNMT3A and DNMT3B in human embry-
onic stem cells. Nat Genet. 2015;47:469–78.
Liu X, Gao Q, Li P, Zhao Q, Zhang J, Li J, Koseki H, Wong J. UHRF1 targets DNMT1 for DNA
methylation through cooperative binding of hemi-methylated DNA and methylated H3K9. Nat
Commun. 2013;4:1563.
Liu YJ, Nakamura T, Nakano T. Essential role of DPPA3 for chromatin condensation in mouse
oocytogenesis. Biol Reprod. 2012;86:40.
Lorthongpanich C, Cheow LF, Balu S, Quake SR, Knowles BB, Burkholder WF, Solter D,
Messerschmidt DM. Single-cell DNA-methylation analysis reveals epigenetic chimerism in
preimplantation embryos. Science. 2013;341:1110–2.
Ma P, Lin S, Bartolomei MS, Schultz RM. Metastasis tumor antigen 2 (MTA2) is involved in
proper imprinted expression of H19 and Peg3 during mouse preimplantation development. Biol
Reprod. 2010;83:1027–35.
Mackay DJ, Callaway JL, Marks SM, White HE, Acerini CL, Boonen SE, Dayanikli P, Firth HV,
Goodship JA, Haemers AP, et al. Hypomethylation of multiple imprinted loci in individuals
with transient neonatal diabetes is associated with mutations in ZFP57. Nat Genet. 2008;40:
949–51.
Marahrens Y, Panning B, Dausman J, Strauss W, Jaenisch R. Xist-defi cient mice are defective in
dosage compensation but not spermatogenesis. Genes Dev. 1997;11:156–66.
McGrath J, Solter D. Completion of mouse embryogenesis requires both the maternal and paternal
genomes. Cell. 1984;37:179–83.
McGraw S, Oakes CC, Martel J, Cirio MC, de Zeeuw P, Mak W, Plass C, Bartolomei MS, Chaillet
JR, Trasler JM. Loss of DNMT1o disrupts imprinted X chromosome inactivation and accentu-
ates placental defects in females. PLoS Genet. 2013;9:e1003873.
Mertineit C, Yoder JA, Taketo T, Laird DW, Trasler JM, Bestor TH. Sex-specifi c exons control
DNA methyltransferase in mammalian germ cells. Development. 1998;125:889–97.
J. Dan and T. Chen
147
Messerschmidt DM, de Vries W, Ito M, Solter D, Ferguson-Smith A, Knowles BB. Trim28
is required for epigenetic stability during mouse oocyte to embryo transition. Science.
2012;335:1499–502.
Mudbhary R, Hoshida Y, Chernyavskaya Y, Jacob V, Villanueva A, Fiel MI, Chen X, Kojima K,
Thung S, Bronson RT, et al. UHRF1 overexpression drives DNA hypomethylation and hepato-
cellular carcinoma. Cancer Cell. 2014;25:196–209.
Muto M, Kanari Y, Kubo E, Takabe T, Kurihara T, Fujimori A, Tatsumi K. Targeted disruption of
Np95 gene renders murine embryonic stem cells hypersensitive to DNA damaging agents and
DNA replication blocks. J Biol Chem. 2002;277:34549–55.
Nakamura T, Arai Y, Umehara H, Masuhara M, Kimura T, Taniguchi H, Sekimoto T, Ikawa M,
Yoneda Y, Okabe M, et al. PGC7/Stella protects against DNA demethylation in early embryo-
genesis. Nat Cell Biol. 2007;9:64–71.
Nakamura T, Liu YJ, Nakashima H, Umehara H, Inoue K, Matoba S, Tachibana M, Ogura A,
Shinkai Y, Nakano T. PGC7 binds histone H3K9me2 to protect against conversion of 5mC to
5hmC in early embryos. Nature. 2012;486:415–9.
Neri F, Krepelova A, Incarnato D, Maldotti M, Parlato C, Galvagni F, Matarese F, Stunnenberg
HG, Oliviero S. Dnmt3L antagonizes DNA methylation at bivalent promoters and favors DNA
methylation at gene bodies in ESCs. Cell. 2013;155:121–34.
Nielsen AL, Ortiz JA, You J, Oulad-Abdelghani M, Khechumian R, Gansmuller A, Chambon P,
Losson R. Interaction with members of the heterochromatin protein 1 (HP1) family and histone
deacetylation are differentially involved in transcriptional silencing by members of the TIF1
family. EMBO J. 1999;18:6385–95.
Nishiyama A, Yamaguchi L, Sharif J, Johmura Y, Kawamura T, Nakanishi K, Shimamura S, Arita
K, Kodama T, Ishikawa F, et al. Uhrf1-dependent H3K23 ubiquitylation couples maintenance
DNA methylation and replication. Nature. 2013;502:249–53.
Noh KM, Wang H, Kim HR, Wenderski W, Fang F, Li CH, Dewell S, Hughes SH, Melnick AM,
Patel DJ, et al. Engineering of a histone-recognition domain in Dnmt3a alters the epigenetic
landscape and phenotypic features of mouse ESCs. Mol Cell. 2015;59:89–103.
Norris DP, Patel D, Kay GF, Penny GD, Brockdorff N, Sheardown SA, Rastan S. Evidence
that random and imprinted Xist expression is controlled by preemptive methylation. Cell.
1994;77:41–51.
Okano M, Bell DW, Haber DA, Li E. DNA methyltransferases Dnmt3a and Dnmt3b are essential
for de novo methylation and mammalian development. Cell. 1999;99:247–57.
Okano M, Xie S, Li E. Cloning and characterization of a family of novel mammalian DNA (cyto-
sine- 5) methyltransferases. Nat Genet. 1998a;19:219–20.
Okano M, Xie S, Li E. Dnmt2 is not required for de novo and maintenance methylation of viral
DNA in embryonic stem cells. Nucleic Acids Res. 1998b;26:2536–40.
Ooi SK, Qiu C, Bernstein E, Li K, Jia D, Yang Z, Erdjument-Bromage H, Tempst P, Lin SP, Allis
CD, et al. DNMT3L connects unmethylated lysine 4 of histone H3 to de novo methylation of
DNA. Nature. 2007;448:714–7.
Ostler KR, Yang Q, Looney TJ, Zhang L, Vasanthakumar A, Tian Y, Kocherginsky M, Raimondi
SL, DeMaio JG, Salwen HR, et al. Truncated DNMT3B isoform DNMT3B7 suppresses
growth, induces differentiation, and alters DNA methylation in human neuroblastoma. Cancer
Res. 2012;72:4714–23.
Otani J, Nankumo T, Arita K, Inamoto S, Ariyoshi M, Shirakawa M. Structural basis for recogni-
tion of H3K4 methylation status by the DNA methyltransferase 3A ATRX-DNMT3-DNMT3L
domain. EMBO Rep. 2009;10:1235–41.
Panning B, Jaenisch R. DNA hypomethylation can activate Xist expression and silence X-linked
genes. Genes Dev. 1996;10:1991–2002.
Payer B, Lee JT. X chromosome dosage compensation: how mammals keep the balance. Annu Rev
Genet. 2008;42:733–72.
Penny GD, Kay GF, Sheardown SA, Rastan S, Brockdorff N. Requirement for Xist in X chromo-
some inactivation. Nature. 1996;379:131–7.
Genetic Studies on Mammalian DNA Methyltransferases
148
Popp C, Dean W, Feng S, Cokus SJ, Andrews S, Pellegrini M, Jacobsen SE, Reik W. Genome-
wide erasure of DNA methylation in mouse primordial germ cells is affected by AID defi -
ciency. Nature. 2010;463:1101–5.
Pradhan S, Bacolla A, Wells RD, Roberts RJ. Recombinant human DNA (cytosine-5) methyltrans-
ferase. I. Expression, purifi cation, and comparison of de novo and maintenance methylation.
J Biol Chem. 1999;274:33002–10.
Qin W, Leonhardt H, Spada F. Usp7 and Uhrf1 control ubiquitination and stability of the mainte-
nance DNA methyltransferase Dnmt1. J Cell Biochem. 2011;112:439–44.
Qin W, Wolf P, Liu N, Link S, Smets M, Mastra FL, Forne I, Pichler G, Horl D, Fellinger K, et al.
DNA methylation requires a DNMT1 ubiquitin interacting motif (UIM) and histone ubiquitina-
tion. Cell Res. 2015;25:911–29.
Quenneville S, Verde G, Corsinotti A, Kapopoulou A, Jakobsson J, Offner S, Baglivo I, Pedone
PV, Grimaldi G, Riccio A, et al. In embryonic stem cells, ZFP57/KAP1 recognize a methylated
hexanucleotide to affect chromatin and DNA methylation of imprinting control regions. Mol
Cell. 2011;44:361–72.
Ratnam S, Mertineit C, Ding F, Howell CY, Clarke HJ, Bestor TH, Chaillet JR, Trasler
JM. Dynamics of Dnmt1 methyltransferase expression and intracellular localization during
oogenesis and preimplantation development. Dev Biol. 2002;245:304–14.
Reese KJ, Lin S, Verona RI, Schultz RM, Bartolomei MS. Maintenance of paternal methylation
and repression of the imprinted H19 gene requires MBD3. PLoS Genet. 2007;3:e137.
Reik W, Walter J. Genomic imprinting: parental infl uence on the genome. Nat Rev Genet.
2001;2:21–32.
Rhee I, Jair KW, Yen RW, Lengauer C, Herman JG, Kinzler KW, Vogelstein B, Baylin SB,
Schuebel KE. CpG methylation is maintained in human cancer cells lacking DNMT1. Nature.
2000;404:1003–7.
Riggs AD. X inactivation, differentiation, and DNA methylation. Cytogenet Cell Genet. 1975;
14:9–25.
Rothbart SB, Dickson BM, Ong MS, Krajewski K, Houliston S, Kireev DB, Arrowsmith CH, Strahl
BD. Multivalent histone engagement by the linked tandem Tudor and PHD domains of UHRF1
is required for the epigenetic inheritance of DNA methylation. Genes Dev. 2013;27:1288–98.
Rothbart SB, Krajewski K, Nady N, Tempel W, Xue S, Badeaux AI, Barsyte-Lovejoy D, Martinez
JY, Bedford MT, Fuchs SM, et al. Association of UHRF1 with methylated H3K9 directs the
maintenance of DNA methylation. Nat Struct Mol Biol. 2012;19:1155–60.
Rottach A, Frauer C, Pichler G, Bonapace IM, Spada F, Leonhardt H. The multi-domain protein Np95
connects DNA methylation and histone modifi cation. Nucleic Acids Res. 2010;38:1796–804.
Rouleau J, Tanigawa G, Szyf M. The mouse DNA methyltransferase 5-region. A unique house-
keeping gene promoter. J Biol Chem. 1992;267:7368–77.
Ryan RF, Schultz DC, Ayyanathan K, Singh PB, Friedman JR, Fredericks WJ, Rauscher 3rd
FJ. KAP-1 corepressor protein interacts and colocalizes with heterochromatic and euchro-
matic HP1 proteins: a potential role for Kruppel-associated box-zinc fi nger proteins in
heterochromatin- mediated gene silencing. Mol Cell Biol. 1999;19:4366–78.
Sado T, Fenner MH, Tan SS, Tam P, Shioda T, Li E. X inactivation in the mouse embryo defi cient
for Dnmt1: distinct effect of hypomethylation on imprinted and random X inactivation. Dev
Biol. 2000;225:294–303.
Sado T, Okano M, Li E, Sasaki H. De novo DNA methylation is dispensable for the initiation and
propagation of X chromosome inactivation. Development. 2004;131:975–82.
Schneider K, Fuchs C, Dobay A, Rottach A, Qin W, Wolf P, Alvarez-Castro JM, Nalaskowski
MM, Kremmer E, Schmid V, et al. Dissection of cell cycle-dependent dynamics of Dnmt1 by
FRAP and diffusion-coupled modeling. Nucleic Acids Res. 2013;41:4860–76.
Schultz DC, Ayyanathan K, Negorev D, Maul GG, Rauscher 3rd FJ. SETDB1: a novel KAP-1-
associated histone H3, lysine 9-specifi c methyltransferase that contributes to HP1-mediated
silencing of euchromatic genes by KRAB zinc-fi nger proteins. Genes Dev. 2002;16:919–32.
J. Dan and T. Chen
149
Schultz DC, Friedman JR, Rauscher 3rd FJ. Targeting histone deacetylase complexes via KRAB-
zinc fi nger proteins: the PHD and bromodomains of KAP-1 form a cooperative unit that recruits
a novel isoform of the Mi-2alpha subunit of NuRD. Genes Dev. 2001;15:428–43.
Seisenberger S, Andrews S, Krueger F, Arand J, Walter J, Santos F, Popp C, Thienpont B, Dean W,
Reik W. The dynamics of genome-wide DNA methylation reprogramming in mouse primordial
germ cells. Mol Cell. 2012;48:849–62.
Shah MY, Vasanthakumar A, Barnes NY, Figueroa ME, Kamp A, Hendrick C, Ostler KR, Davis
EM, Lin S, Anastasi J, et al. DNMT3B7, a truncated DNMT3B isoform expressed in human
tumors, disrupts embryonic development and accelerates lymphomagenesis. Cancer Res.
2010;70:5840–50.
Sharif J, Muto M, Takebayashi S, Suetake I, Iwamatsu A, Endo TA, Shinga J, Mizutani-Koseki Y,
Toyoda T, Okamura K, et al. The SRA protein Np95 mediates epigenetic inheritance by recruit-
ing Dnmt1 to methylated DNA. Nature. 2007;450:908–12.
Smallwood SA, Tomizawa S, Krueger F, Ruf N, Carli N, Segonds-Pichon A, Sato S, Hata K,
Andrews SR, Kelsey G. Dynamic CpG island methylation landscape in oocytes and preimplan-
tation embryos. Nat Genet. 2011;43:811–4.
Smith EY, Futtner CR, Chamberlain SJ, Johnstone KA, Resnick JL. Transcription is required to
establish maternal imprinting at the Prader-Willi syndrome and Angelman syndrome locus.
PLoS Genet. 2011;7:e1002422.
Smith ZD, Chan MM, Mikkelsen TS, Gu H, Gnirke A, Regev A, Meissner A. A unique regulatory
phase of DNA methylation in the early mammalian embryo. Nature. 2012;484:339–44.
Splinter E, de Wit E, Nora EP, Klous P, van de Werken HJ, Zhu Y, Kaaij LJ, van Ijcken W, Gribnau
J, Heard E, et al. The inactive X chromosome adopts a unique three-dimensional conformation
that is dependent on Xist RNA. Genes Dev. 2011;25:1371–83.
Suetake I, Shinozaki F, Miyagawa J, Takeshima H, Tajima S. DNMT3L stimulates the DNA
methylation activity of Dnmt3a and Dnmt3b through a direct interaction. J Biol Chem.
2004;279:27816–23.
Surani MA, Barton SC, Norris ML. Development of reconstituted mouse eggs suggests imprinting
of the genome during gametogenesis. Nature. 1984;308:548–50.
Takebayashi S, Tamura T, Matsuoka C, Okano M. Major and essential role for the DNA methyla-
tion mark in mouse embryogenesis and stable association of DNMT1 with newly replicated
regions. Mol Cell Biol. 2007;27:8243–58.
Takeshita K, Suetake I, Yamashita E, Suga M, Narita H, Nakagawa A, Tajima S. Structural insight
into maintenance methylation by mouse DNA methyltransferase 1 (Dnmt1). Proc Natl Acad
Sci U S A. 2011;108:9055–9.
Tesar PJ, Chenoweth JG, Brook FA, Davies TJ, Evans EP, Mack DL, Gardner RL, McKay
RD. New cell lines from mouse epiblast share defi ning features with human embryonic stem
cells. Nature. 2007;448:196–9.
Tomizawa S, Sasaki H. Genomic imprinting and its relevance to congenital disease, infertility,
molar pregnancy and induced pluripotent stem cell. J Hum Genet. 2012;57:84–91.
Tsumura A, Hayakawa T, Kumaki Y, Takebayashi S, Sakaue M, Matsuoka C, Shimotohno K,
Ishikawa F, Li E, Ueda HR, et al. Maintenance of self-renewal ability of mouse embryonic
stem cells in the absence of DNA methyltransferases Dnmt1, Dnmt3a and Dnmt3b. Genes
Cells. 2006;11:805–14.
Tucker KL, Talbot D, Lee MA, Leonhardt H, Jaenisch R. Complementation of methylation defi -
ciency in embryonic stem cells by a DNA methyltransferase minigene. Proc Natl Acad Sci U
S A. 1996;93:12920–5.
Van den Wyngaert I, Sprengel J, Kass SU, Luyten WH. Cloning and analysis of a novel human
putative DNA methyltransferase. FEBS Lett. 1998;426:283–9.
Vlachogiannis G, Niederhuth CE, Tuna S, Stathopoulou A, Viiri K, de Rooij DG, Jenner RG,
Schmitz RJ, Ooi SK. The Dnmt3L ADD domain controls cytosine methylation establishment
during spermatogenesis. Cell Rep. 2015;10:944–56.
Walter J, Paulsen M. Imprinting and disease. Semin Cell Dev Biol. 2003;14:101–10.
Genetic Studies on Mammalian DNA Methyltransferases
150
Weber M, Hellmann I, Stadler MB, Ramos L, Paabo S, Rebhan M, Schubeler D. Distribution,
silencing potential and evolutionary impact of promoter DNA methylation in the human
genome. Nat Genet. 2007;39:457–66.
Williamson CM, Ball ST, Dawson C, Mehta S, Beechey CV, Fray M, Teboul L, Dear TN, Kelsey
G, Peters J. Uncoupling antisense-mediated silencing and DNA methylation in the imprinted
Gnas cluster. PLoS Genet. 2011;7:e1001347.
Wossidlo M, Nakamura T, Lepikhov K, Marques CJ, Zakhartchenko V, Boiani M, Arand J, Nakano
T, Reik W, Walter J. 5-Hydroxymethylcytosine in the mammalian zygote is linked with epigen-
etic reprogramming. Nat Commun. 2011;2:241.
Wutz A. Gene silencing in X-chromosome inactivation: advances in understanding facultative het-
erochromatin formation. Nat Rev Genet. 2011;12:542–53.
Yamaguchi S, Hong K, Liu R, Inoue A, Shen L, Zhang K, Zhang Y. Dynamics of 5-methylcytosine
and 5-hydroxymethylcytosine during germ cell reprogramming. Cell Res. 2013;23:329–39.
Yang F, Babak T, Shendure J, Disteche CM. Global survey of escape from X inactivation by RNA-
sequencing in mouse. Genome Res. 2010;20:614–22.
Yoder JA, Bestor TH. A candidate mammalian DNA methyltransferase related to pmt1p of fi ssion
yeast. Hum Mol Genet. 1998;7:279–84.
Zhang Y, Jurkowska R, Soeroes S, Rajavelu A, Dhayalan A, Bock I, Rathert P, Brandt O, Reinhardt
R, Fischle W, et al. Chromatin methylation activity of Dnmt3a and Dnmt3a/3L is guided
by interaction of the ADD domain with the histone H3 tail. Nucleic Acids Res. 2010;38:
4246–53.
Zhao H, Chen T. Tet family of 5-methylcytosine dioxygenases in mammalian development. J Hum
Genet. 2013;58:421–7.
Zhao J, Sun BK, Erwin JA, Song JJ, Lee JT. Polycomb proteins targeted by a short repeat RNA to
the mouse X chromosome. Science. 2008;322:750–6.
Zuo X, Sheng J, Lau HT, McDonald CM, Andrade M, Cullen DE, Bell FT, Iacovino M, Kyba M,
Xu G, et al. Zinc fi nger protein ZFP57 requires its co-factor to recruit DNA methyltransferases
and maintains DNA methylation imprint in embryonic stem cells via its transcriptional repres-
sion domain. J Biol Chem. 2012;287:2107–18.
J. Dan and T. Chen
... Once the 5mC methyl group is attached, it can be physically located in the major groove of a DNA helix where it can be accessed by DNA methylation readers, the methyl-CpG-binding domain (MBD) protein family (Table 3) [474]. De novo methylation is described to occur by the DNMT3 family of proteins, namely DNMT3A/B, with DNA methylation maintenance being performed by DNMT1 [442,[475][476][477][478][479] (a detailed review of the DNMT family can be found in reference [480]). Despite these canonical roles, evidence has emerged that each protein is capable of having a compensatory function for the other DNMTs should the cell require it; i.e., DNMT1 can show de novo methylation function, while DNMT3A/B can show DNA maintenance function [481,482]. ...
... Following synthesis, DNMT1 is recruited to the daughter strand to restore the methylation pattern read on the parental strand [485]. If for any reason DNA methylation maintenance is inhibited at this stage, owing to mutation, molecular inhibition, or metabolite insufficiency, the integrity of the methylation signature can be diluted or lost if this persists for subsequent cellular divisions [473,480,487]. ...
Article
Full-text available
Cellular metabolism (or energetics) and epigenetics are tightly coupled cellular processes. It is arguable that of all the described cancer hallmarks, dysregulated cellular energetics and epigenetics are the most tightly coregulated. Cellular metabolic states regulate and drive epigenetic changes while also being capable of influencing, if not driving, epigenetic reprogramming. Conversely, epigenetic changes can drive altered and compensatory metabolic states. Cancer cells meticulously modify and control each of these two linked cellular processes in order to maintain their tumorigenic potential and capacity. This review aims to explore the interplay between these two processes and discuss how each affects the other, driving and enhancing tumorigenic states in certain contexts.
... This process is predominantly managed by three DNA methyltransferases (DNMTs), maintenance DNA methyltransferase (DNMT1), and de novo DNA methyltransferases (DNMT3A and DNMT3B), which transfer a methyl group to the fifth carbon of cytosine, forming 5-methylcytosine (5mC). [13][14][15][16] It is reported that there is a significantly higher methylation level in PE placenta than in the normal term placenta. Additionally, global CpG hypermethylation coupled with hypomethylation in promoters is observed in PE placenta. ...
Article
Full-text available
The etiology of preeclampsia (PE), a complex and multifactorial condition, remains incompletely understood. DNA methylation, which is primarily regulated by three DNA methyltransferases (DNMTs), DNMT1, DNMT3A, and DNMT3B, plays a vital role in early embryonic development and trophectoderm differentiation. Yet, how DNMTs modulate trophoblast fusion and PE development remains unclear. In this study, we found that the DNMTs expression was downregulated during trophoblast cells fusion. Downregulation of DNMTs was observed during the reconstruction of the denuded syncytiotrophoblast (STB) layer of placental explants. Additionally, overexpression of DNMTs inhibited trophoblast fusion. Conversely, treatment with the DNA methylation inhibitor 5‐aza‐CdR decreased the expression of DNMTs and promoted trophoblast fusion. A combined analysis of DNA methylation data and gene transcriptome data obtained from the primary cytotrophoblasts (CTBs) fusion process identified 104 potential methylation‐regulated differentially expressed genes (MeDEGs) with upregulated expression due to DNA demethylation, including CD59, TNFAIP3, SDC1, and CDK6. The transcription regulation region (TRR) of TNFAIP3 showed a hypomethylation with induction of 5‐aza‐CdR, which facilitated CREB recruitment and thereby participated in regulating trophoblast fusion. More importantly, clinical correlation analysis of PE showed that the abnormal increase in DNMTs may be involved in the development of PE. This study identified placental DNA methylation‐regulated genes that may contribute to PE, offering a novel perspective on the role of epigenetics in trophoblast fusion and its implication in PE development.
... DNA methylation primarily occurs at CpG sites, cytosines preceding guanine nucleotides, and is catalyzed by DNA methyltransferases (DNMTs), including DNMT1, DNMT3A, and DNMT3B . DNMT1 maintains methylation patterns during cell division, while DNMT3A and DNMT3B are involved in de novo methylation during development (Dan and Chen, 2016;. Notably, DNMT3A plays a pivotal role in several biological processes, particularly in reproductive biology (Dura et al., 2022). ...
Article
Full-text available
DNMT3A participates in de novo methylation, yet its impact on the proliferation of testicular Sertoli cells remains unclear. Development-specific methylation has been proven to be associated with cellular development. Therefore, in this study, we simulated DNMT3A expression pattern during testicular development by DNMT3A interference. Then, RRBS and RNA-seq were used to decipher DNMT3A regulatory mechanisms on Sertoli cell proliferation. Immunofluorescence staining revealed the expression of DNMT3A in the Sertoli cells of the prepubertal testis. DNMT3A was demonstrated to inhibit the cell cycle and proliferation of Sertoli cells, while promoting cell apoptosis. After transfected with DNMT3A interference, a total of 560 DEGs and 2,091 DMGs produced by DNMT3A interference were identified between two treated groups, respectively. Integrating the results from RRBS and RNA-seq, the overlapping genes between DMGs and DEGs were found to be enriched in the Gene Ontology (GO) terms related to cellular development and the Apelin signaling pathway. The present study demonstrated the impact of DNMT3A on the proliferation of porcine testicular Sertoli cells, suggesting that DNMT3A primarily acts through the Apelin signaling pathway. These findings provide valuable insights into how DNMT3A influences testicular development and health, offering new perspectives.
... In mammalian development, DNA meth ylation is regulated by three conserved DNA methyltransferase (DNMT) enzymes: DNMT3a and DNMT3b, which typically act as de novo methylators, and DNMT1, which is primarily responsible for maintenance methylation. Severe malformations observed in Dnmt1 and Dnmt3b knockout mice demonstrated the developmental require ment of DNMT activity, but early embryonic lethality left the role of DNMTs and DNA methylation in orofacial morphogenesis unclear (13,(27)(28)(29). ...
Article
Full-text available
Orofacial clefts of the lip and palate are widely recognized to result from complex gene–environment interactions, but inadequate understanding of environmental risk factors has stymied development of prevention strategies. We interrogated the role of DNA methylation, an environmentally malleable epigenetic mechanism, in orofacial development. Expression of the key DNA methyltransferase enzyme DNMT1 was detected throughout palate morphogenesis in the epithelium and underlying cranial neural crest cell (cNCC) mesenchyme, a highly proliferative multipotent stem cell population that forms orofacial connective tissue. Genetic and pharmacologic manipulations of DNMT activity were then applied to define the tissue- and timing-dependent requirement of DNA methylation in orofacial development. cNCC-specific Dnmt1 inactivation targeting initial palate outgrowth resulted in OFCs, while later targeting during palatal shelf elevation and elongation did not. Conditional Dnmt1 deletion reduced cNCC proliferation and subsequent differentiation trajectory, resulting in attenuated outgrowth of the palatal shelves and altered development of cNCC-derived skeletal elements. Finally, we found that the cellular mechanisms of cleft pathogenesis observed in vivo can be recapitulated by pharmacologically reducing DNA methylation in multipotent cNCCs cultured in vitro. These findings demonstrate that DNA methylation is a crucial epigenetic regulator of cNCC biology, define a critical period of development in which its disruption directly causes OFCs, and provide opportunities to identify environmental influences that contribute to OFC risk.
... Li Xu and Biao Peng contributed equally to this work. Methyltransferases are a distinct group of proteins described by the attendance of a structurally conserved S-adenosyl methionine (SAM) binding domain and methyltransferase-like domain (Dan and Chen 2016;Holoch and Moazed 2015;Horning et al. 2016). It has been proven that methylation can regulate gene transcription and influence chromatin organization without gene mutations (Edwards et al. 2017). ...
Article
Full-text available
Glioblastoma (GBM), a predominant central nervous system (CNS) malignancy, is correlated with high mortality and severe morbidity. Mammalian methyltransferase-like 7B (METTL7B) as a methyltransferase has been identified to participate in cancer progression. However, its function in GBM is elusive. Accordingly, we aimed to explore the effect of METTL7B on GBM. The expression of METTL7B and EGR2 in GBM patients and GBM cells were detected by qPCR, western blots and immunohistochemical staining. Cell viability was assessed by CCK-8 assays. Cell proliferation was determined by EdU, colony formation, and tumor sphere formation assays. METTL7B shRNA was injected into the Balb/c nude mice. The size and weight of isolated tumor was measured. And the expression levels of Ki67, METTL7B and EGR1 were examined by immunohistochemical staining. METTL7B was significantly elevated, while EGR1 was downregulated in clinical GBM tissues. METTL7B upregulation was associated with the low overall survival of GBM patients. Moreover, METTL7B depletion remarkably attenuated GBM cell proliferation. Mechanistically, METTL7B overexpression inhibited EGR1 expression in GBM cells. EGR1 knockdown rescued the inhibitory effect of METTL7B depletion on GBM cell proliferation. Meanwhile, METTL7B depletion arrested more GBM cells at the G0/G1, but fewer cells at the S phase, which EGR1 knockdown reversed these effects. Furthermore, tumorigenicity analysis revealed that METTL7B promotes tumor growth of GBM cells in vivo. METTL7B contributes to the malignant progression of GBM by inhibiting EGR1 expression. METTL7B and EGR1 may be utilized as the treatment targets for GBM therapy.
... Methyltransferases contains a S-Adenosyl Methionine (SAM) binding domain, known as a family with a structurally conserved methyltransferase similar domains (6)(7)(8)(9). Research studies have shown that methylation could influence chromatin organization and directly regulate transcription of gene, but modulating the mutations in the gene itself (8,10). In addition, studies have demonstrated that methyltransferases could influence the progression of metabolic diseases, genetic diseases, and cancers (11)(12)(13). ...
Article
Full-text available
Methyltransferase-like 18 (METTL18), a METTL family member, is abundant in hepatocellular carcinoma (HCC). Studies have indicated the METTL family could regulate the progress of diverse malignancies while the role of METTL18 in HCC remains unclear. Data of HCC patients were acquired from the cancer genome atlas (TCGA) and gene expression omnibus (GEO). The expression level of METTL18 in HCC patients was compared with normal liver tissues by Wilcoxon test. Then, the logistic analysis was used to estimate the correlation between METTL18 and clinicopathological factors. Besides, Gene Ontology (GO), Gene Set Enrichment Analysis (GSEA), and single-sample Gene Set Enrichment Analysis (ssGSEA) were used to explore relevant functions and quantify the degree of immune infiltration for METTL18. Univariate and Multivariate Cox analyses and Kaplan–Meier analysis were used to estimate the association between METTL18 and prognosis. Besides, by cox multivariate analysis, a nomogram was conducted to forecast the influence of METTL18 on survival rates. METTL18-high was associated with Histologic grade, T stage, Pathologic stage, BMI, Adjacent hepatic tissue inflammation, AFP, Vascular invasion, and TP53 status (P < 0.05). HCC patients with METTL18-high had a poor Overall-Survival [OS; hazard ratio (HR): 1.87, P < 0.001), Disease-Specific Survival (DSS, HR: 1.76, P = 0.015), and Progression-Free Interval (PFI, HR: 1.51, P = 0.006). Multivariate analysis demonstrated that METTL18 was an independent factor for OS (HR: 2.093, P < 0.001), DSS (HR: 2.404, P = 0.015), and PFI (HR: 1.133, P = 0.006). Based on multivariate analysis, the calibration plots and C-indexes of nomograms showed an efficacious predictive effect for HCC patients. GSEA demonstrated that METTL18-high could activate G2M checkpoint, E2F targets, KRAS signaling pathway, and Mitotic Spindle. There was a positive association between the METTL18 and abundance of innate immunocytes (T helper 2 cells) and a negative relation to the abundance of adaptive immunocytes (Dendritic cells, Cytotoxic cells etc.). Finally, we uncovered knockdown of METTL18 significantly suppressed the proliferation, invasion, and migration of HCC cells in vitro. This research indicates that METTL18 could be a novel biomarker to evaluate HCC patients’ prognosis and an important regulator of immune responses in HCC.
... Das Resultat der Hypomethylierung ist das schwere Immundefizit- (Dan & Chen, 2016). ...
Thesis
Currently, the vulnerability-stress model, in the sense of a multifactorial explanatory model, is considered to be the most appropriate to represent the etiopathogenesis of anxiety disorders. Epigenetic mechanisms are understood as a bridge between genetic factors and environmental factors. This includes the methylation of specific DNA regions, which is mediated by DNA methyltransferases. These enzymes have rarely been the focus of psychiatric research in relation to anxiety disorders. Therefore, this work deals with selected single nucleotide polymorphisms of the DNMT3A and DNMT3B gene and investigates whether these SNPs and/or their haplotypes are associated panic disorder and/or with dimensional psychological characteristics, such as anxiety-related cognition or anxiety sensitivity. In summary, a significant or nominally significant association of two SNPs with anxiety-related characteristics such was shown. To better assess these associations, replications with sufficient test strength are required . Given the demonstrated association with PSWQ, investigation of another anxiety phenotype, Generalized Anxiety Disorder, is also sensible. As a further step, the functionality of the significantly associated SNPs should be performed. In addition, another DNMT, Dnmt1, is associated with fear conditioning, and the methylation patterns of the DNMTs themselves also appear to have an impact on the development of anxiety disorders. Therefore, an investigation of the DNMT1 gene and the methylation patterns of the DNMT genes are further reasonable steps to better understand a possible influence of DNMTs on the development of anxiety disorders and on anxiety-related psychological characteristics.
Article
Full-text available
One of the main challenges of current research on aging is to identify the complex epigenetic mechanisms involved in the acquisition of the cellular senescent phenotype. Despite some evidence suggested that epigenetic changes of DNA repetitive elements, including transposable elements (TE) sequences, are associated with replicative senescence of fibroblasts, data on different types of cells are scarce. We previously analysed genome-wide DNA methylation of young and replicative senescent human endothelial cells (HUVECs), highlighting increased levels of demethylated sequences in senescent cells. Here, we aligned the most significantly demethylated single CpG sites to the reference genome and annotated their localization inside TE sequences and found a significant hypomethylation of sequences belonging to the Long-Interspersed Element-1 (LINE-1 or L1) subfamilies L1M, L1P, and L1HS. To verify the hypothesis that L1 demethylation could be associated with increased transcription/activation of L1s and/or Alu elements (non-autonomous retroelements that usually depend on L1 sequences for reverse transcription and retrotransposition), we quantified the RNA expression levels of both L1 (generic L1 elements or site-specific L1PA2 on chromosome 14) and Alu elements in young and senescent HUVECs and human dermal fibroblasts (NHDFs). The RNA expression of Alu and L1 sequences was significantly increased in both senescent HUVECs and NHDFs, whereas the RNA transcript of L1PA2 on chromosome 14 was not significantly modulated in senescent cells. Moreover, we found an increased amount of TE DNA copies in the cytoplasm of senescent HUVECs and NHDFs. Our results support the hypothesis that TE, which are significantly increased in senescent cells, could be retrotranscribed to DNA sequences.
Article
Small B-cell lymphoma is the classification of B-cell chronic lymphoproliferative disorders that include chronic lymphocytic leukaemia/small lymphocytic lymphoma, follicular lymphoma, mantle cell lymphoma, marginal zone lymphoma, lymphoplasmacytic lymphoma/Waldenstrom macroglobulinemia. The clinical presentation is somewhat heterogeneous, and its occurrence and development mechanisms are not yet precise and may involve epigenetic changes. Epigenetic alterations mainly include DNA methylation, histone modification, and non-coding RNA, which are essential for genetic detection, early diagnosis, and assessment of treatment resistance in small B-cell lymphoma. As chronic lymphocytic leukemia/small lymphocytic lymphoma has already been reported in the literature, this article focuses on small B-cell lymphomas such as follicular lymphoma, mantle cell lymphoma, marginal zone lymphoma, and Waldenstrom macroglobulinemia. It discusses recent developments in epigenetic research to diagnose and treat this group of lymphomas. This review provides new ideas for the treatment and prognosis assessment of small B-cell lymphoma by exploring the connection between small B-cell lymphoma and epigenetics.
Article
Full-text available
Diet bioactive components, in the concept of nutrigenetics and nutrigenomics, consist of food constituents, which can transfer information from the external environment and influence gene expression in the cell and thus the function of the whole organism. It is crucial to regard food not only as the source of energy and basic nutriments, crucial for living and organism development, but also as the factor influencing health/disease, biochemical mechanisms, and activation of biochemical pathways. Bioactive components of the diet regulate gene expression through changes in the chromatin structure (including DNA methylation and histone modification), non-coding RNA, activation of transcription factors by signalling cascades, or direct ligand binding to the nuclear receptors. Analysis of interactions between diet components and human genome structure and gene activity is a modern approach that will help to better understand these relations and will allow designing dietary guidances, which can help maintain good health.
Article
Full-text available
Down syndrome is the leading genetic cause of intellectual disabilities, occurring in 1 out of 700 live births. Given that Down syndrome is caused by an extra copy of chromosome 21 that involves over-expression of 400 genes across a whole chromosome, it precludes any possibility of a genetic therapy. Our lab has long studied the natural dosage compensation mechanism for X chromosome inactivation. To “dosage compensate” X-linked genes between females and males, the X-linked XIST gene produces a large non-coding RNA that silences one of the two X chromosomes in female cells. The initial motivation of this study was to translate the natural mechanisms of X chromosome inactivation into chromosome therapy for Down syndrome. Using genome editing with zinc finger nucleases, we have successfully inserted a large XIST transgene into Chromosome 21 in Down syndrome iPS cells, which results in chromosome-wide transcriptional silencing of the extra Chromosome 21. Remarkably, deficits in proliferation and neural growth are rapidly reversed upon silencing one chromosome 21. Successful trisomy silencing in vitro surmounts the major first step towards potential development of “chromosome therapy” for Down syndrome. The human iPSC-based trisomy correction system we established opens a unique opportunity to identify therapeutic targets and study transplantation therapies for Down syndrome.
Article
Full-text available
We have cloned and characterized 5'-flanking sequences of the DNA methyltransferase (MeTase) gene. DNA MeTase gene transcription is initiated at a few discrete sites: 343 and 90 base pairs upstream of the translation initiation site as determined by RNase protection and primer extension assays. The promoter sequences that regulate expression of DNA MeTase, as defined by chloramphenicol acetyltransferase assays, reside between position -171 and the transcription start site. The promoter of DNA MeTase does not contain TATAA or CAAT boxes and is unusual because it does not contain the CG-rich elements characteristic of TATAA-less housekeeping genes. The 5'-flanking region of DNA MeTase contains AP-1, AP-2 and glucocorticoid response elements, suggesting possible regulation by cellular signal transduction pathways. The base composition of the DNA MeTase promoter is markedly different from that of other housekeeping genes. Whereas most housekeeping genes are characterized by CG-rich areas in their 5'-flanking regions, the TG dinucleotide is over-represented in DNA MeTase 5'-flanking sequences, including a perfect tandem repeat of T/G between positions -685 and -650. DNA methylation patterns play an important role in the developmental regulation of gene expression in vertebrates. DNA MeTase activity is probably regulated to maintain this pattern of methylation. We suggest that the DNA MeTase promoter represents a new class of housekeeping gene promoters that was designed to ensure high fidelity regulation of gene expression.
Article
It has been a controversial issue as to how many DNA cytosine methyltransferase mammalian cells have and whether de novo methylation and maintenance methylation activities are encoded by a single gene or two different genes. To address these questions, we have generated a null mutation of the only known mammalian DNA methyltransferase gene through homologous recombination in mouse embryonic stem cells and found that the development of the homozygous embryos is arrested prior to the 8-somite stage. Surprisingly, the null mutant embryonic stem cells are viable and contain low but stable levels of methyl cytosine and methyltransferase activity, suggesting the existence of a second DNA methyltransferase in mammalian cells. Further studies indicate that de novo methylation activity is not impaired by the mutation as integrated provirus DNA in MoMuLV-infected homozygous embryonic stem cells become methylated at a similar rate as in wild-type cells. Differentiation of mutant cells results in further reduction of methyl cytosine levels, consistent with the de novo methylation activity being down regulated in differentiated cells. These results provide the first evidence that an independently encoded DNA methyltransferase is present in mammalian cells which is capable of de novo methylating cellular and viral DNA in vivo.
Article
Epigenetic reprogramming, characterized by loss of cytosine methylation and histone modifications, occurs during mammalian development in primordial germ cells (PGCs), yet the targets and kinetics of this process are poorly characterized. Here we provide a map of cytosine methylation on a large portion of the genome in developing male and female PGCs isolated from mouse embryos. We show that DNA methylation erasure is global and affects genes of various biological functions. We also reveal complex kinetics of demethylation that are initiated at most genes in early PGC precursors around embryonic day 8.0-9.0. In addition, besides intracisternal A-particles (IAPs), we identify rare LTR-ERV1 retroelements and single-copy sequences that resist global methylation erasure in PGCs as well as in preimplantation embryos. Our data provide important insights into the targets and dynamics of DNA methylation reprogramming in mammalian germ cells.
Chapter
In 1975, two papers suggested a role for DNA methylation in X chromosome inactivation. In one paper (Riggs, 1975), I argued that: 1) DNA methylation should affect protein-DNA interactions; 2) methylation patterns and a maintenance methylase should exist; and 3) DNA methylation should be involved in mammalian cellular differentiative processes. Holliday and Pugh (1975) argued similarly, although less weight was given to X inactivation and more weight was given to the possibility that 5-methylcytosine (5-meCyt) might be deaminated to thymidine; thus a specific mutational change would be generated, as suggested by Scarano (1971). Recently, several studies of X chromosome inactivation have contributed to the emerging body of evidence supporting a role for DNA methylation in mammalian gene regulation; it is these studies that will be reviewed in this chapter. More comprehensive reviews of X chromosome inactivation have been published recently (Gartler and Riggs, 1983; Graves, 1983).