ArticlePDF AvailableLiterature Review

TET enzymes, DNA demethylation and pluripotency

Authors:
  • Andalusian Center for Developmental Biology (CABD)

Abstract and Figures

Ten-eleven translocation (TET) methylcytosine dioxygenases (TET1, TET2, TET3) actively cause demethylation of 5-methylcytosine (5mC) and produce and safeguard hypomethylation at key regulatory regions across the genome. This 5mC erasure is particularly important in pluripotent embryonic stem cells (ESCs) as they need to maintain self-renewal capabilities while retaining the potential to generate different cell types with diverse 5mC patterns. In this review, we discuss the multiple roles of TET proteins in mouse ESCs, and other vertebrate model systems, with a particular focus on TET functions in pluripotency, differentiation, and developmental DNA methylome reprogramming. Furthermore, we elaborate on the recently described non-catalytic roles of TET proteins in diverse biological contexts. Overall, TET proteins are multifunctional regulators that through both their catalytic and non-catalytic roles carry out myriad functions linked to early developmental processes.
Content may be subject to copyright.
TET enzymes, DNA demethylation and pluripotency
Samuel E Ross1, 2 and Ozren Bogdanovic1, 3
1 Genomics and Epigenetics Division, Garvan Institute of Medical Research, Sydney, New South
Wales, 2010, Australia
2 St Vincent's Clinical School, Faculty of Medicine, University of New South Wales, Sydney, New
South Wales, 2010, Australia.
3 School of Biotechnology and Biomolecular Sciences, University of New South Wales, Sydney, New
South Wales, 2052, Australia.
Correspondence to: o.bogdanovic@garvan.org.au
Abstract
Ten-eleven translocation (TET) methylcytosine dioxygenases (TET1, TET2, TET3) actively
cause demethylation of 5-methylcytosine (5mC) and produce and safeguard hypomethylation
at key regulatory regions across the genome. This 5mC erasure is particularly important in
pluripotent embryonic stem cells (ESCs) as they need to maintain self-renewal capabilities
while retaining the potential to generate different cell types with diverse 5mC patterns. In this
Review, we discuss the multiple roles of TET proteins in mouse ESCs, and other vertebrate
model systems, with a particular focus on TET functions in pluripotency, differentiation, and
developmental DNA methylome reprogramming. Furthermore, we elaborate on the recently
described non-catalytic roles of TET proteins in diverse biological contexts. Overall, TET
proteins are multifunctional regulators that through both their catalytic and non-catalytic roles
carry out myriad functions linked to early developmental processes.
Introduction
Ten-eleven translocation (TET) methylcytosine dioxygenases were first described when TET1
was identified as a fusion partner of the mixed lineage leukaemia gene (MLL) in acute myeloid
leukaemia [1]. Since then TET proteins have been associated with other myeloid and lymphoid
malignancies as well as solid cancers including melanoma, breast, and prostate cancers [2].
TET proteins play key roles in the regulation of self-renewal capacities of diverse stem cell
types, and mutations in genes coding for TET proteins can lead to oncogenic transformation
[3]. The major catalytic function of TET proteins was first described in a landmark study, which
revealed that TET1 could catalyse the conversion of 5-methylcytosine (5mC) to 5-
hydroxymethylcytosine (5hmC) [4]. It is now known that TET proteins can cause the sequential
oxidation of 5mC to 5hmC, 5-formylcytosine (5fC), and 5-carboxylcytosine (5caC) [5, 6]. The
oxidised intermediates, 5fC and 5caC, can be removed by thymine DNA glycosylases (TDG)
and base-excision repair (BER) machinery to regenerate unmethylated cytosines at targeted
sites [7, 8]. Specific reader proteins have been described for each of these oxidised
intermediates, which is suggestive of their potential function in gene regulation [9].
This ability of targeted 5mC removal associated with TET function is particularly
important in embryonic stem cells (ESCs) as they need to maintain self-renewal capabilities as
well as adopt diverse 5mC patterns upon differentiation. Mammals and vertebrates such as
zebrafish encode three TET protein copies (TET1, TET2, TET3) [4, 10] whereas in the
Xenopus genus only TET2 and TET3 have been described [11]. Invertebrate chordates such as
amphioxus (Branchiostoma lanceolatum) display a single TET orthologue of yet
uncharacterised function [12]. TET proteins are characterised by their core catalytic domain
and can exist in cell-type specific isoforms with or without their CXXC binding domain
(present in TET1 and TET3, and supplemented in TET2 by its association with CXXC4/IDAX)
[13, 14] (Figure 1). In this Review, we focus on insights related to TET protein function in
mouse embryonic stem cells (mESCs) and vertebrate embryos. More extensive reviews of TET
function have recently been published elsewhere [2, 15].
TET expression and binding dynamics in mESCs
ESCs are derived from epiblasts of mammalian blastocysts and cultured in conditions that
promote propagation of the pluripotent state. ESCs are considered to be either ‘naïve’ when
isolated from pre-implantation epiblasts (E3.5-E4.5), or ‘primed’ when obtained from post-
implantation epiblasts (E5.5-E6.5) [16, 17]. Naïve mouse ESCs are generally cultured in two
types of media. The original culture conditions consisted of leukaemia inhibitory factor (LIF)
and foetal calf serum. While these conditions efficiently maintained the potential for blastocyst
chimera formation, cells in these cultures displayed heterogeneous expression patterns of
pluripotency genes. Furthermore, such cells exhibited 5mC patterns similar to post-
implantation embryos and somatic cells [17]. More recently, culture conditions consisting of
serum-free media and two small molecular inhibitors (2i) that block FGF/ERK pathway and
partly inhibit glycogen synthase 3 (Gsk3), have been developed [18]. Cells grown in 2i
conditions are homogeneous in terms of pluripotency marker expression and morphology, in
addition to displaying low 5mC levels characteristic of pre-implantation embryos [19-21]. Such
cells are thus naïve both in terms of their cellular potency and their epigenome, and are often
referred to as “ground state” naïve ESCs. The majority of earlier studies (e.g. before 2010)
routinely used serum/LIF conditions for mESC culture, thus unless otherwise stated, the studies
discussed below refer to serum/LIF-cultured mESCs.
In mESCs, Tet1 and Tet2 constitute the majority of Tet transcripts while Tet3 is barely
detectable [5, 22, 23] (Figure 2). When LIF is removed from the mESC culture media, mESCs
spontaneously differentiate and this is marked by a significant decrease in Tet1 and 5hmC
levels [4, 22]. In line with these findings, Tet1 levels displayed a progressive drop during the
eight days of embryoid body (EB) culture [22]. This is in contrast to Tet2 transcripts that
decrease during first four days but are fully restored by day eight, and Tet3 transcripts that
displayed a progressive and significant (> 20 fold) increase during this process [22] (Figure
2). It is thus likely that the relatively high 5hmC content (~ 4%) observed in undifferentiated
ESCs [4] is predominantly caused by high levels of TET1, and to a lesser extent, TET2. In
agreement with these findings, in 2i-cultured ESCs, Tet1 is the most highly expressed Tet
transcript whereas Tet3 is undetectable [19]. Interestingly, 2i and serum/LIF conditions differ
in the levels of Tet2, which appears to be more highly expressed in 2i [19]. During serum to 2i
reprogramming, TET1 and TET2 are required for the production of 5hmC even though they
are not the major determinants of the profound 5mC loss associated with this process [19, 20].
Both serum to 2i reprogramming as well as the reprogramming of epiblast stem cells (EpiSCs)
to naïve pluripotency can be enhanced by the addition of retinol and ascorbate to the growth
medium [24]. These components augment TET activity by different mechanisms; ascorbate
directly stimulates TET activity by the reduction of non-enzyme bound Fe3+ to Fe2+, whereas
retinol and retinoid acid (RA) increase expression of Tet2 and Tet3 through binding to
conserved RA-responsive elements [24].
Genome wide profiling of 5hmC and TET1 protein binding sites have revealed an
enrichment of TET1 at CpG island promoters and gene bodies in mESCs, in agreement with
the presence of the CXXC domain in TET1 [25-27] (Figure 1). Nevertheless, the exact role of
CXXC in TET1 targeting to chromatin is unclear, and it appears that TET1 largely depends on
transcription factors such as NANOG, PRDM14, TEX10, and others for its chromatin
recruitment [28-30]. TET1 also exhibits strong enrichment at bivalent promoters that are
marked by both active (H3K4me3) and repressive (H3K27me3) histone marks [27, 31]. These
regulatory signatures frequently decorate promoters of developmental and pluripotency genes
[32]. Similarly to TET1, TET2 can also associate with gene bodies [33] and distal regulatory
elements such as enhancers [34]. The exact roles of TET protein function in mESCs are far
from being fully understood. So far, TETs have been attributed both activator [25] and
repressor [26] function, and have been associated with the regulation of bivalent chromatin
[31] and Polycomb-marked regions [27]. These functions, when impaired, are believed to
impact diverse features of the pluripotent transcriptome with profound consequences for
differentiation processes [23, 34] (Table 1).
TET proteins play key roles during mESC differentiation
Despite the coordinated expression of TETs during pluripotency, TET triple knockout (TKO)
mESCs maintained self-renewal capabilities, normal ESC morphology, and expressed
pluripotency markers such as Oct4 and Nanog [35, 36]. However, embryoid bodies (EBs)
formed from TKO mESCs displayed reduced levels of endo- and mesodermal markers and
resulted in poorly differentiated tissues in general [35]. Moreover, TKO teratomas lacked
mesodermal and endodermal structures whereas TKO mESCs were characterised by low
efficiency of chimeric embryo formation [35]. A number of studies have explored the effects
of single and double TET knockouts (DKO) to reveal specialised roles of TET proteins during
ESC differentiation (Table 1). TET1 targeting studies have reproducibly found that TET1
depletion in mESCs results in a skew towards extraembryonic, mesodermal or endodermal
lineages in embryoid body differentiation assays and teratoma formation assays [5, 23, 25, 36-
38] (Figure 3A). This can be partly explained by the observation that TET1 associates with
the transcriptional repressor and activator SIN3A to activate the nodal antagonist Lefty1 in
mESCs [39]. Depletion of TET1 results in Lefty1 promoter hypermethylation and decreased
expression of Lefty1 and Lefty2 transcripts as well as in an increase in mesoderm/endoderm
transcription factors T and Foxa2 [23, 37, 39]. Moreover, TET1 depletion causes an increase
in trophectoderm markers such as Cdx2, Eomes, and Hand1, and a decrease in neuroectoderm
markers like Pax6 [23] (Figure 3A). The preference toward mesodermal and endodermal
lineage could also be a result of decreased expression of pluripotency markers associated with
restricting endoderm formation such as Esrrb and Prdm14 [25]. TET1 depletion in mESCs also
leads to bivalent promoter DNA hypermethylation followed by an unexpected increase in the
accompanying gene expression levels [27, 36, 37]. It is hypothesised that TET1 at bivalent
promoters recruits Polycomb Repressive Complex 2 (PRC2) to suppress gene expression and
that hypermethylation of these loci would inhibit PRC2 binding. This is supported by observed
physical interactions between TET1 and PRC2 and overlaps in PRC2, TET1 and 5hmC
genomic profiles in mESCs [40]. Unlike TET1, TET2 depletion in mESCs has no visible effect
on the differentiation outcomes of mESCs in embryoid body and teratoma formation assays
[23, 34]. TET2 depletion, however, results in delayed expression of genes associated with early
differentiation stages and also causes a significant reduction in cellular 5hmC levels [34]. As
TET2 mainly associates with enhancers, many of which exhibit low transcription factor
occupancy, its role in mESCs appears to be the priming of regulatory regions for activation
upon differentiation [33, 34].
Unlike Tet1 and Tet2, Tet3 is expressed at low levels in mESCs [5, 22, 23]. TET3 only
contributes to 2% of the total 5hmC in mESCs as assayed by mass spectrometry [36]. However,
knockout of Tet3 in mESCs causes impaired neuroectoderm formation in serum-free embryoid
body assays [41]. This observed skewing is likely caused by the lack of Wnt signalling
suppression, driven partly by promoter hypermethylation and decreased expression of Wnt
inhibitor secreted frizzled-related protein 4 (Sfrp4) upon Tet3 depletion [41].
Additionally, studies in induced pluripotent stem cells (iPSCs) have also supported the
notion of TET proteins as regulators of differentiation. Tet1 alone can be used as a substitute
for Oct4 in the renown OKSM (Oct4, Klf4, Sox2, c-myc ) reprogramming cocktail [42], and
TET proteins have proven essential for the reprogramming of multiple somatic cell types such
as mouse embryonic fibroblasts (MEFs), neural cells, and B cells [28, 43, 44]. Unexpectedly,
however, TET proteins only promote reprogramming in the absence of vitamin C despite it
being known to increase TET activity [45, 46].
In summary, TET TKO and single TET KO studies reveal that while TET proteins are
not required for mESC pluripotency, they are essential for the maintenance of proper
differentiation capacity and the generation of functional embryonic structures. Notably, TET
proteins seem to participate in the regulation of genes with well-established roles in the
suppression of developmental pathways.
TET proteins and developmental DNA methylome reprogramming
While DNA methylomes are generally stable in mammalian somatic cells, the mammalian life
cycle is characterised by two global 5mC reprogramming events [47, 48]. The first genome-
wide 5mC erasure takes place shortly after fertilisation and is characterised by the rapid
removal of 5mC from the paternal pronucleus [49, 50], followed by a progressive drop in 5mC
levels associated with both maternal and paternal genomic contributions [51]. 5mC levels reach
their lowest point during the blastocyst stage, after which the methylomes are gradually re-
established during gastrulation. Similarly, during primordial germ cell (PGC) formation, 5mC
is globally erased and re-established for the second time, and this event is thought to be crucial
for sex-specific imprint establishment [52, 53]. Given the roles of TET proteins in DNA
demethylation, a number of studies have interrogated the links between TETs and global 5mC
erasure in mammals. Initial studies have established links between TET3-mediated DNA
hydroxylation and the reprogramming of the zygotic paternal DNA methylome, following
fertilisation [54]. However, more recent work suggests that inhibition of TET3 activity, as well
as Tet3 deletion, reduces the amount of zygotic 5hmC, without affecting paternal 5mC erasure
[55]. This implies that TET3 plays a protective role in safeguarding hypomethylated genomic
sites rather than being the initiator of this major demethylation event. Similarly, PGC-specific
DNA demethylation can occur in the absence of TET1 and TET2 activity in vitro [56] and in
vivo [57, 58]. While TET1 loss resulted in ~50% reduction of global PGC 5hmC, the TET1
KO PGC genome reached a hypomethylated state at E13.5 comparable to its wild type
counterpart [58]. It is worth noting, however, that Tet1-deficient embryos display abnormalities
associated with imprint erasure [59]. In summary, TET proteins are important regulators of the
genomic 5hmC content, yet their contribution to mammalian global DNA demethylation events
is likely only auxiliary.
Gastrulation and body plan formation in vertebrates is TET-dependent
To assess the roles of TET proteins in vivo, diverse knockout and knockdown strategies have
been employed in mouse, zebrafish and Xenopus (Table 1). Adult germline-specific deletion
of all three TET proteins in mice followed by crossing resulted in TKO progeny that were still
able to form all three embryonic germ layers [60 ]. However, these TET TKO embryos did not
develop past gastrulation and had no discernible early organ structures [60]. These mice also
displayed diffused expression of extraembryonic markers like Eomes and downregulation of
nodal agonists Lefty1 and Lefty2 [60]. This was followed by another mouse TKO study that
demonstrated hyperactive Wnt signalling in neuromesodermal progenitors at similar
embryonic stages (E7.5), suggestive of both Wnt and Nodal signalling being involved in TET-
dependent regulation of embryonic patterning [41]. Whereas TET TKOs are characterised by
lethal gastrulation defects, phenotypes observed upon deletion of single TET proteins are less
severe and their importance for survival are varied. TET2 KO mice develop normally but are
prone to myeloid malignancies and this in agreement with the absence of phenotypes observed
upon TET2 KO in mESCs [61, 62]. Maternal deletion of Tet3 resulted in normal
preimplantation development, however the affected embryos displayed morphological
abnormalities starting from midgestation, and a severely reduced viability rate [54]. Tet1/3
double knockouts (DKOs) were embryonically lethal and no viable embryos could be detected
after E10.5 [63]. The phenotypes of single Tet1 knockouts appear to depend on the genetic
background. Tet1 KOs on a mixed genetic background are viable [37], whereas C57BL/6 KO
Tet1 KO mice display partial embryonic lethality [63, 64]. The majority of the defects observed
in Tet1 KO were associated with gastrulation and involved impairment of primitive streak
formation and misregulation of metabolic genes in the extraembryonic ectoderm [64]. Finally,
the combined deficiency of TET1 and TET2 causes severe developmental abnormalities
including exencephaly and growth retardation and is perinatally lethal; the majority of
Tet1/Tet2 DKO mice die within two days after birth [65]. Interestingly, a fraction of Tet1/Tet2
DKO mice can survive to adulthood with only minor abnormalities. These mice displayed
imprinting defects and somewhat higher levels of global 5mC as measured by MeDIP-seq
approaches. Furthermore, these mice exhibited increased levels of Tet3, which is believed to
compensate for Tet1/Tet2 loss [65].
In zebrafish, tet expression starts between late gastrula and early somitogenesis stages
and coincides with a global increase in 5hmC abundance [10]. tet1/2/3 TKO zebrafish
generated using TALE technology develop past the gastrulation stage and die in the larval
period displaying eye, brain, and Notch signalling defects [66]. Another study, which used
morpholino knockdown (MO) approaches to deplete tet1/2/3 in zebrafish, demonstrated that
affected embryos do not pass gastrulation [67], as previously shown in TET TKO mice [60].
These tet1/2/3 MO embryos also displayed specific DNA hypermethylation of conserved
developmental enhancers related to genes involved in TGF-β, Notch/Delta, and Wnt signalling
pathways [67]. Furthermore, this TET-dependent demethylation of embryonic enhancers is a
conserved feature of the vertebrate phylotypic stage, the most conserved period of vertebrate
embryogenesis [67]. In the frog Xenopus laevis, tet3 MO depletion results in severe eye and
neural phenotypes and greatly reduces the expression of neural development genes like pax6
[11]. These phenotypes are in agreement with phenotypes observed upon zebrafish tet3 MO
knockdown that was characterised by microphthalmia [67]. In summary, vertebrates display
diverse requirements for TET proteins during embryogenesis with specific TET proteins
playing both common and organism-specific roles within the vertebrate subphylum. A major
feature, consistent across all vertebrates, is the lack of requirement for TET proteins during
pluripotency, and the absolute requirement for TET function during gastrulation and body plan
formation (Figure 3B). This requirement is underpinned by the timely activation of distal
regulatory regions by means of active DNA demethylation, associated with key developmental
pathways.
!
DNA Demethylation-Independent Functions of TETs
While the major described function of TET proteins is the sequential 5mC oxidation, a growing
body of work has now identified important non-catalytic functions for TETs in diverse model
systems. A recent study using full-length TET1 knockout mice found that their phenotypes
differed from other catalytic TET1 KO mice and therefore proposed an important non-catalytic
role for the N-terminus of TET1 in embryo development [64]. TET2 can also regulate gene
expression by directly interacting with O-linked N-acetylglucosamine (O-GlcNAc) transferase
(OGT) to promote histone O-GlcNAcylation, which is a process that appears to be independent
of TET2 enzymatic activity [68]. This can subsequently promote H3K4me3 deposition through
the SET1/COMPASS complex [69]. Notably, this interaction with OGT has also been proposed
to enhance and regulate the catalytic activity of TET1 as demonstrated in a recent study that
utilised zebrafish embryos and mESCs [70]. TET proteins may also possess catalytic activity
towards RNA and not just DNA. Recently it has been shown that Paraspeckle Component 1
(PSPC1), an RNA binding protein, directly binds TET2 and directs it to MERVL transcripts
where TET2 then demethylates and destabilizes the RNA leading to a decrease in its expression
[71].
Conclusions and perspectives
TET depletion studies in mESCs and other vertebrate model systems have revealed
both specialised and redundant roles for TET proteins in differentiation, gastrulation,
and body plan formation but not during pluripotency; TETs appear to be dispensable
for pluripotency both in vivo and in vitro [35, 60, 67]. The roles of TET proteins are
currently under intense investigation because of their apparent links with cancer [2] and
clonal hematopoiesis [72].
Many questions related to TET function remain unanswered. For example, no unifying
mechanism that describes TET targeting to chromatin has been revealed to date.
Biochemical studies have identified a number of transcription factors that are proposed
to aid TET recruitment to DNA [28-30, 44], however it is not yet clear how
evolutionarily conserved these mechanisms are. Moreover, the exact role of 5hmC in
gene regulation is far from being understood. TKO studies in mice and tet1/2/3
morphants in zebrafish suggest that TET activity is required for demethylation and
activation of distal regulatory elements associated with developmental pathways [35,
41, 60, 66, 67]. It has been proposed that such 5mC to 5hmC conversion could
participate in the reconfiguration of chromatin structure that might be required for
subsequent transcription factor binding [73]. This is supported by notions that hmC-
marked nucleosomes are less stable in vitro [74] and that the methyl-CpG-binding
domain (MBD) of a key 5mC-dependent repressor, MeCP2, binds 5hmC ~20 fold less
stronger than 5mC [75].
With the advent of single cell epigenome profiling technologies [76] TET proteins and
5hmC are again gaining significant traction. A recent study demonstrated the utility of
single-cell 5hmC sequencing in lineage reconstruction during early mammalian
embryogenesis [77]. It is expected that such high-resolution epigenome profiling
techniques will greatly aid in disentangling the complex relationships between 5mC,
5hmC, and genome regulation, during vertebrate embryonic development and disease
formation.
Funding
Australian Research Council (ARC) Discovery Project (DP190103852) supported this work.
O.B. is supported by NHRMC (R.D. Wright Biomedical CDF APP1162993) and CINSW
(Career Development Fellowship CDF181229).
Competing Interests
The Authors declare that there are no competing interests associated with the manuscript
Table 1. Overview of TET depletion studies in mESCs and vertebrate embryos
Model
Depletion strategy
Phenotype
Reference
mESC
Tet1, Tet2, Tet3
shRNA KD
Impaired mESC self-renewal and maintenance, and
morphological abnormalities upon Tet1 KD
[5]
Tet1, Tet2, Tet3
siRNA KD; Tet1,
Tet2 shRNA KD
Global loss of 5hmC, increase in trophectoderm (Cdx2,
Eomes, Hand1), decrease in neuroectoderm (Pax6 and
Neurod1), and decreased Nodal antagonist (Lefty1 and
Lefty2) expression in Tet1/Tet2 KD. Minor increase in
Pax6, Neurod1, Lefty1/2 expression upon Tet2 KD. Tet3
depletion causes Lefty2 repression.
[23]
Tet1/2 siRNA KD;
Tet1 shRNA KD
Global loss of 5hmC, down-regulation of pluripotency
genes, increased expression of extraembryonic endoderm
differentiation markers, and increased Cdx2 expression
upon Tet1/Tet2 knockdown
[25]
Tet1 KO
Altered expression of lineage specification markers T and
Pax6, low efficiency of EB formation, loss of 5hmC and
trophectoderm-like cells in Tet KO teratomas
[37]
Tet1 siRNA KD
Loss of stem cell identity, morphological changes, global
5hmC loss, impaired LIF/Stat3 signalling, and upregulation
of differentiation markers
[38]
Tet1/2 DKO
Global reduction of 5hmC levels, trophectdoerm-like cells
in Tet1/2 DKO teratomas
[65]
Tet1/2/3 TKO
Global loss of 5hmC , defects in EB and teratoma
formation, poor contribution to chimeric embryos,
increased promoter 5mC, deregulation of developmental
genes
[35]
Tet1 TKO; Tet2 KO
hmC loss of 44% and 90,7% in Tet1 KO and Tet2 KO
respectively, increased 5mC on enhancers and delayed
gene induction during ES-NPC differentiation. in Tet2 KO,
[34]
Tet1/2/3 TKO
Lower (< 5%) global 5mC levels, increased 2C-like
population in Tet TKO ESCs
[36]
Tet3 KO; Tet1/2/3
TKO
Skewed differentiation toward cardiac mesoderm and
down-regulation of Wnt signalling inhibitor Sfrp4 during
differentiation of Tet3 KO mESCs, decreased
neuroectdoerm formation in Tet1/2/3 TKO
[41]
mouse
Tet1 KO
75% of homozygous pups displaying smaller body size
[37]
Conditional Tet2
KO (hematopoietic
compartment)
Enlargement of the hematopoietic stem cell compartment,
splenomegaly, monocytosis, and extramedullary
hematopoiesis, increased capacity od stem cell self-
renewal, extramedulary hematopoiesis.
[61]
Tet2 KO
Global loss of 5hmC and increase in 5mC in bone marrow
cells, chronic myelomonocytic leukemia - like phenotype at
2-4 months, splenomegaly and hepatomegaly, lethal
myeloid malignancies
[62]
Maternal Tet3 KO
Defects in paternal 5mC reprogramming, morphological
abnormalities at mid-gestation, reduced 5hmC levels, >
50% less viable pups at birth
[54]
Tet1 KO
Loss of oocytes, smaller ovaries, small litter size, reduced
fertility, impaired meiotic gene activation
[57, 59]
Tet1/2 DKO
Perinatal lethal (60%) with exencephaly, head
haemorrhage, growth retardation. 40% survivors display
smaller ovaries, reduced fertility, decreased global 5hmC,
partially compromised imprinting
[65]
Tet1 KO
Reduced global 5hmC levels in the brain, impairment in
memory extinction, increased hippocampal long-term
depression, down regulation of neuronal genes (Npas4, c-
Fos, Egr2, Egr4, Npas4)
[78]
Tet1 KO
Impaired spatial learning and memory, impaired
maintenance of the neural progenitor pool, defective adult
neurogenesis
[79]
Tet1/3 KO
Embryonic lethality, delayed early development, mitotic
defects, apoptosis, global loss of 5hmC and gain of 5mC,
transcriptome variability, loss of Nanog expression, poor
separation of germ layers
[63]
Tet1/2/3 TKO
Complete embryonic lethality with severe gastrulation
defects, hyperactive Nodal signalling due to abnormal
methylation of Lefty1 and Lefty2 regulatory regions
[60]
Tet1/2/3 TKO
Embryonic lethality, hyperactivation of Wnt singling in
neuromesodermal progenitors resulting in a skew towards
mesoderm fate.
[41]
zebrafish
(Danio
rerio)
tet1/2/3 TKO;
tet1/2 DKO
Lethality following larval period, global reduction in 5hmC
(> 30 fold) levels, smaller eyes, abnormal brain
morphology, altered pigmentation at 36hpf, loss of
differentiated definitive blood cells, aberrant Notch
signalling in both tet1/2/3 TKO and tet1/2 DKO embryos
[66]
tet1/2/3 MO; tet3
MO
Embryonic lethality (> 75%) with severe gastrulation
defects, short and blended axes, impaired head structures,
small eyes and reduced pigmentation in tet1/2/3 MO.
Minor defects such as microphthalmia in tet3 MO
[67]
Xenopus
laevis
tet3 MO
Eye malformations, microcephaly, reduced pigmentation,
reduced expression of neural crest and neuronal markers,
[11]
Abbreviations
5mC, 5-methylcytosine; 5hmC, 5-hydroxymethylcytosine; 5fC, 5-formylcytosine; 5caC, 5-
carboxylcytosine; BER, base excision repair; DKO, double knock out; EB, embryoid body;
ESC, embryonic stem cells; KD, knockdown; KO, knock out; LIF, Leukemia inhibitory factor;
mESC, mouse embryonic stem cells; MO, morpholino; PRC2, polycomb repressive complex
2; TDG, thymine DNA glycosylase; TET, Ten-eleven translocation methylcytosine
dioxygenase; TKO, triple knock out; siRNA, small interfering RNA; shRNA, small hairpin
RNA; 2C - 2 cell.
References
1 Lorsbach, R. B., Moore, J., Mathew, S., Raimondi, S. C., Mukatira, S. T. and Downing,
J. R. (2003) TET1, a member of a novel protein family, is fused to MLL in acute myeloid
leukemia containing the t(10;11)(q22;q23). Leukemia. 17, 637
2 Rasmussen, K. D. and Helin, K. (2016) Role of TET enzymes in DNA methylation,
development, and cancer. Genes Dev. 30, 733-750
3 Cimmino, L., Abdel-Wahab, O., Levine, R. L. and Aifantis, I. (2011) TET family
proteins and their role in stem cell differentiation and transformation. Cell Stem Cell. 9, 193-
204
4 Tahiliani, M., Koh, K. P., Shen, Y., Pastor, W. A., Bandukwala, H., Brudno, Y.,
Agarwal, S., Iyer, L. M., Liu, D. R., Aravind, L. and Rao, A. (2009) Conversion of 5-
methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1.
Science. 324, 930-935
5 Ito, S., D'Alessio, A. C., Taranova, O. V., Hong, K., Sowers, L. C. and Zhang, Y. (2010)
Role of Tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass
specification. Nature. 466, 1129-1133
6 Ito, S., Shen, L., Dai, Q., Wu, S. C., Collins, L. B., Swenberg, J. A., He, C. and Zhang,
Y. (2011) Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-
carboxylcytosine. Science. 333, 1300-1303
7 Maiti, A. and Drohat, A. C. (2011) Thymine DNA glycosylase can rapidly excise 5-
formylcytosine and 5-carboxylcytosine: potential implications for active demethylation of CpG
sites. J Biol Chem. 286, 35334-35338
8 He, Y. F., Li, B. Z., Li, Z., Liu, P., Wang, Y., Tang, Q., Ding, J., Jia, Y., Chen, Z., Li,
L., Sun, Y., Li, X., Dai, Q., Song, C. X., Zhang, K., He, C. and Xu, G. L. (2011) Tet-mediated
formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science. 333,
1303-1307
9 Spruijt, C. G., Gnerlich, F., Smits, A. H., Pfaffeneder, T., Jansen, P. W., Bauer, C.,
Munzel, M., Wagner, M., Muller, M., Khan, F., Eberl, H. C., Mensinga, A., Brinkman, A. B.,
Lephikov, K., Muller, U., Walter, J., Boelens, R., van Ingen, H., Leonhardt, H., Carell, T. and
Vermeulen, M. (2013) Dynamic readers for 5-(hydroxy)methylcytosine and its oxidized
derivatives. Cell. 152, 1146-1159
10 Almeida, R. D., Loose, M., Sottile, V., Matsa, E., Denning, C., Young, L., Johnson, A.
D., Gering, M. and Ruzov, A. (2012) 5-hydroxymethyl-cytosine enrichment of non-committed
cells is not a universal feature of vertebrate development. Epigenetics. 7, 383-389
11 Xu, Y., Xu, C., Kato, A., Tempel, W., Abreu, J. G., Bian, C., Hu, Y., Hu, D., Zhao, B.,
Cerovina, T., Diao, J., Wu, F., He, H. H., Cui, Q., Clark, E., Ma, C., Barbara, A., Veenstra, G.
J. C., Xu, G., Kaiser, U. B., Liu, X. S., Sugrue, S. P., He, X., Min, J., Kato, Y. and Shi, Y. G.
(2012) Tet3 CXXC domain and dioxygenase activity cooperatively regulate key genes for
Xenopus eye and neural development. Cell. 151, 1200-1213
12 Marlétaz, F., Firbas, P. N., Maeso, I., Tena, J. J., Bogdanovic, O., Perry, M., Wyatt, C.
D. R., de la Calle-Mustienes, E., Bertrand, S., Burguera, D., Acemel, R. D., van Heeringen, S.
J., Naranjo, S., Herrera-Ubeda, C., Skvortsova, K., Jimenez-Gancedo, S., Aldea, D., Marquez,
Y., Buono, L., Kozmikova, I., Permanyer, J., Louis, A., Albuixech-Crespo, B., Le Petillon, Y.,
Leon, A., Subirana, L., Balwierz, P. J., Duckett, P. E., Farahani, E., Aury, J.-M., Mangenot, S.,
Wincker, P., Albalat, R., Benito-Gutiérrez, È., Cañestro, C., Castro, F., D’Aniello, S., Ferrier,
D. E. K., Huang, S., Laudet, V., Marais, G. A. B., Pontarotti, P., Schubert, M., Seitz, H.,
Somorjai, I., Takahashi, T., Mirabeau, O., Xu, A., Yu, J.-K., Carninci, P., Martinez-Morales,
J. R., Crollius, H. R., Kozmik, Z., Weirauch, M. T., Garcia-Fernàndez, J., Lister, R., Lenhard,
B., Holland, P. W. H., Escriva, H., Gómez-Skarmeta, J. L. and Irimia, M. (2018) Amphioxus
functional genomics and the origins of vertebrate gene regulation. Nature. 564, 64-70
13 Akahori, H., Guindon, S., Yoshizaki, S. and Muto, Y. (2015) Molecular Evolution of
the TET Gene Family in Mammals. Int J Mol Sci. 16, 28472-28485
14 Melamed, P., Yosefzon, Y., David, C., Tsukerman, A. and Pnueli, L. (2018) Tet
Enzymes, Variants, and Differential Effects on Function. Frontiers in cell and developmental
biology. 6, 22-22
15 Wu, X. and Zhang, Y. (2017) TET-mediated active DNA demethylation: mechanism,
function and beyond. Nature Reviews Genetics. 18, 517
16 Davidson, K. C., Mason, E. A. and Pera, M. F. (2015) The pluripotent state in mouse
and human. Development. 142, 3090-3099
17 Kinoshita, M. and Smith, A. (2018) Pluripotency Deconstructed. Dev Growth Differ.
60, 44-52
18 Ying, Q. L., Wray, J., Nichols, J., Batlle-Morera, L., Doble, B., Woodgett, J., Cohen,
P. and Smith, A. (2008) The ground state of embryonic stem cell self-renewal. Nature. 453,
519-523
19 Ficz, G., Hore, T. A., Santos, F., Lee, H. J., Dean, W., Arand, J., Krueger, F., Oxley,
D., Paul, Y. L., Walter, J., Cook, S. J., Andrews, S., Branco, M. R. and Reik, W. (2013) FGF
signaling inhibition in ESCs drives rapid genome-wide demethylation to the epigenetic ground
state of pluripotency. Cell Stem Cell. 13, 351-359
20 Habibi, E., Brinkman, A. B., Arand, J., Kroeze, L. I., Kerstens, H. H., Matarese, F.,
Lepikhov, K., Gut, M., Brun-Heath, I., Hubner, N. C., Benedetti, R., Altucci, L., Jansen, J. H.,
Walter, J., Gut, I. G., Marks, H. and Stunnenberg, H. G. (2013) Whole-genome bisulfite
sequencing of two distinct interconvertible DNA methylomes of mouse embryonic stem cells.
Cell Stem Cell. 13, 360-369
21 Leitch, H. G., McEwen, K. R., Turp, A., Encheva, V., Carroll, T., Grabole, N.,
Mansfield, W., Nashun, B., Knezovich, J. G., Smith, A., Surani, M. A. and Hajkova, P. (2013)
Naive pluripotency is associated with global DNA hypomethylation. Nat Struct Mol Biol. 20,
311-316
22 Szwagierczak, A., Bultmann, S., Schmidt, C. S., Spada, F. and Leonhardt, H. (2010)
Sensitive enzymatic quantification of 5-hydroxymethylcytosine in genomic DNA. Nucleic
acids research. 38, e181-e181
23 Koh, K. P., Yabuuchi, A., Rao, S., Huang, Y., Cunniff, K., Nardone, J., Laiho, A.,
Tahiliani, M., Sommer, C. A., Mostoslavsky, G., Lahesmaa, R., Orkin, S. H., Rodig, S. J.,
Daley, G. Q. and Rao, A. (2011) Tet1 and Tet2 regulate 5-hydroxymethylcytosine production
and cell lineage specification in mouse embryonic stem cells. Cell stem cell. 8, 200-213
24 Hore, T. A., von Meyenn, F., Ravichandran, M., Bachman, M., Ficz, G., Oxley, D.,
Santos, F., Balasubramanian, S., Jurkowski, T. P. and Reik, W. (2016) Retinol and ascorbate
drive erasure of epigenetic memory and enhance reprogramming to naive pluripotency by
complementary mechanisms. Proc Natl Acad Sci U S A. 113, 12202-12207
25 Ficz, G., Branco, M. R., Seisenberger, S., Santos, F., Krueger, F., Hore, T. A., Marques,
C. J., Andrews, S. and Reik, W. (2011) Dynamic regulation of 5-hydroxymethylcytosine in
mouse ES cells and during differentiation. Nature. 473, 398-402
26 Williams, K., Christensen, J., Pedersen, M. T., Johansen, J. V., Cloos, P. A., Rappsilber,
J. and Helin, K. (2011) TET1 and hydroxymethylcytosine in transcription and DNA
methylation fidelity. Nature. 473, 343-348
27 Wu, H., D'Alessio, A. C., Ito, S., Xia, K., Wang, Z., Cui, K., Zhao, K., Sun, Y. E. and
Zhang, Y. (2011) Dual functions of Tet1 in transcriptional regulation in mouse embryonic stem
cells. Nature. 473, 389-393
28 Costa, Y., Ding, J., Theunissen, T. W., Faiola, F., Hore, T. A., Shliaha, P. V., Fidalgo,
M., Saunders, A., Lawrence, M., Dietmann, S., Das, S., Levasseur, D. N., Li, Z., Xu, M., Reik,
W., Silva, J. C. R. and Wang, J. (2013) NANOG-dependent function of TET1 and TET2 in
establishment of pluripotency. Nature. 495, 370-374
29 Okashita, N., Kumaki, Y., Ebi, K., Nishi, M., Okamoto, Y., Nakayama, M., Hashimoto,
S., Nakamura, T., Sugasawa, K., Kojima, N., Takada, T., Okano, M. and Seki, Y. (2014)
PRDM14 promotes active DNA demethylation through the ten-eleven translocation (TET)-
mediated base excision repair pathway in embryonic stem cells. Development. 141, 269-280
30 Ding, J., Huang, X., Shao, N., Zhou, H., Lee, D. F., Faiola, F., Fidalgo, M., Guallar, D.,
Saunders, A., Shliaha, P. V., Wang, H., Waghray, A., Papatsenko, D., Sanchez-Priego, C., Li,
D., Yuan, Y., Lemischka, I. R., Shen, L., Kelley, K., Deng, H., Shen, X. and Wang, J. (2015)
Tex10 Coordinates Epigenetic Control of Super-Enhancer Activity in Pluripotency and
Reprogramming. Cell Stem Cell. 16, 653-668
31 Verma, N., Pan, H., Doré, L. C., Shukla, A., Li, Q. V., Pelham-Webb, B., Teijeiro, V.,
González, F., Krivtsov, A., Chang, C.-J., Papapetrou, E. P., He, C., Elemento, O. and Huangfu,
D. (2018) TET proteins safeguard bivalent promoters from de novo methylation in human
embryonic stem cells. Nature genetics. 50, 83-95
32 Voigt, P., Tee, W. W. and Reinberg, D. (2013) A double take on bivalent promoters.
Genes Dev. 27, 1318-1338
33 Huang, Y., Chavez, L., Chang, X., Wang, X., Pastor, W. A., Kang, J., Zepeda-Martínez,
J. A., Pape, U. J., Jacobsen, S. E., Peters, B. and Rao, A. (2014) Distinct roles of the
methylcytosine oxidases Tet1 and Tet2 in mouse embryonic stem cells. Proceedings of the
National Academy of Sciences of the United States of America. 111, 1361-1366
34 Hon, G. C., Song, C. X., Du, T., Jin, F., Selvaraj, S., Lee, A. Y., Yen, C. A., Ye, Z.,
Mao, S. Q., Wang, B. A., Kuan, S., Edsall, L. E., Zhao, B. S., Xu, G. L., He, C. and Ren, B.
(2014) 5mC oxidation by Tet2 modulates enhancer activity and timing of transcriptome
reprogramming during differentiation. Molecular cell. 56, 286-297
35 Dawlaty, M. M., Breiling, A., Le, T., Barrasa, M. I., Raddatz, G., Gao, Q., Powell, B.
E., Cheng, A. W., Faull, K. F., Lyko, F. and Jaenisch, R. (2014) Loss of Tet enzymes
compromises proper differentiation of embryonic stem cells. Developmental cell. 29, 102-111
36 Lu, F., Liu, Y., Jiang, L., Yamaguchi, S. and Zhang, Y. (2014) Role of Tet proteins in
enhancer activity and telomere elongation. Genes Dev. 28, 2103-2119
37 Dawlaty, M. M., Ganz, K., Powell, B. E., Hu, Y.-C., Markoulaki, S., Cheng, A. W.,
Gao, Q., Kim, J., Choi, S.-W., Page, D. C. and Jaenisch, R. (2011) Tet1 is dispensable for
maintaining pluripotency and its loss is compatible with embryonic and postnatal development.
Cell stem cell. 9, 166-175
38 Freudenberg, J. M., Ghosh, S., Lackford, B. L., Yellaboina, S., Zheng, X., Li, R.,
Cuddapah, S., Wade, P. A., Hu, G. and Jothi, R. (2012) Acute depletion of Tet1-dependent 5-
hydroxymethylcytosine levels impairs LIF/Stat3 signaling and results in loss of embryonic
stem cell identity. Nucleic acids research. 40, 3364-3377
39 Zhu, F., Zhu, Q., Ye, D., Zhang, Q., Yang, Y., Guo, X., Liu, Z., Jiapaer, Z., Wan, X.,
Wang, G., Chen, W., Zhu, S., Jiang, C., Shi, W. and Kang, J. (2018) Sin3a-Tet1 interaction
activates gene transcription and is required for embryonic stem cell pluripotency. Nucleic acids
research. 46, 6026-6040
40 Neri, F., Incarnato, D., Krepelova, A., Rapelli, S., Pagnani, A., Zecchina, R., Parlato,
C. and Oliviero, S. (2013) Genome-wide analysis identifies a functional association of Tet1
and Polycomb repressive complex 2 in mouse embryonic stem cells. Genome Biol. 14, R91
41 Li, X., Yue, X., Pastor, W. A., Lin, L., Georges, R., Chavez, L., Evans, S. M. and Rao,
A. (2016) Tet proteins influence the balance between neuroectodermal and mesodermal fate
choice by inhibiting Wnt signaling. Proc Natl Acad Sci U S A. 113, E8267-e8276
42 Gao, Y., Chen, J., Li, K., Wu, T., Huang, B., Liu, W., Kou, X., Zhang, Y., Huang, H.,
Jiang, Y., Yao, C., Liu, X., Lu, Z., Xu, Z., Kang, L., Chen, J., Wang, H., Cai, T. and Gao, S.
(2013) Replacement of Oct4 by Tet1 during iPSC induction reveals an important role of DNA
methylation and hydroxymethylation in reprogramming. Cell Stem Cell. 12, 453-469
43 Di Stefano, B., Sardina, J. L., van Oevelen, C., Collombet, S., Kallin, E. M., Vicent, G.
P., Lu, J., Thieffry, D., Beato, M. and Graf, T. (2014) C/EBPalpha poises B cells for rapid
reprogramming into induced pluripotent stem cells. Nature. 506, 235-239
44 Sardina, J. L., Collombet, S., Tian, T. V., Gomez, A., Di Stefano, B., Berenguer, C.,
Brumbaugh, J., Stadhouders, R., Segura-Morales, C., Gut, M., Gut, I. G., Heath, S., Aranda,
S., Di Croce, L., Hochedlinger, K., Thieffry, D. and Graf, T. (2018) Transcription Factors Drive
Tet2-Mediated Enhancer Demethylation to Reprogram Cell Fate. Cell Stem Cell. 23, 727-
741.e729
45 Chen, J., Guo, L., Zhang, L., Wu, H., Yang, J., Liu, H., Wang, X., Hu, X., Gu, T., Zhou,
Z., Liu, J., Liu, J., Wu, H., Mao, S. Q., Mo, K., Li, Y., Lai, K., Qi, J., Yao, H., Pan, G., Xu, G.
L. and Pei, D. (2013) Vitamin C modulates TET1 function during somatic cell reprogramming.
Nat Genet. 45, 1504-1509
46 Blaschke, K., Ebata, K. T., Karimi, M. M., Zepeda-Martinez, J. A., Goyal, P.,
Mahapatra, S., Tam, A., Laird, D. J., Hirst, M., Rao, A., Lorincz, M. C. and Ramalho-Santos,
M. (2013) Vitamin C induces Tet-dependent DNA demethylation and a blastocyst-like state in
ES cells. Nature. 500, 222-226
47 Lee, H. J., Hore, T. A. and Reik, W. (2014) Reprogramming the methylome: erasing
memory and creating diversity. Cell Stem Cell. 14, 710-719
48 Skvortsova, K., Iovino, N. and Bogdanovic, O. (2018) Functions and mechanisms of
epigenetic inheritance in animals. Nat Rev Mol Cell Biol. 19, 774-790
49 Oswald, J., Engemann, S., Lane, N., Mayer, W., Olek, A., Fundele, R., Dean, W., Reik,
W. and Walter, J. (2000) Active demethylation of the paternal genome in the mouse zygote.
Current biology : CB. 10, 475-478
50 Mayer, W., Niveleau, A., Walter, J., Fundele, R. and Haaf, T. (2000) Demethylation of
the zygotic paternal genome. Nature. 403, 501-502
51 Santos, F., Hendrich, B., Reik, W. and Dean, W. (2002) Dynamic reprogramming of
DNA methylation in the early mouse embryo. Developmental biology. 241, 172-182
52 Guibert, S., Forne, T. and Weber, M. (2012) Global profiling of DNA methylation
erasure in mouse primordial germ cells. Genome Res. 22, 633-641
53 Hackett, J. A., Sengupta, R., Zylicz, J. J., Murakami, K., Lee, C., Down, T. A. and
Surani, M. A. (2013) Germline DNA demethylation dynamics and imprint erasure through 5-
hydroxymethylcytosine. Science. 339, 448-452
54 Gu, T. P., Guo, F., Yang, H., Wu, H. P., Xu, G. F., Liu, W., Xie, Z. G., Shi, L., He, X.,
Jin, S. G., Iqbal, K., Shi, Y. G., Deng, Z., Szabo, P. E., Pfeifer, G. P., Li, J. and Xu, G. L. (2011)
The role of Tet3 DNA dioxygenase in epigenetic reprogramming by oocytes. Nature. 477, 606-
610
55 Amouroux, R., Nashun, B., Shirane, K., Nakagawa, S., Hill, P. W., D'Souza, Z.,
Nakayama, M., Matsuda, M., Turp, A., Ndjetehe, E., Encheva, V., Kudo, N. R., Koseki, H.,
Sasaki, H. and Hajkova, P. (2016) De novo DNA methylation drives 5hmC accumulation in
mouse zygotes. Nat Cell Biol. 18, 225-233
56 Vincent, J. J., Huang, Y., Chen, P. Y., Feng, S., Calvopina, J. H., Nee, K., Lee, S. A.,
Le, T., Yoon, A. J., Faull, K., Fan, G., Rao, A., Jacobsen, S. E., Pellegrini, M. and Clark, A. T.
(2013) Stage-specific roles for tet1 and tet2 in DNA demethylation in primordial germ cells.
Cell Stem Cell. 12, 470-478
57 Yamaguchi, S., Hong, K., Liu, R., Shen, L., Inoue, A., Diep, D., Zhang, K. and Zhang,
Y. (2012) Tet1 controls meiosis by regulating meiotic gene expression. Nature. 492, 443-447
58 Hill, P. W. S., Leitch, H. G., Requena, C. E., Sun, Z., Amouroux, R., Roman-Trufero,
M., Borkowska, M., Terragni, J., Vaisvila, R., Linnett, S., Bagci, H., Dharmalingham, G.,
Haberle, V., Lenhard, B., Zheng, Y., Pradhan, S. and Hajkova, P. (2018) Epigenetic
reprogramming enables the transition from primordial germ cell to gonocyte. Nature. 555, 392-
396
59 Yamaguchi, S., Shen, L., Liu, Y., Sendler, D. and Zhang, Y. (2013) Role of Tet1 in
erasure of genomic imprinting. Nature. 504, 460-464
60 Dai, H. Q., Wang, B. A., Yang, L., Chen, J. J., Zhu, G. C., Sun, M. L., Ge, H., Wang,
R., Chapman, D. L., Tang, F., Sun, X. and Xu, G. L. (2016) TET-mediated DNA demethylation
controls gastrulation by regulating Lefty-Nodal signalling. Nature. 538, 528-532
61 Moran-Crusio, K., Reavie, L., Shih, A., Abdel-Wahab, O., Ndiaye-Lobry, D., Lobry,
C., Figueroa, M. E., Vasanthakumar, A., Patel, J., Zhao, X., Perna, F., Pandey, S., Madzo, J.,
Song, C., Dai, Q., He, C., Ibrahim, S., Beran, M., Zavadil, J., Nimer, S. D., Melnick, A.,
Godley, L. A., Aifantis, I. and Levine, R. L. (2011) Tet2 loss leads to increased hematopoietic
stem cell self-renewal and myeloid transformation. Cancer Cell. 20, 11-24
62 Li, Z., Cai, X., Cai, C. L., Wang, J., Zhang, W., Petersen, B. E., Yang, F. C. and Xu,
M. (2011) Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and
subsequent development of myeloid malignancies. Blood. 118, 4509-4518
63 Kang, J., Lienhard, M., Pastor, W. A., Chawla, A., Novotny, M., Tsagaratou, A.,
Lasken, R. S., Thompson, E. C., Surani, M. A., Koralov, S. B., Kalantry, S., Chavez, L. and
Rao, A. (2015) Simultaneous deletion of the methylcytosine oxidases Tet1 and Tet3 increases
transcriptome variability in early embryogenesis. Proc Natl Acad Sci U S A. 112, E4236-4245
64 Khoueiry, R., Sohni, A., Thienpont, B., Luo, X., Velde, J. V., Bartoccetti, M., Boeckx,
B., Zwijsen, A., Rao, A., Lambrechts, D. and Koh, K. P. (2017) Lineage-specific functions of
TET1 in the postimplantation mouse embryo. Nat Genet. 49, 1061-1072
65 Dawlaty, M. M., Breiling, A., Le, T., Raddatz, G., Barrasa, M. I., Cheng, A. W., Gao,
Q., Powell, B. E., Li, Z., Xu, M., Faull, K. F., Lyko, F. and Jaenisch, R. (2013) Combined
deficiency of Tet1 and Tet2 causes epigenetic abnormalities but is compatible with postnatal
development. Developmental cell. 24, 310-323
66 Li, C., Lan, Y., Schwartz-Orbach, L., Korol, E., Tahiliani, M., Evans, T. and Goll, M.
G. (2015) Overlapping Requirements for Tet2 and Tet3 in Normal Development and
Hematopoietic Stem Cell Emergence. Cell Rep. 12, 1133-1143
67 Bogdanovic, O., Smits, A. H., de la Calle Mustienes, E., Tena, J. J., Ford, E., Williams,
R., Senanayake, U., Schultz, M. D., Hontelez, S., van Kruijsbergen, I., Rayon, T., Gnerlich, F.,
Carell, T., Veenstra, G. J., Manzanares, M., Sauka-Spengler, T., Ecker, J. R., Vermeulen, M.,
Gomez-Skarmeta, J. L. and Lister, R. (2016) Active DNA demethylation at enhancers during
the vertebrate phylotypic period. Nat Genet. 48, 417-426
68 Chen, Q., Chen, Y., Bian, C., Fujiki, R. and Yu, X. (2013) TET2 promotes histone O-
GlcNAcylation during gene transcription. Nature. 493, 561-564
69 Deplus, R., Delatte, B., Schwinn, M. K., Defrance, M., Mendez, J., Murphy, N.,
Dawson, M. A., Volkmar, M., Putmans, P., Calonne, E., Shih, A. H., Levine, R. L., Bernard,
O., Mercher, T., Solary, E., Urh, M., Daniels, D. L. and Fuks, F. (2013) TET2 and TET3
regulate GlcNAcylation and H3K4 methylation through OGT and SET1/COMPASS. Embo j.
32, 645-655
70 Hrit, J., Li, C., Martin, E. A., Simental, E., Goll, M. and Panning, B. (2018) OGT binds
a conserved C-terminal domain of TET1 toregulate TET1 activity and function in development.
bioRxiv, 253872
71 Guallar, D., Bi, X., Pardavila, J. A., Huang, X., Saenz, C., Shi, X., Zhou, H., Faiola, F.,
Ding, J., Haruehanroengra, P., Yang, F., Li, D., Sanchez-Priego, C., Saunders, A., Pan, F.,
Valdes, V. J., Kelley, K., Blanco, M. G., Chen, L., Wang, H., Sheng, J., Xu, M., Fidalgo, M.,
Shen, X. and Wang, J. (2018) RNA-dependent chromatin targeting of TET2 for endogenous
retrovirus control in pluripotent stem cells. Nature genetics. 50, 443-451
72 Fuster, J. J., MacLauchlan, S., Zuriaga, M. A., Polackal, M. N., Ostriker, A. C.,
Chakraborty, R., Wu, C. L., Sano, S., Muralidharan, S., Rius, C., Vuong, J., Jacob, S.,
Muralidhar, V., Robertson, A. A., Cooper, M. A., Andres, V., Hirschi, K. K., Martin, K. A.
and Walsh, K. (2017) Clonal hematopoiesis associated with TET2 deficiency accelerates
atherosclerosis development in mice. Science. 355, 842-847
73 Mahe, E. A., Madigou, T. and Salbert, G. (2018) Reading cytosine modifications within
chromatin. Transcription. 9, 240-247
74 Mendonca, A., Chang, E. H., Liu, W. and Yuan, C. (2014) Hydroxymethylation of
DNA influences nucleosomal conformation and stability in vitro. Biochim Biophys Acta. 1839,
1323-1329
75 Khrapunov, S., Warren, C., Cheng, H., Berko, E. R., Greally, J. M. and Brenowitz, M.
(2014) Unusual characteristics of the DNA binding domain of epigenetic regulatory protein
MeCP2 determine its binding specificity. Biochemistry. 53, 3379-3391
76 Karemaker, I. D. and Vermeulen, M. (2018) Single-Cell DNA Methylation Profiling:
Technologies and Biological Applications. Trends in Biotechnology. 36, 952-965
77 Mooijman, D., Dey, S. S., Boisset, J. C., Crosetto, N. and van Oudenaarden, A. (2016)
Single-cell 5hmC sequencing reveals chromosome-wide cell-to-cell variability and enables
lineage reconstruction. Nat Biotechnol. 34, 852-856
78 Rudenko, A., Dawlaty, M. M., Seo, J., Cheng, A. W., Meng, J., Le, T., Faull, K. F.,
Jaenisch, R. and Tsai, L. H. (2013) Tet1 is critical for neuronal activity-regulated gene
expression and memory extinction. Neuron. 79, 1109-1122
79 Zhang, R. R., Cui, Q. Y., Murai, K., Lim, Y. C., Smith, Z. D., Jin, S., Ye, P., Rosa, L.,
Lee, Y. K., Wu, H. P., Liu, W., Xu, Z. M., Yang, L., Ding, Y. Q., Tang, F., Meissner, A., Ding,
C., Shi, Y. and Xu, G. L. (2013) Tet1 regulates adult hippocampal neurogenesis and cognition.
Cell Stem Cell. 13, 237-245
80 Ko, M., An, J., Bandukwala, H. S., Chavez, L., Aijö, T., Pastor, W. A., Segal, M. F.,
Li, H., Koh, K. P., Lähdesmäki, H., Hogan, P. G., Aravind, L. and Rao, A. (2013) Modulation
of TET2 expression and 5-methylcytosine oxidation by the CXXC domain protein IDAX.
Nature. 497, 122-126
81 Zhang, W., Xia, W., Wang, Q., Towers, A. J., Chen, J., Gao, R., Zhang, Y., Yen, C. A.,
Lee, A. Y., Li, Y., Zhou, C., Liu, K., Zhang, J., Gu, T. P., Chen, X., Chang, Z., Leung, D., Gao,
S., Jiang, Y. H. and Xie, W. (2016) Isoform Switch of TET1 Regulates DNA Demethylation
and Mouse Development. Molecular cell. 64, 1062-1073
Figure 1. Structure and function of the Mus musculus TET protein family. All three TET
family members display a conserved C-terminal domain that consists of a cysteine-rich region
and a double-stranded β-helix domain (DSBH) domain, which harbours a low complexity
region. The DBSH domain is responsible for iron ion binding and dioxygenease activity of
TET proteins. Additionally, TET1 and TET3 exhibit a CXXC domain that is associated with
binding to CpG dinucleotides. TET2 has lost its CXXC domain due to a chromosomal
inversion, and the ancestral TET2 CXXC domain in mammals is encoded by a separate gene
(Idax). TET proteins can exist in isoforms with or without their associated CXXC domain, and
such isoforms can be tissue- / cell type-specific [14, 80, 81].
Figure 1. Structure and function of the Mus musculus TET protein family. All three TET
family members display a conserved C-terminal domain that consists of a cysteine-rich region
and a double-stranded β-helix domain (DSBH) domain, which harbours a low complexity
region. The DBSH domain is responsible for iron ion binding and dioxygenease activity of
TET proteins. Additionally, TET1 and TET3 exhibit a CXXC domain that is associated with
binding to CpG dinucleotides. TET2 has lost its CXXC domain due to a chromosomal
inversion, and the ancestral TET2 CXXC domain in mammals is encoded by a separate gene
(Idax). TET proteins can exist in isoforms with or without their associated CXXC domain, and
such isoforms can be tissue- / cell type-specific [14, 80, 81].
Figure 2. Tet expression dynamics during embryoid body formation. Expression of Tet1,
Tet2, and Tet3 in mESCs and embryoid bodies at four and eight days of differentiation. The
dashed line tracks global 5hmC levels during this process.
Figure 3. A) TET protein function in vertebrates. Germ layer markers and transcription
factors regulated by Tet1 and Tet3 during embryoid body differentiation [23, 41]. B) A
unifying mechanism of TET protein function in vertebrates. TETs are responsible for
demethylation and activation of regulatory elements associated with key signalling pathways
during gastrulation and body plan formation in diverse vertebrates [41, 60, 67].
... TET enzymes are dynamically expressed during embryogenesis and play essential roles in gene regulation during development and in embryonic stem cell (ESC) biology 8,9 . Tet1 and Tet2 are highly expressed in ESCs, whereas Tet3 is not detectable 10 . ...
... Tet1 and Tet2 are highly expressed in ESCs, whereas Tet3 is not detectable 10 . While Tet1/2 levels decline during ESC differentiation, Tet3 is induced 9,11 . Loss of TET1 or TET2 in ESCs does not block pluripotency but compromises expression of lineage specific genes and ESC differentiation [12][13][14] . ...
Article
Full-text available
The ten-eleven-translocation family of proteins (TET1/2/3) are epigenetic regulators of gene expression. They regulate genes by promoting DNA demethylation (i.e., catalytic activity) and by partnering with regulatory proteins (i.e., non-catalytic functions). Unlike Tet1 and Tet2, Tet3 is not expressed in mouse embryonic stem cells (ESCs) but is induced upon ESC differentiation. However, the significance of its dual roles in lineage specification is less defined. By generating TET3 catalytic-mutant (Tet3m/m) and knockout (Tet3–/–) mouse ESCs and differentiating them to neuroectoderm (NE), we identify distinct catalytic-dependent and independent roles of TET3 in NE specification. We find that the catalytic activity of TET3 is important for activation of neural genes while its non-catalytic functions are involved in suppressing mesodermal programs. Interestingly, the vast majority of differentially methylated regions (DMRs) in Tet3m/m and Tet3–/– NE cells are hypomethylated. The hypo-DMRs are associated to aberrantly upregulated genes while the hyper-DMRs are linked to downregulated neural genes. We find the maintenance methyltransferase Dnmt1 as a direct target of TET3, which is downregulated in TET3-deficient NE cells and may contribute to the increased DNA hypomethylation. Our findings establish that the catalytic-dependent and -independent roles of TET3 have distinct contributions to NE specification with potential implications in development.
... 36 TET enzymes oxidise 5mC into 5-hydroxymethylcytosine, potentially leading to DNA demethylation. 37 To explore the potential roles of DNA demethylase in regulating IGF2BP3 promoter methylation levels in TNBC, we knocked down DNMT1/3A/3B and TET1/2/3 in MDA-MB-231 and HCC-1806 cell lines using specific their siRNAs ( Figure S1a-f). We found that only knockdown of TET3, and not DNMT1/3A/3B and TET1/2, resulted in a decrease in IGF2BP3 expression ( Figures 2D-F and S1g-i). ...
Article
Full-text available
Background N6‐methyladenosine (m6A) is an abundant reversible modification in eukaryotic mRNAs. Emerging evidences indicate that m6A modification plays a vital role in tumourigenesis. As a crucial reader of m6A, IGF2BP3 usually mediates the stabilisation of mRNAs via an m6A‐dependent manner. But the underlying mechanism of IGF2BP3 in the tumourigenesis of triple‐negative breast cancer (TNBC) is unclear. Methods TCGA cohorts were analysed for IGF2BP3 expression and IGF2BP3 promoter methylation levels in different breast cancer subtypes. Colony formation, flow cytometry assays and subcutaneous xenograft were performed to identify the phenotype of IGF2BP3 in TNBC. RNA/RNA immunoprecipitation (RIP)/methylated RNA immunoprecipitation (MeRIP) sequencing and luciferase assays were used to certify the target of IGF2BP3 in TNBC cells. Results IGF2BP3 was highly expressed in TNBC cell lines and tissues. TET3‐mediated IGF2BP3 promoter hypomethylation led to the upregulation of IGF2BP3. Knocking down IGF2BP3 markedly reduced the proliferation of TNBC in vitro and in vivo. Intersection co‐assays revealed that IGF2BP3 decreased neurofibromin 1 (NF1) stabilisation via an m6A‐dependent manner. NF1 knockdown could rescue the phenotypes of IGF2BP3 knockdown cells partially. Conclusion TET3‐mediated IGF2BP3 accelerated the proliferation of TNBC by destabilising NF1 mRNA via an m6A‐dependent manner. This suggests that IGF2BP3 could be a potential therapeutic target for TNBC.
... Because DNA demethylation and the 5hmC mark involve in various biological reactions, TETs play a very important role in both physiological and pathological processes, which have been elucidated by many studies. 13,[30][31][32][33][34][35][36] For example, TET2 loss resulted in hypermutagenicity in haematopoietic progenitor cells, unveiling a key role of TET2 in safeguarding cells against genomic mutagenicity. 37 Dysfunctions of TET2 in cancer are associated with TET2 mutation and abnormal expression of TET2 regulators. ...
Article
Full-text available
Ten-eleven translocation (TET) family proteins (TETs), specifically, TET1, TET2 and TET3, can modify DNA by oxidizing 5-methylcytosine (5mC) iteratively to yield 5-hydroxymethylcytosine (5hmC), 5-formylcytosine (5fC), and 5-carboxycytosine (5caC), and then two of these intermediates (5fC and 5caC) can be excised and return to unmethylated cytosines by thymine-DNA glycosylase (TDG)-mediated base excision repair. Because DNA methylation and demethylation play an important role in numerous biological processes, including zygote formation, embryogenesis, spatial learning and immune homeostasis, the regulation of TETs functions is complicated, and dysregulation of their functions is implicated in many diseases such as myeloid malignancies. In addition, recent studies have demonstrated that TET2 is able to catalyze the hydroxymethylation of RNA to perform post-transcriptional regulation. Notably, catalytic-independent functions of TETs in certain biological contexts have been identified, further highlighting their multifunctional roles. Interestingly, by reactivating the expression of selected target genes, accumulated evidences support the potential therapeutic use of TETs-based DNA methylation editing tools in disorders associated with epigenetic silencing. In this review, we summarize recent key findings in TETs functions, activity regulators at various levels, technological advances in the detection of 5hmC, the main TETs oxidative product, and TETs emerging applications in epigenetic editing. Furthermore, we discuss existing challenges and future directions in this field.
Article
Full-text available
The D2 dopamine receptor (DRD2) gene has garnered substantial attention as one of the most extensively studied genes across various neuropsychiatric disorders. Since its initial association with severe alcoholism in 1990, particularly through the identification of the DRD2 Taq A1 allele, numerous international investigations have been conducted to elucidate its role in different conditions. As of February 22, 2024, there are 5485 articles focusing on the DRD2 gene listed in PUBMED. There have been 120 meta-analyses with mixed results. In our opinion, the primary cause of negative reports regarding the association of various DRD2 gene polymorphisms is the inadequate screening of controls, not adequately eliminating many hidden reward deficiency syndrome behaviors. Moreover, pleiotropic effects of DRD2 variants have been identified in neuropsychologic, neurophysiologic, stress response, social stress defeat, maternal deprivation, and gambling disorder, with epigenetic DNA methylation and histone post-translational negative methylation identified as discussed in this article. There are 70 articles listed in PUBMED for DNA methylation and 20 articles listed for histone methylation as of October 19, 2022. For this commentary, we did not denote DNA and/or histone methylation; instead, we provided a brief summary based on behavioral effects. Based on the fact that Blum and Noble characterized the DRD2 Taq A1 allele as a generalized reward gene and not necessarily specific alcoholism, it now behooves the field to find ways to either use effector moieties to edit the neuroepigenetic insults or possibly harness the idea of potentially removing negative mRNA-reduced expression by inducing “dopamine homeostasis.”
Preprint
Full-text available
The TET family is a ten-eleven translocation family of dioxygenases that oxidize 5-methylcytosine (5mC) to 5-hydroxymethylcytosine (5hmC) and other oxidation products to regulate DNA methylation. Our data revealed significant downregulation of TET1 expression in CRC issues and SW480 cells. The database highlighted mutations as the primary mode of alteration of TET1 in CRC. The bioinformatics analysis results revealed a significant association between TET1 and immune cell infiltration, while indicating that the expression levels of immune checkpoint-related genes in CRC tissues tend to be elevated in comparison to normal tissues. Upon transfection, overexpression of TET1 exerted a comprehensive inhibitory effect by suppressing cell proliferation, inducing apoptosis, hindering migration and invasion, arresting cell cycle progression, and attenuating the activity of the Wnt/β-catenin signaling pathway as well as in nuclear β-catenin expression. Overexpression of TET1 increased 5hmC levels while simultaneously decreasing 5mC levels. We revealed antagonistic genes SFRP2 and WIF1 within the Wnt/β-catenin signaling pathway, which have a significant increase in expression level and a decrease in hypermethylation level upon TET1 overexpression. In conclusion, TET1 exerts its antitumor function by inhibiting the activity of Wnt/β-catenin signaling pathways through demethylation of the antagonistic genes SFRP2 and WIF1. This modulation has a significant impact on the biological properties of CRC.
Article
Background SPRY1 is associated with the invasiveness and prognosis of various tumors, and TET3 affects aging by regulating gene expression. Aims We investigated the roles of SPRY1 and TET3 in natural skin aging, replicative aging, and photoaging, along with the effect of UVA on genome‐wide DNA methylation in HaCaT cells. Methods TET3 and SPRY1 expression were measured in the skin of patients of different age groups, as well as in vitro human skin, HaCaT cell replicative senescence, and HaCaT and HaCaT‐si TET3 cell photoaging models. Senescence was verified using β‐galactosidase staining, and DNA damage was detected using immunofluorescence staining for γ‐H2A.X. 5‐Methyl cytosine (5‐mC) content in the genome was determined using ELISA. Results SPRY1 expression increased with age, whereas TET3 expression decreased. Similarly, SPRY1 was upregulated and TET3 was downregulated with increasing cell passages. TET3‐siRNA upregulated SPRY1 expression in HaCaT cells. UVA irradiation promoted HaCaT cell senescence and induced cellular DNA damage. SPRY1 was upregulated and TET3 was downregulated upon UVA irradiation. Genome‐wide 5‐mC content increased upon TET3 silencing and UVA irradiation, indicating a surge in overall methylation. Conclusions SPRY1 and TET3 are natural skin aging‐related genes that counteract to regulate replicative aging and UVA‐induced photoaging in HaCaT cells. The cell photoaging model may limit experimental bias caused by different exposure times of skin model samples.
Chapter
In recent years, many studies have focused on understanding the effects of genetic and epigenetic mechanisms on carcinogenesis, diagnosing the disease at an early stage, and determining personalized treatment strategies. Epigenetic and genetic alterations are effective in the initiation and progression of cancer, the second most common cause of death worldwide. Epigenetics is defined as heritable changes in gene expression without DNA sequence alterations. Epigenetic mechanisms include DNA methylation, histone modifications, and non-coding RNAs. Disruption of the balance in epigenetic processes, which are necessary for the normal maintenance of tissue-specific gene expression, may cause cancer formation and progression. The reversibility of epigenetic abnormalities is a promising feature for epigenetic cancer therapy studies. This chapter aims to summarize information about epigenetic mechanisms, their role in cancer initiation and progression, and their potential use in cancer therapy.
Chapter
The occurrence of pancreatic cancer (PC) is presented to have risen in the past few years. Pancreatic cancer includes 5% of all cancer-related deaths and almost 2% of existing cancer types. Pancreatic tumors can be categorized as endocrine pancreatic tumors and non-endocrine pancreatic tumors. The significant symptoms will usually not be determined until the advanced metastasis stage. Research in understanding the mechanism of pancreatic cancer focuses on genetic and epigenetic changes using high-throughput genomic sequencing techniques. The epigenetic alterations and genetic abnormalities could lead to tumor progression through increasing oncogene expression, the proliferation of tumor cells, or suppressing tumor suppressor gene expressions with various adaptations. In pancreatic cancer, the progression of tumor metastasis could be related to epigenetic changes, including hypomethylation and hypermethylation of DNA and histone modifications. In today’s world, which is also described as the post-genomic era, the epigenetic foundations of cancer development reveal revolutionary results in cancer genetics and provide the development of promising new methods in cancer treatment. This book chapter discusses epigenetic changes based on methylation, demethylation, and histone modification mechanisms in pancreatic cancer.
Article
Full-text available
The interplay between genetic alterations and metabolic dysregulation is increasingly recognized as a pivotal axis in cancer pathogenesis. Both elements are mutually reinforcing, thereby expediting the ontogeny and progression of malignant neoplasms. Intriguingly, recent findings have highlighted the translocation of metabolites and metabolic enzymes from the cytoplasm into the nuclear compartment, where they appear to be intimately associated with tumor cell proliferation. Despite these advancements, significant gaps persist in our understanding of their specific roles within the nuclear milieu, their modulatory effects on gene transcription and cellular proliferation, and the intricacies of their coordination with the genomic landscape. In this comprehensive review, we endeavor to elucidate the regulatory landscape of metabolic signaling within the nuclear domain, namely nuclear metabolic signaling involving metabolites and metabolic enzymes. We explore the roles and molecular mechanisms through which metabolic flux and enzymatic activity impact critical nuclear processes, including epigenetic modulation, DNA damage repair, and gene expression regulation. In conclusion, we underscore the paramount significance of nuclear metabolic signaling in cancer biology and enumerate potential therapeutic targets, associated pharmacological interventions, and implications for clinical applications. Importantly, these emergent findings not only augment our conceptual understanding of tumoral metabolism but also herald the potential for innovative therapeutic paradigms targeting the metabolism–genome transcriptional axis.
Article
Full-text available
Vertebrates have greatly elaborated the basic chordate body plan and evolved highly distinctive genomes that have been sculpted by two whole-genome duplications. Here we sequence the genome of the Mediterranean amphioxus (Branchiostoma lanceolatum) and characterize DNA methylation, chromatin accessibility, histone modifications and transcriptomes across multiple developmental stages and adult tissues to investigate the evolution of the regulation of the chordate genome. Comparisons with vertebrates identify an intermediate stage in the evolution of differentially methylated enhancers, and a high conservation of gene expression and its cis-regulatory logic between amphioxus and vertebrates that occurs maximally at an earlier mid-embryonic phylotypic period. We analyse regulatory evolution after whole-genome duplications, and find that-in vertebrates-over 80% of broadly expressed gene families with multiple paralogues derived from whole-genome duplications have members that restricted their ancestral expression, and underwent specialization rather than subfunctionalization. Counter-intuitively, paralogues that restricted their expression increased the complexity of their regulatory landscapes. These data pave the way for a better understanding of the regulatory principles that underlie key vertebrate innovations.
Article
Full-text available
The idea that epigenetic determinants such as DNA methylation, histone modifications or RNA can be passed to the next generation through meiotic products (gametes) is long standing. Such meiotic epigenetic inheritance (MEI) is fairly common in yeast, plants and nematodes, but its extent in mammals has been much debated. Advances in genomics techniques are now driving the profiling of germline and zygotic epigenomes, thereby improving our understanding of MEI in diverse species. Whereas the role of DNA methylation in MEI remains unclear, insights from genome-wide studies suggest that a previously underappreciated fraction of mammalian genomes bypass epigenetic reprogramming during development. Notably, intergenerational inheritance of histone modifications, tRNA fragments and microRNAs can affect gene regulation in the offspring. It is important to note that MEI in mammals rarely constitutes transgenerational epigenetic inheritance (TEI), which spans multiple generations. In this Review, we discuss the examples of MEI in mammals, including mammalian epigenome reprogramming, and the molecular mechanisms of MEI in vertebrates in general. We also discuss the implications of the inheritance of histone modifications and small RNA for embryogenesis in metazoans, with a particular focus on insights gained from genome-wide studies.
Article
Full-text available
TET enzymes convert 5-methylcytosine to 5-hydroxymethylcytosine and higher oxidized derivatives. TETs stably associate with and are post-translationally modified by the nutrient-sensing enzyme OGT, suggesting a connection between metabolism and the epigenome. Here, we show for the first time that modification by OGT enhances TET1 activity in vitro. We identify a TET1 domain that is necessary and sufficient for binding to OGT and report a point mutation that disrupts the TET1-OGT interaction. We show that this interaction is necessary for TET1 to rescue hematopoetic stem cell production in tet mutant zebrafish embryos, suggesting that OGT promotes TET1's function during development. Finally, we show that disrupting the TET1-OGT interaction in mouse embryonic stem cells changes the abundance of TET2 and 5-methylcytosine, which is accompanied by alterations in gene expression. These results link metabolism and epigenetic control, which may be relevant to the developmental and disease processes regulated by these two enzymes.
Article
Full-text available
Sin3a is a core component of histone-deacetylation-activity-associated transcriptional repressor complex, playing important roles in early embryo development. Here, we reported that down-regulation of Sin3a led to the loss of embryonic stem cell (ESC) self-renewal and skewed differentiation into mesendoderm lineage. We found that Sin3a functioned as a transcriptional coactivator of the critical Nodal antagonist Lefty1 through interacting with Tet1 to de-methylate the Lefty1 promoter. Further studies showed that two amino acid residues (Phe147, Phe182) in the PAH1 domain of Sin3a are essential for Sin3a-Tet1 interaction and its activity in regulating pluripotency. Furthermore, genome-wide analyses of Sin3a, Tet1 and Pol II ChIP-seq and of 5mC MeDIP-seq revealed that Sin3a acted with Tet1 to facilitate the transcription of a set of their co-target genes. These results link Sin3a to epigenetic DNA modifications in transcriptional activation and have implications for understanding mechanisms underlying versatile functions of Sin3a in mouse ESCs.
Article
Full-text available
Gametes are highly specialized cells that can give rise to the next generation through their ability to generate a totipotent zygote. In mice, germ cells are first specified in the developing embryo around embryonic day (E) 6.25 as primordial germ cells (PGCs). Following subsequent migration into the developing gonad, PGCs undergo a wave of extensive epigenetic reprogramming around E10.5-E11.5, including genome-wide loss of 5-methylcytosine. The underlying molecular mechanisms of this process have remained unclear, leading to our inability to recapitulate this step of germline development in vitro. Here we show, using an integrative approach, that this complex reprogramming process involves coordinated interplay among promoter sequence characteristics, DNA (de)methylation, the polycomb (PRC1) complex and both DNA demethylation-dependent and -independent functions of TET1 to enable the activation of a critical set of germline reprogramming-responsive genes involved in gamete generation and meiosis. Our results also reveal an unexpected role for TET1 in maintaining but not driving DNA demethylation in gonadal PGCs. Collectively, our work uncovers a fundamental biological role for gonadal germline reprogramming and identifies the epigenetic principles of the PGC-to-gonocyte transition that will help to guide attempts to recapitulate complete gametogenesis in vitro.
Article
Full-text available
Discovery of the ten-eleven translocation 1 (TET) methylcytosine dioxygenase family of enzymes, nearly 10 years ago, heralded a major breakthrough in understanding the epigenetic modifications of DNA. Initially described as catalyzing the oxidation of methyl cytosine (5mC) to hydroxymethyl cytosine (5hmC), it is now clear that these enzymes can also catalyze additional reactions leading to active DNA demethylation. The association of TET enzymes, as well as the 5hmC, with active regulatory regions of the genome has been studied extensively in embryonic stem cells, although these enzymes are expressed widely also in differentiated tissues. However, TET1 and TET3 are found as various isoforms, as a result of utilizing alternative regulatory regions in distinct tissues. Some of these isoforms, like TET2, lack the CXXC domain which probably has major implications on their recruitment to specific loci in the genome, while in certain contexts TET1 is seen paradoxically to repress transcription. In this review we bring together these novel aspects of the differential regulation of these Tet isoforms and the likely consequences on their activity.
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.
Article
(Cell Stem Cell 23, 727–741.e1–e9, November 1, 2018) In the initial online and print versions of our manuscript, we mistakenly mislabeled the heatmaps in Figure 6B. The original and corrected versions of Figure 6B appear below, and the online version of our manuscript now contains the correct labeling. We apologize for the confusion. [Figure presented]
Article
Here, we report DNA methylation and hydroxymethylation dynamics at nucleotide resolution using C/EBPα-enhanced reprogramming of B cells into induced pluripotent cells (iPSCs). We observed successive waves of hydroxymethylation at enhancers, concomitant with a decrease in DNA methylation, suggesting active demethylation. Consistent with this finding, ablation of the DNA demethylase Tet2 almost completely abolishes reprogramming. C/EBPα, Klf4, and Tfcp2l1 each interact with Tet2 and recruit the enzyme to specific DNA sites. During reprogramming, some of these sites maintain high levels of 5hmC, and enhancers and promoters of key pluripotency factors become demethylated as early as 1 day after Yamanaka factor induction. Surprisingly, methylation changes precede chromatin opening in distinct chromatin regions, including Klf4 bound sites, revealing a pioneer factor activity associated with alternation in DNA methylation. Rapid changes in hydroxymethylation similar to those in B cells were also observed during compound-accelerated reprogramming of fibroblasts into iPSCs, highlighting the generality of our observations.
Article
DNA methylation is an epigenetic modification that plays an important role in gene expression regulation, development, and disease. Recent technological innovations have spurred the development of methods that enable us to study the occurrence and biology of this mark at the single-cell level. Apart from answering fundamental biological questions about heterogeneous systems or rare cell types, low-input methods also bring clinical applications within reach. Ultimately, integrating these data with other single-cell data sets will allow deciphering multiple layers of gene expression regulation within each individual cell. Here, we review the approaches that have been developed to facilitate single-cell DNA methylation profiling, their biological applications, and how these will further our understanding of the biology of DNA methylation.