Skip to main content

Decoding the role of TET family dioxygenases in lineage specification

Abstract

Since the discovery of methylcytosine oxidase ten-eleven translocation (TET) proteins, we have witnessed an exponential increase in studies examining their roles in epigenetic regulation. TET family proteins catalyze the sequential oxidation of 5-methylcytosine (5mC) to oxidized methylcytosines including 5-hydroxymethylcytosine (5hmC), 5-formylcytosine, and 5-carboxylcytosine. TETs contribute to the regulation of lineage-specific gene expression via modulating DNA 5mC/5hmC balances at the proximal and distal regulatory elements of cell identity genes, and therefore enhance chromatin accessibility and gene transcription. Emerging evidence suggests that TET dioxygenases participate in the establishment and/or maintenance of hypomethylated bivalent domains at multiple differentiation-associated genes, and thus ensure developmental plasticity. Here, we review the current state of knowledge concerning TET family proteins, DNA hydroxymethylation, their distribution, and function in endoderm, mesoderm, and neuroectoderm specification. We will summarize the evidence pertaining to their crucial regulatory roles in lineage commitment and development.

Background

DNA methylation and the recently identified hydroxymethylation are essential epigenetic modifications in cells. DNA methylation is catalyzed by DNA cytosine-5-methyltransferases (DNMTs) via transferring a methyl group to the 5′ position of cytosine to form 5-methylcytosine (5mC). Heavily methylated DNA is often associated with repressed gene expression. In addition, DNA methylation can be actively demethylated by DNA methylcytosine dioxygenases, the ten-eleven translocation (TET) proteins, through oxidizing 5mC into 5-hydroxymethylcytosine (5hmC) and further to 5-formylcytosine (5fC) and then to 5-carboxylcytosine (5caC). Finally, 5fC and 5caC are removed by thymine DNA glycosylase (TDG) and cytosine is replaced by base excision repair (BER) [1,2,3,4,5]. Successful oxidation of 5mC by TET proteins influences various biological properties such as chromatin accessibility, nucleosome positioning, genomic stability, and rates of gene transcription [6,7,8,9]. TET proteins hence serve as important epigenetic modifiers that participate in a number of biological processes including embryogenesis, lineage specification, and disease development. Herein, we specifically review how TET proteins and their enzymatic products contribute to the regulation of cell lineage commitment and development.

Structural basis of TET family proteins

TET proteins are widely expressed in various organisms including human, mouse, Xenopus, and zebrafish [6, 10,11,12,13,14]. The three TET family members (TET1, TET2, and TET3) share a highly conserved catalytic domain at the C-termini, which comprises cysteine-rich and double-stranded β-helix (DSBH) regions [15, 16]. The DSBH region contains ferrous (Fe2+) and α-ketoglutarate (α-KG) binding sites that are critical for TET catalytic activity [17, 18]. TET1 and TET3 both possess a zinc finger cysteine-X-X-cysteine (CXXC) domain at the N-termini, which allows them to bind to cytosine and its modified forms (e.g., 5mC, 5hmC, 5fC, and 5caC) in DNA [6, 19, 20]. Although TET2 does not encode the CXXC domain, it targets deoxynucleotides through another CXXC-containing protein known as IDAX or CXXC4. Therefore, TET2 appears to bind to DNA in a similar fashion as TET1 and TET3 [21]. Recently, a CXXC-domain deficient short form of TET1 (TET1S) has been identified in both mouse and human somatic cells [22, 23]. TET1S retains reduced catalytic activity of TET dioxygenase and displays a weaker binding affinity for DNA [22]. Despite their differences in DNA binding properties, these TET isoforms are comparable in terms of 5mC oxidation capability.

Dynamic distribution of TETs and 5hmC during development

The three TET family members have distinct expression patterns among different cell types and are tightly regulated during development [24,25,26]. TET1 protein is highly expressed in both human and mouse embryonic stem cells (ESCs), whereas TET2 is expressed at extremely low levels in human ESCs similar to TET3 in mouse ESCs [26,27,28]. The expression levels of TET proteins change dynamically during development. For instance, TET3 is expressed at high levels in oocytes and zygotes, and undergoes rapid downregulation in two-cell-stage embryos [29]. It has been shown that TET3 expression increases dramatically in human ESC-derived neuroectoderm and pancreatic endoderm [24, 28], which is consistent with its progressive increase in mouse embryos from e6.5 to e9.5 [30]. In contrast, TET2 is widely expressed in a variety of somatic organs and cell types, especially in hematopoietic cells [31]. More interestingly, recent studies have illustrated that two TET1 isoforms also display distinct expression patterns, in which the full-length isoforms of TET1 are preferentially expressed in ESCs, early embryos, and primordial germ cells (PGC), while the short form of TET1 (TET1S) is restricted to somite cells and overexpressed in cancer [22, 23, 32]. An isoform switch of TET1 has been implicated in influencing gametic imprinting, PGC development, and epigenetic memory erasure [22].

Since the discovery of TET family dioxygenases, studies examining the function of DNA hydroxymethylation have exponentially increased. Emerging evidence indicates that 5hmCs are enriched at low-to-intermediate CpG density regions of promoters and enhancers of developmental regulatory genes [33,34,35,36,37,38,39,40,41,42]. Similar to TET family proteins, the distribution of 5hmC is dynamically changed and positively correlated with active gene transcription during lineage specification [35, 43, 44]. For example, high levels of 5hmC are found in ESCs and in the central nervous system [16, 45]. Global 5hmC levels decrease during ESCs differentiation toward neuroectoderm fate, while enrichment of 5hmC at the gene body of transcriptionally active genes is identified in neural progenitor cells (NPCs) [43, 44]. When NPCs further differentiate into neurons, overall 5hmC increases are accompanied by a loss of H3K27me3 at promoters of genes that are important for neuronal function [34]. Likewise, a human ESC-based model of pancreatic differentiation reveals that global 5hmC levels rapidly decrease at the first step toward definitive endoderm, then gradually increase toward pancreatic endoderm specification. 5hmC enriching peaks significantly overlap with poised and active enhancers, as well as the boundaries of hypomethylated functional genomic regions [28]. Furthermore, enrichment of 5hmC at tissue-specific enhancers has also been demonstrated in cardiomyocyte and hematopoietic cell differentiation [10, 35, 46]. Other than at promoters and enhancers, enrichment of 5hmC is also found over the gene bodies of actively transcribed genes [35, 36, 47]. Taken together, these studies suggest that expression of TET family proteins and distribution of 5hmC is differentially regulated to meet the needs of cellular functions during lineage commitment.

Mechanisms of TET family proteins in the regulation of gene transcription

The establishment of cellular identities requires precise control of gene expression. Mounting evidence suggests that TET proteins work as epigenetic players to alter DNA methylation, histone modification, and chromatin accessibility, which regulate the transcription of key developmental genes. Strikingly, promoters, enhancers, and DNase I-hypersensitive sites accumulate significantly more 5mC in Tet1, Tet2, and Tet3 triple knockout mouse ESCs, suggesting that these gene regulatory regions are the major targets of TETs [13, 48, 49]. In the following subsections, we discuss the current mechanistic understanding of how TET proteins regulate lineage-specific gene expression.

Maintenance of hypomethylated promoters of developmental genes

Accumulated evidence indicates that DNA methylation status at promoter regions influences gene transcription [50]. Promoter hypermethylation is believed to contribute to the establishment of a transcriptionally poised/inactive state [50]. For example, pluripotent genes, such as OCT4, are actively expressed in human ESCs and suppressed by promoter hypermethylation upon differentiation [44]. It has been widely documented that genetic ablation of TET induces promoter hypermethylation and aberrant gene expression in multiple lineage differentiation systems [6, 25, 27, 38, 51,52,53]. For example, TET2 is required for the maintenance of NANOG expression in the spontaneous differentiation of mesodermal lineage cells from human ESCs. TET2 ChIP-seq revealed that TET2 associated with the NANOG promoter prevents DNA methylation [27]. In helper T cell differentiation, Tet2 is recruited to the promoters of cytokine genes in a lineage-specific transcriptional factor-dependent manner, which stimulates active DNA demethylation and expression of these cytokine genes [38]. TET1 and TET3 also regulate DNA methylation status in the promoter regions. For instance, Tet3 directly binds to the promoters of genes critical for neural development in Xenopus, such as Pax6, Rx, and Ngn2, and sustains high levels of 5hmC at promoters [6]. Furthermore, simultaneous deletion of Tet2 and Tet3 downregulates P2rX7 expression along with reduction of 5hmC at the P2rX7 promoter during bone marrow mesenchymal stem cell differentiation [53], indicating that TET-mediated rapid and specific oxidation of 5mC at promoter loci is biologically relevant.

More interestingly, TET-mediated DNA demethylation has been suggested in association with the establishment and/or maintenance of bivalent promoters of developmental genes, in which H3K4me3 and H3K27me3 histone modifications take place simultaneously [14, 54]. In general, bivalent promoters of developmental genes are hypomethylated in ESCs [55]. They are preferentially repressed by trimethylation of histone H3 at lysine 27, which is easier to be reversed than DNA methylation. These low-methylated genetic loci can extend beyond promoter regions, forming H3K27me3-marked DNA methylation valleys (DMVs) that provide binding sites for a large set of transcription factors to mediate complex regulation during development [56]. It has been recently shown that TET proteins and Polycomb Repressive Complex 2 (PRC2), which is responsible for H3K27 methylation, can recruit each other to maintain a hypomethylated status at bivalent promoters and DMVs [57,58,59]. In addition, TETs can interact with OGT (O-linked β-N-acetylglucosamine transferase), which enhances methylation of histone H3 at lysine 4 by promoting the binding of a component of the H3K4 methyltransferase SET1/COMPASS complex to active promoters [60]. It has been further demonstrated that overexpression of TET2, but not the catalytically inactive TET2, results in an increase in 5hmC at a particular set of key developmental gene promoters, which is sufficient to promote DNA demethylation and de novo bivalent modifications [54]. This discovery was subsequently supported by a recent report which illustrated that TETs safeguard bivalent promoters of many lineage determinants, such as FOXA2, GATA2, PAX6, and SOX17, by preventing their aberrant hypermethylation to ensure developmental competency [14]. Together, these studies suggest that TET-mediated DNA demethylation retains developmental plasticity at the promoters and/or DMVs of developmentally important genes, and therefore ensure robust induction of lineage-specific transcription upon differentiation (Fig. 1).

Fig. 1
figure 1

The role of TET proteins on lineage-specific bivalent promoters and enhancers. a In the presence of TET dioxygenases, PRC2 recruits TETs to bivalent promoters to maintain their hypomethylated status. In the absence of TETs, binding of DNMT3B at the bivalent promoters causes de novo DNA methylation, which leads to stable gene silencing and loss of developmental plasticity. b A model of TET-mediated enhancer priming and activation. Upon differentiation, pioneer transcription factors that are not sensitive to DNA modifications can bind to distal enhancers of lineage-specific genes and recruit TETs to demethylate methylcytosines. Other epigenetic modifiers, such as p300 and SET1/COMPASS, subsequently bind to these sites and establish poised (H3K4me1) and active (H3K27ac) enhancers, which in turn increases chromatin accessibility and allow other transcription factors binding to occur

Regulation of chromatin accessibility and enhancer architecture

In addition to facilitate promoter hypomethylation, TET proteins also play critical roles in regulating chromatin accessibility and enhancer architecture (Fig. 1). 5hmC co-localizes with enhancers and open chromosome regions during different biological processes, such as B cell differentiation and pancreatic endocrine differentiation [7, 28, 61]. By examining genome-wide DNA methylation and hydroxymethylation in the context of Tet2 deletion in mESCs, Ren and colleagues have found that depletion of Tet2 leads to enhancer hypermethylation, accompanied by the loss of active enhancer mark H3K27ac and delayed the induction of Slit3, Lmo4, and Irx3 upon differentiation to a neural progenitor fate [62]. A similar observation has been made in cardiomyocytes where deletion of Tet2 causes loss of 5hmC at enhancers and is accompanied by extensive elevation of DNA methylation, reduction of H3K27ac, and impaired gene expression during heart development [10]. Re-expression of Tet2 catalytic domain in Tet2/3 double knockout pro-B cells restores chromatin accessibility at a genome-wide level as well as at the Igκ enhancer [7]. In agreement with the above studies, application of TET inhibitor dimethyloxalylglycine (DMOG) reduces chromatin accessibility at specific enhancers of P19 embryonic carcinoma cells when differentiated to NPCs [63]. These data implicate the functional relevance of TETs and TET-mediated 5hmC to chromatin accessibility at distal regulatory elements, particularly enhancers. Although the precise molecular mechanisms remain unclear, epigenetic readers of 5hmC, such as MeCP2 (methyl-CpG-binding protein 2), might constitute a mechanism of TET-regulated chromatin opening [64].

In addition to chromatin accessibility, TETs may also contribute to enhancer priming. Prior to activation, the enhancer region is methylated and not accessible to general transcription factors. However, pioneer transcription factors, such as FOXA1, MEIS1, and PBX1, preferentially bind to oligonucleotide probes which contain methylated cytosines [63, 65]. It has also been suggested that pioneer transcription factors can physically interact with TETs which catalyze oxidation of 5mC at gene regulatory regions [7, 66, 67]. For example, pioneer transcription factors, PU.1, recruit Tet2 proteins to the Igκ enhancer to facilitate DNA demethylation during early B cell maturation [7]. Furthermore, removal of 5mC or deposition of 5hmC coincides with increased accessibility of enhancers and monomethylation of H3K4 [12, 28, 63]. Accordingly, H3K4 methyltransferases are repelled by 5mC [68, 69], indicating that TET-mediated DNA demethylation is necessary for the recruitment of H3K4 methyltransferases to prime enhancers. Additionally, it has been demonstrated that TETs can recruit SET1/COMPASS H3K4 methyltransferase as well as histone acetyltransferase p300 to gene regulatory regions [38, 60]. Therefore, H3K4 monomethyltransferases MLL3/4, which are COMPASS family members, might also interact with TETs and become recruited to enhancer regions [70].

Despite their plausible contribution to chromatin architecture, TETs may regulate gene transcription by interfering with RNA polymerization and RNA splicing. In CD4+ T cells, TET1 and TET2 alter CTCF-dependent alternative pre-mRNA splicing through oxidation of 5mC to 5hmC and 5caC at corresponding intragenic CTCF-binding sites in the PTPRC (protein tyrosine phosphatase CD45) locus [71, 72]. Additionally, it has been suggested that oxidation derivatives of 5hmC are concentrated on the gene bodies of transcribed genes and support transcriptional consistency [35, 47].

TET dioxygenases modulate cell fate commitment

Growing evidence confirms that TET family proteins are important regulators for embryogenesis. Tet1/2/3 triple knockout mice are embryonic lethal with severe gastrulation defects during embryogenesis [13, 30, 48]. Although the formation of three germ layers is initiated, Tet-null mouse embryos cannot further develop, accompanied by impaired patterning of axial mesoderm, neuroectoderm, and definitive endoderm [13, 30]. Furthermore, knockdown of TET2 skews spontaneous differentiation of human ESCs into neuroectoderm with the loss of mesoderm and endoderm [27], while inactivation of Tet2 in mouse hematopoietic stem cells leads to an increase in the granulocytic and monocytic population [31]. Thus, deficiency of TET proteins disturbs the 5mC and 5hmC landscapes, which causes differentiation to switch from one to another lineage [30, 61, 73, 74]. These studies illustrate the critical function of TET family proteins in development. Below, we will discuss the roles of TET proteins in each lineage commitment (Table 1).

Table 1 Phenotypes resulting from the depletion of TETs in ectoderm, mesoderm, and endoderm lineages

Neuroectoderm lineage specification

The brain is one of the places in mammals with the most abundant 5hmC, suggesting that TET proteins and TET-mediated 5hmC might have significant impacts in neurogenesis. Tet1 knockout mice exhibit defects in neuron function including learning, memory consolidation, storage, and extinction [52, 75, 76]. Ablation of Tet1 leads to downregulation of neuronal activity-regulated genes such as Npas4 [75]. Although Tet1 is highly expressed in multiple regions and cells of the brain like hippocampus, isocortex, and cerebellar granule cells [77, 78], Tet1 knockout does not alter brain morphology. It has been revealed that compensatory upregulations of Tet2 and Tet3 were observed in the Tet1 knockout mouse brain [76]. In addition, overexpression of Tet3 can facilitate the reprogramming of MEFs (mouse embryo fibroblasts) into neuronal cells, which is accompanied with an active demethylation process at the promoters of genes encoding neuron-specific transcription factors such Ascl1, Brn2, and Ngn2 [79]. Therefore, TET family members are functionally redundant in neurogenesis. In the absence of Tet1, other Tet family members can compensate for the activity of Tet1 to regulate neurogenesis.

A recent study from Huangfu’s group has nicely demonstrated that TET1/2/3 triple knockout human ESCs lose their ability to differentiate into PAX6+/SOX1+ neuroectoderm cells, which is mainly caused by aberrant hypermethylation at the PAX6 promoter [14]. In human ESCs, TET1 binding at PAX6 bivalent promoter leads to the progressive oxidation of 5-methylcytosine. In the absence of 5hmC, de novo DNA methyltransferase DNMT3B anchors at PAX6 bivalent promoter P0 and induces promoter hypermethylation, which suppresses PAX6 induction and subsequent neuroectoderm differentiation. Targeted demethylation of the PAX6 P0 promoter by a catalytically inactive Cas9 (dCas9) fused with a TET1 catalytic domain partially restored PAX6 expression and rescued neuron differentiation defects [14]. These results clearly illustrate that TET-mediated hydroxymethylation prevents repressive DNA methylation and ensures key lineage-specific transcription factor expression in neuron differentiation.

Mesoderm lineage specification

TET2 is ubiquitously expressed in multiple hematopoietic cells, and its mutation is frequently found in hematological malignancies [31, 80, 81]. It has been well documented that TET2 is critical for hematopoiesis. In patients who develop chronic myelomonocytic leukemia and carry TET2 mutations in CD34+ hematopoietic progenitor cells, the TET2-mutated CD34+ progenitors preferentially develop into myeloid instead of erythroid cells upon differentiation [82]. Furthermore, it has been shown that single deletion of Tet2 or double deletion of Tet2/3 or Tet1/2 alters T cell, B cell, NKT cell, red blood cell, or mast cell differentiation and maturation [38, 61, 74, 83,84,85]. For instance, genetic depletion of Tet2 perturbs the induction of signature cytokine genes upon differentiation of CD4+ T cells toward helper T (Th) cells, which is associated with differential enrichment of 5hmC and p300 at the promoters of Ifng and Il17 [38]. In contrast, Tet1/Tet2 or Tet2/Tet3 double knockout mice contain less regulatory T (Treg) cells in the spleen. It has been further shown that Tet proteins can demethylate the conserved noncoding sequences (CNS) in Foxp3 loci to maintain the expression of Foxp3 [83, 86]. One of the Foxp3 CNS functions as a super enhancer and docking site for chromatin organizer Satb1 binding [87].

Moreover, TET dioxygenases have also been implicated to play crucial roles during cardiomyocytes and skeletal myoblast differentiation [10, 51, 88]. Knockdown of Tet2 in undifferentiated C2C12 myoblasts significantly reduces the induction of myoblast differentiation-associated genes Myog and myoM with elevated methylation at their promoters [88]. Additionally, TET2 is shown to control the expression of contractile genes, such as SRF, MYOCD, and MYH11, in human coronary artery smooth muscle cells [51]. Similarly, loss of gene expression in smooth muscle cells is correlated with a significant decrease in TET2 binding and 5hmC levels at the promoters of corresponding genes. In cardiomyocyte differentiation, Tet2 knockdown deregulates a large number of genes associated with heart development and contraction. In particular, induction of the key cardiac gene Myh7 is suppressed upon cardiomyocyte differentiation in Tet2 knockdown cells, presumably due to the aberrant methylation at its enhancer [10].

Endoderm lineage specification

To date, functional analyses of TETs and 5hmC in the context of endodermal lineage specification are still limited. In the intestine, Tet1 is highly expressed in intestinal stem cells and positively regulates the expression of Wnt target genes such as Axin2 and Lgr5 [25]. In a model of hepatocyte differentiation from the human hepatic progenitor HepaRG cells, Hernandez-Vargas and colleagues demonstrated that TET1 dioxygenase is necessary for HNF4A promoter P1 demethylation and activation upon hepatocyte differentiation [89]. They found that TET1 binds to the P1 locus via the pioneer transcription factor FOXA2, which is required for the establishment of the hepatocyte program [89]. Additionally, to understand the precise role of TET-mediated hydroxymethylation during pancreatic lineage progression, we recently characterized each lineage intermediate, including definitive endoderm, primitive gut tube, posterior foregut, and pancreatic progenitors, in great detail during their stepwise differentiation from human ESCs [28]. We developed genome-wide maps for each stage, encompassing 5mC/5hmC, gene expression, and chromatin architecture/accessibility. We identified that 5hmC is positively correlated with enhancer activities and chromatin accessibility during pancreatic differentiation [28], and further discovered that TET dioxygenases promoted pancreatic endocrine differentiation but had less effect on the early endoderm formation (unpublished results).

Conclusion and perspectives

In summary, TETs and TET-mediated 5hmC are dynamically redistributed in lineage descendants during development. A substantial number of studies have provided data supporting the role of TET proteins in the establishment of cell identity during differentiation. TET family members alter lineage specification as epigenetic modifiers through their dioxygenase activity to convert 5mC to oxidized methylcytosine. They modulate DNA methylation/hydroxymethylation balance at promoters and enhancers of cell fate-determining genes to further increase chromatin accessibility and facilitate gene transcription in lineage commitment. Strikingly, TET dioxygenases participate in the establishment and/or maintenance of bivalent domains of many differentiation-associated genes, and thusly ensure developmental plasticity. However, direct connections between TET dioxygenases and chromatin architecture are still not clear. Although 5hmC is enriched at distal regulatory elements, how the poised and active enhancer states are influenced by TETs remains elusive. It will be of great interest to identify whether oxidized methylcytosines can act as docking sites and recruit epigenetic readers to modulate histone modifications. Further studies aimed at unraveling the precise role of TET dioxygenases in the control of epigenetic machinery and gene regulation will contribute to the knowledge of how lineage specification is precisely regulated during development.

Abbreviations

TET:

ten-eleven translocation protein

5mC:

5-methylcytosine

5hmC:

5-hydroxymethylcytosine

5fC:

5-formylcytosine

5caC:

5-carboxylcytosine

DNMT:

DNA methyltransferase

TDG:

thymine DNA glycosylase

BER:

base excision repair

DSBH:

double-stranded β-helix

CXXC:

cysteine-X-X-cysteine

ESCs:

embryonic stem cells

NPCs:

neural progenitor cells

DMVs:

DNA methylation valleys

PRC2:

Polycomb Repressive Complex 2

OGT:

O-linked β-N-acetylglucosamine transferase

DMOG:

dimethyloxalylglycine

H3K27ac:

acetylation of histone H3 at lysine 27

H3K4me1:

monomethylation of histone H3 at lysine 4

H3K4me3:

trimethylation of histone H3 at lysine 4

H3K27me3:

trimethylation of histone H3 at lysine 27

References

  1. Maiti A, Drohat AC. Thymine DNA glycosylase can rapidly excise 5-formylcytosine and 5-carboxylcytosine: potential implications for active demethylation of CpG sites. J Biol Chem. 2011;286:35334–8. https://doi.org/10.1074/jbc.C111.284620.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Zhang L, Lu X, Lu J, Liang H, Dai Q, Xu GL, et al. Thymine DNA glycosylase specifically recognizes 5-carboxylcytosine-modified DNA. Nat Chem Biol. 2012;8:328–30. https://doi.org/10.1038/nchembio.914.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Pastor WA, Aravind L, Rao A. TETonic shift: biological roles of TET proteins in DNA demethylation and transcription. Nat Rev Mol Cell Biol. 2013;14:341–56. https://doi.org/10.1038/nrm3589.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Williams K, Christensen J, Pedersen MT, Johansen JV, Cloos PA, Rappsilber J, et al. TET1 and hydroxymethylcytosine in transcription and DNA methylation fidelity. Nature. 2011;473:343–8. https://doi.org/10.1038/nature10066.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. He YF, Li BZ, Li Z, Liu P, Wang Y, Tang Q, et al. Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science. 2011;333:1303–7. https://doi.org/10.1126/science.1210944.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Xu Y, Xu C, Kato A, Tempel W, Abreu JG, Bian C, et al. Tet3 CXXC domain and dioxygenase activity cooperatively regulate key genes for Xenopus eye and neural development. Cell. 2012;151:1200–13. https://doi.org/10.1016/j.cell.2012.11.014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Lio CW, Zhang J, Gonzalez-Avalos E, Hogan PG, Chang X, Rao A. Tet2 and Tet3 cooperate with B-lineage transcription factors to regulate DNA modification and chromatin accessibility. Elife. 2016. https://doi.org/10.7554/elife.18290.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Teif VB, Beshnova DA, Vainshtein Y, Marth C, Mallm JP, Hofer T, et al. Nucleosome repositioning links DNA (de)methylation and differential CTCF binding during stem cell development. Genome Res. 2014;24:1285–95. https://doi.org/10.1101/gr.164418.113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Kafer GR, Li X, Horii T, Suetake I, Tajima S, Hatada I, et al. 5-Hydroxymethylcytosine marks sites of DNA damage and promotes genome stability. Cell Rep. 2016;14:1283–92. https://doi.org/10.1016/j.celrep.2016.01.035.

    Article  CAS  PubMed  Google Scholar 

  10. Greco CM, Kunderfranco P, Rubino M, Larcher V, Carullo P, Anselmo A, et al. DNA hydroxymethylation controls cardiomyocyte gene expression in development and hypertrophy. Nat Commun. 2016;7:12418. https://doi.org/10.1038/ncomms12418.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Ge L, Zhang RP, Wan F, Guo DY, Wang P, Xiang LX, et al. TET2 plays an essential role in erythropoiesis by regulating lineage-specific genes via DNA oxidative demethylation in a zebrafish model. Mol Cell Biol. 2014;34:989–1002. https://doi.org/10.1128/MCB.01061-13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Bogdanovic O, Smits AH, Mustienes ED, Tena JJ, Ford E, Williams R, et al. Active DNA demethylation at enhancers during the vertebrate phylotypic period. Nat Genet. 2016;48:417–26. https://doi.org/10.1038/ng.3522.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Dai HQ, Wang BA, Yang L, Chen JJ, Zhu GC, Sun ML, et al. TET-mediated DNA demethylation controls gastrulation by regulating Lefty-Nodal signalling. Nature. 2016;538:528–32. https://doi.org/10.1038/nature20095.

    Article  CAS  PubMed  Google Scholar 

  14. Verma N, Pan H, Dore LC, Shukla A, Li QV, Pelham-Webb B, et al. TET proteins safeguard bivalent promoters from de novo methylation in human embryonic stem cells. Nat Genet. 2018;50:83–95. https://doi.org/10.1038/s41588-017-0002-y.

    Article  CAS  PubMed  Google Scholar 

  15. Iyer LM, Tahiliani M, Rao A, Aravind L. Prediction of novel families of enzymes involved in oxidative and other complex modifications of bases in nucleic acids. Cell Cycle. 2009;8:1698–710. https://doi.org/10.4161/cc.8.11.8580.

    Article  CAS  PubMed  Google Scholar 

  16. Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H, Brudno Y, et al. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science. 2009;324:930–5. https://doi.org/10.1126/science.1170116.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Yin R, Mo J, Dai J, Wang H. Nickel(II) inhibits tet-mediated 5-methylcytosine oxidation by high affinity displacement of the cofactor iron(II). ACS Chem Biol. 2017;12:1494–8. https://doi.org/10.1021/acschembio.7b00261.

    Article  CAS  PubMed  Google Scholar 

  18. Xiao MT, Yang H, Xu W, Ma SH, Lin HP, Zhu HG, et al. Inhibition of alpha-KG-dependent histone and DNA demethylases by fumarate and succinate that are accumulated in mutations of FH and SDH tumor suppressors. Gene Dev. 2012;26:1326–38. https://doi.org/10.1101/gad.191056.112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Xu Y, Wu F, Tan L, Kong L, Xiong L, Deng J, et al. Genome-wide regulation of 5hmC, 5mC, and gene expression by Tet1 hydroxylase in mouse embryonic stem cells. Mol Cell. 2011;42:451–64. https://doi.org/10.1016/j.molcel.2011.04.005.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Jin SG, Zhang ZM, Dunwell TL, Harter MR, Wu X, Johnson J, et al. Tet3 reads 5-carboxylcytosine through its CXXC domain and is a potential guardian against neurodegeneration. Cell Rep. 2016;14:493–505. https://doi.org/10.1016/j.celrep.2015.12.044.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Ko M, An J, Bandukwala HS, Chavez L, Aijo T, Pastor WA, et al. Modulation of TET2 expression and 5-methylcytosine oxidation by the CXXC domain protein IDAX. Nature. 2013;497:122–6. https://doi.org/10.1038/nature12052.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Zhang W, Xia W, Wang Q, Towers AJ, Chen J, Gao R, et al. Isoform switch of TET1 regulates DNA demethylation and mouse development. Mol Cell. 2016;64:1062–73. https://doi.org/10.1016/j.molcel.2016.10.030.

    Article  CAS  PubMed  Google Scholar 

  23. Good CR, Madzo J, Patel B, Maegawa S, Engel N, Jelinek J, et al. A novel isoform of TET1 that lacks a CXXC domain is overexpressed in cancer. Nucleic Acids Res. 2017;45:8269–81. https://doi.org/10.1093/nar/gkx435.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Li T, Yang D, Li J, Tang Y, Yang J, Le W. Critical role of Tet3 in neural progenitor cell maintenance and terminal differentiation. Mol Neurobiol. 2015;51:142–54. https://doi.org/10.1007/s12035-014-8734-5.

    Article  CAS  PubMed  Google Scholar 

  25. Kim R, Sheaffer KL, Choi I, Won KJ, Kaestner KH. Epigenetic regulation of intestinal stem cells by Tet1-mediated DNA hydroxymethylation. Genes Dev. 2016;30:2433–42. https://doi.org/10.1101/gad.288035.116.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Koh KP, Yabuuchi A, Rao S, Huang Y, Cunniff K, Nardone J, et al. Tet1 and Tet2 regulate 5-hydroxymethylcytosine production and cell lineage specification in mouse embryonic stem cells. Cell Stem Cell. 2011;8:200–13. https://doi.org/10.1016/j.stem.2011.01.008.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Langlois T, daCostaReis M-MB, Lenglet G, Droin N, Marty C, LeCouedic JP, et al. TET2 deficiency inhibits mesoderm and hematopoietic differentiation in human embryonic stem cells. Stem Cells. 2014;32:2084–97. https://doi.org/10.1002/stem.1718.

    Article  CAS  PubMed  Google Scholar 

  28. Li J, Wu X, Zhou Y, Lee M, Guo L, Han W, et al. Decoding the dynamic DNA methylation and hydroxymethylation landscapes in endodermal lineage intermediates during pancreatic differentiation of hESC. Nucleic Acids Res. 2018;46:2883–900. https://doi.org/10.1093/nar/gky063.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Iqbal K, Jin SG, Pfeifer GP, Szabo PE. Reprogramming of the paternal genome upon fertilization involves genome-wide oxidation of 5-methylcytosine. Proc Natl Acad Sci USA. 2011;108:3642–7. https://doi.org/10.1073/pnas.1014033108.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Li X, Yue XJ, Pastor WA, Lin LZ, Georges R, Chavez L, et al. Tet proteins influence the balance between neuroectodermal and mesodermal fate choice by inhibiting Wnt signaling. Proc Natl Acad Sci USA. 2016;113:E8267–76. https://doi.org/10.1073/pnas.1617802113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Moran-Crusio K, Reavie L, Shih A, Abdel-Wahab O, Ndiaye-Lobry D, Lobry C, et al. Tet2 loss leads to increased hematopoietic stem cell self-renewal and myeloid transformation. Cancer Cell. 2011;20:11–24. https://doi.org/10.1016/j.ccr.2011.06.001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Neri F, Incarnato D, Krepelova A, Dettori D, Rapelli S, Maldotti M, et al. TET1 is controlled by pluripotency-associated factors in ESCs and downmodulated by PRC2 in differentiated cells and tissues. Nucleic Acids Res. 2015;43:6814–26. https://doi.org/10.1093/nar/gkv392.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Serandour AA, Avner S, Oger F, Bizot M, Percevault F, Lucchetti-Miganeh C, et al. Dynamic hydroxymethylation of deoxyribonucleic acid marks differentiation-associated enhancers. Nucleic Acids Res. 2012;40:8255–65. https://doi.org/10.1093/nar/gks595.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hahn MA, Qiu R, Wu X, Li AX, Zhang H, Wang J, et al. Dynamics of 5-hydroxymethylcytosine and chromatin marks in Mammalian neurogenesis. Cell Rep. 2013;3:291–300. https://doi.org/10.1016/j.celrep.2013.01.011.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Tsagaratou A, Aijo T, Lio CW, Yue X, Huang Y, Jacobsen SE, et al. Dissecting the dynamic changes of 5-hydroxymethylcytosine in T-cell development and differentiation. Proc Natl Acad Sci USA. 2014;111:E3306–15. https://doi.org/10.1073/pnas.1412327111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Nestor CE, Lentini A, Hagg Nilsson C, Gawel DR, Gustafsson M, Mattson L, et al. 5-Hydroxymethylcytosine remodeling precedes lineage specification during differentiation of human CD4(+) T cells. Cell Rep. 2016;16:559–70. https://doi.org/10.1016/j.celrep.2016.05.091.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Taylor SE, Li YH, Smeriglio P, Rath M, Wong WH, Bhutani N. Stable 5-hydroxymethylcytosine (5hmC) acquisition marks gene activation during chondrogenic differentiation. J Bone Miner Res. 2016;31:524–34. https://doi.org/10.1002/jbmr.2711.

    Article  CAS  PubMed  Google Scholar 

  38. Ichiyama K, Chen T, Wang X, Yan X, Kim BS, Tanaka S, et al. The methylcytosine dioxygenase Tet2 promotes DNA demethylation and activation of cytokine gene expression in T cells. Immunity. 2015;42:613–26. https://doi.org/10.1016/j.immuni.2015.03.005.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Wu H, D’Alessio AC, Ito S, Wang Z, Cui K, Zhao K, et al. Genome-wide analysis of 5-hydroxymethylcytosine distribution reveals its dual function in transcriptional regulation in mouse embryonic stem cells. Genes Dev. 2011;25:679–84. https://doi.org/10.1101/gad.2036011.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Booth MJ, Branco MR, Ficz G, Oxley D, Krueger F, Reik W, et al. Quantitative sequencing of 5-methylcytosine and 5-hydroxymethylcytosine at single-base resolution. Science. 2012;336:934–7. https://doi.org/10.1126/science.1220671.

    Article  CAS  PubMed  Google Scholar 

  41. Pastor WA, Pape UJ, Huang Y, Henderson HR, Lister R, Ko M, et al. Genome-wide mapping of 5-hydroxymethylcytosine in embryonic stem cells. Nature. 2011;473:394–7. https://doi.org/10.1038/nature10102.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Yu M, Hon GC, Szulwach KE, Song CX, Zhang L, Kim A, et al. Base-resolution analysis of 5-hydroxymethylcytosine in the mammalian genome. Cell. 2012;149:1368–80. https://doi.org/10.1016/j.cell.2012.04.027.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Tan L, Xiong LJ, Xu WQ, Wu FZ, Huang N, Xu YF, et al. Genome-wide comparison of DNA hydroxymethylation in mouse embryonic stem cells and neural progenitor cells by a new comparative hMeDIP-seq method. Nucleic Acids Res. 2013. https://doi.org/10.1093/nar/gkt091.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Kim M, Park YK, Kang TW, Lee SH, Rhee YH, Park JL, et al. Dynamic changes in DNA methylation and hydroxymethylation when hES cells undergo differentiation toward a neuronal lineage. Hum Mol Genet. 2014;23:657–67. https://doi.org/10.1093/hmg/ddt453.

    Article  CAS  PubMed  Google Scholar 

  45. Kriaucionis S, Heintz N. The nuclear DNA base 5-hydroxymethylcytosine is present in purkinje neurons and the brain. Science. 2009;324:929–30. https://doi.org/10.1126/science.1169786.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Han D, Lu XY, Shih AH, Nie J, You QC, Xu MM, et al. A highly sensitive and robust method for genome-wide 5hmC profiling of rare cell populations. Mol Cell. 2016;63:711–9. https://doi.org/10.1016/j.molcel.2016.06.028.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Colquitt BM, Allen WE, Barnea G, Lomvardas S. Alteration of genic 5-hydroxymethylcytosine patterning in olfactory neurons correlates with changes in gene expression and cell identity. Proc Natl Acad Sci USA. 2013;110:14682–7. https://doi.org/10.1073/pnas.1302759110.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Dawlaty MM, Breiling A, Le T, Barrasa MI, Raddatz G, Gao Q, et al. Loss of Tet enzymes compromises proper differentiation of embryonic stem cells. Dev Cell. 2014;29:102–11. https://doi.org/10.1016/j.devcel.2014.03.003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Lu F, Liu Y, Jiang L, Yamaguchi S, Zhang Y. Role of Tet proteins in enhancer activity and telomere elongation. Genes Dev. 2014;28:2103–19. https://doi.org/10.1101/gad.248005.114.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Hackett JA, Reddington JP, Nestor CE, Dunican DS, Branco MR, Reichmann J, et al. Promoter DNA methylation couples genome-defence mechanisms to epigenetic reprogramming in the mouse germline. Development. 2012;139:3623–32. https://doi.org/10.1242/dev.081661.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Liu R, Jin Y, Tang WH, Qin L, Zhang X, Tellides G, et al. Ten-eleven translocation-2 (TET2) is a master regulator of smooth muscle cell plasticity. Circulation. 2013;128:2047–57. https://doi.org/10.1161/CIRCULATIONAHA.113.002887.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Zhang R, Cui Q, Murai K, Lim Yen C, Smith Zachary D, Jin S, et al. Tet1 regulates adult hippocampal neurogenesis and cognition. Cell Stem Cell. 2013;13:237–45. https://doi.org/10.1016/j.stem.2013.05.006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Yang R, Yu T, Kou X, Gao X, Chen C, Liu D, et al. Tet1 and Tet2 maintain mesenchymal stem cell homeostasis via demethylation of the P2rX7 promoter. Nat Commun. 2018;9:2143. https://doi.org/10.1038/s41467-018-04464-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Kong LC, Tan L, Lv RT, Shi ZN, Xiong LJ, Wu FZ, et al. A primary role of TET proteins in establishment and maintenance of de novo bivalency at CpG islands. Nucleic Acids Res. 2016;44:8682–92. https://doi.org/10.1093/nar/gkw529.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Lister R, Pelizzola M, Dowen RH, Hawkins RD, Hon G, Tonti-Filippini J, et al. Human DNA methylomes at base resolution show widespread epigenomic differences. Nature. 2009;462:315–22. https://doi.org/10.1038/nature08514.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Xie W, Schultz MD, Lister R, Hou Z, Rajagopal N, Ray P, et al. Epigenomic analysis of multiline age differentiation of human embryonic stem cells. Cell. 2013;153:1134–48. https://doi.org/10.1016/j.cell.2013.04.022.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Wu H, D’Alessio AC, Ito S, Xia K, Wang Z, Cui K, et al. Dual functions of Tet1 in transcriptional regulation in mouse embryonic stem cells. Nature. 2011;473:389–93. https://doi.org/10.1038/nature09934.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Li Y, Zheng H, Wang Q, Zhou C, Wei L, Liu X, et al. Genome-wide analyses reveal a role of Polycomb in promoting hypomethylation of DNA methylation valleys. Genome Biol. 2018;19:18. https://doi.org/10.1186/s13059-018-1390-8.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Neri F, Incarnato D, Krepelova A, Rapelli S, Pagnani A, Zecchina R, et al. Genome-wide analysis identifies a functional association of Tet1 and Polycomb repressive complex 2 in mouse embryonic stem cells. Genome Biol. 2013;14:R91. https://doi.org/10.1186/gb-2013-14-8-r91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Deplus R, Delatte B, Schwinn MK, Defrance M, Mendez J, Murphy N, et al. TET2 and TET3 regulate GlcNAcylation and H3K4 methylation through OGT and SET1/COMPASS. EMBO J. 2013;32:645–55. https://doi.org/10.1038/emboj.2012.357.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Orlanski S, Labi V, Reizel Y, Spiro A, Lichtenstein M, Levin-Klein R, et al. Tissue-specific DNA demethylation is required for proper B-cell differentiation and function. Proc Natl Acad Sci USA. 2016;113:5018–23. https://doi.org/10.1073/pnas.1604365113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Hon GC, Song CX, Du T, Jin F, Selvaraj S, Lee AY, et al. 5mC oxidation by Tet2 modulates enhancer activity and timing of transcriptome reprogramming during differentiation. Mol Cell. 2014;56:286–97. https://doi.org/10.1016/j.molcel.2014.08.026.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Mahe EA, Madigou T, Serandour AA, Bizot M, Avner S, Chalmel F, et al. Cytosine modifications modulate the chromatin architecture of transcriptional enhancers. Genome Res. 2017;27:947–58. https://doi.org/10.1101/gr.211466.116.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Mellen M, Ayata P, Dewell S, Kriaucionis S, Heintz N. MeCP2 binds to 5hmC enriched within active genes and accessible chromatin in the nervous system. Cell. 2012;151:1417–30. https://doi.org/10.1016/j.cell.2012.11.022.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Hu S, Wan J, Su Y, Song Q, Zeng Y, Nguyen HN, et al. DNA methylation presents distinct binding sites for human transcription factors. Elife. 2013;2:e00726. https://doi.org/10.7554/eLife.00726.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Yang YQA, Zhao JC, Fong KW, Kim J, Li SZ, Song CX, et al. FOXA1 potentiates lineage-specific enhancer activation through modulating TET1 expression and function. Nucleic Acids Res. 2016;44:8153–64. https://doi.org/10.1093/nar/gkw498.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Suzuki T, Shimizu Y, Furuhata E, Maeda S, Kishima M, Nishimura H, et al. RUNX1 regulates site specificity of DNA demethylation by recruitment of DNA demethylation machineries in hematopoietic cells. Blood Adv. 2017;1:1699–711. https://doi.org/10.1182/bloodadvances.2017005710.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Birke M, Schreiner S, Garcia-Cuellar MP, Mahr K, Titgemeyer F, Slany RK. The MT domain of the proto-oncoprotein MLL binds to CpG-containing DNA and discriminates against methylation. Nucleic Acids Res. 2002;30:958–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Bach C, Mueller D, Buhl S, Garcia-Cuellar MP, Slany RK. Alterations of the CxxC domain preclude oncogenic activation of mixed-lineage leukemia 2. Oncogene. 2009;28:815–23. https://doi.org/10.1038/onc.2008.443.

    Article  CAS  PubMed  Google Scholar 

  70. Hu DQ, Gao X, Morgan MA, Herz HM, Smith ER, Shilatifard A. The MLL3/MLL4 branches of the COMPASS family function as major histone H3K4 monomethylases at enhancers. Mol Cell Biol. 2013;33:4745–54. https://doi.org/10.1128/Mcb.01181-13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Shukla S, Kavak E, Gregory M, Imashimizu M, Shutinoski B, Kashlev M, et al. CTCF-promoted RNA polymerase II pausing links DNA methylation to splicing. Nature. 2011;479:74–9. https://doi.org/10.1038/nature10442.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Marina RJ, Sturgill D, Bailly MA, Thenoz M, Varma G, Prigge MF, et al. TET-catalyzed oxidation of intragenic 5-methylcytosine regulates CTCF-dependent alternative splicing. EMBO J. 2016;35:335–55. https://doi.org/10.15252/embj.201593235.

    Article  CAS  PubMed  Google Scholar 

  73. Ito S, D’Alessio AC, Taranova OV, Hong K, Sowers LC, Zhang Y. Role of tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass specification. Nature. 2010;466:1129–33. https://doi.org/10.1038/nature09303.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Tsagaratou A, Gonzalez-Avalos E, Rautio S, Scott-Browne JP, Togher S, Pastor WA, et al. TET proteins regulate the lineage specification and TCR-mediated expansion of iNKT cells. Nat Immunol. 2017;18:45–53. https://doi.org/10.1038/ni.3630.

    Article  CAS  PubMed  Google Scholar 

  75. Rudenko A, Dawlaty MM, Seo J, Cheng AW, Meng J, Le T, et al. Tet1 is critical for neuronal activity-regulated gene expression and memory extinction. Neuron. 2013;79:1109–22. https://doi.org/10.1016/j.neuron.2013.08.003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Kumar D, Aggarwal M, Kaas GA, Lewis J, Wang J, Ross DL, et al. Tet1 oxidase regulates neuronal gene transcription, active DNA hydroxy-methylation, object location memory, and threat recognition memory. Neuroepigenetics. 2015;4:12–27. https://doi.org/10.1016/j.nepig.2015.10.002.

    Article  PubMed  PubMed Central  Google Scholar 

  77. Zhu X, Girardo D, Govek EE, John K, Mellen M, Tamayo P, et al. Role of Tet1/3 genes and chromatin remodeling genes in cerebellar circuit formation. Neuron. 2016;89:100–12. https://doi.org/10.1016/j.neuron.2015.11.030.

    Article  CAS  PubMed  Google Scholar 

  78. Karuppagounder SS, Kumar A, Shao DS, Zille M, Bourassa MW, Caulfield JT, et al. Metabolism and epigenetics in the nervous system: creating cellular fitness and resistance to neuronal death in neurological conditions via modulation of oxygen-, iron-, and 2-oxoglutarate-dependent dioxygenases. Brain Res. 2015;1628:273–87. https://doi.org/10.1016/j.brainres.2015.07.030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Zhang J, Chen SQ, Zhang DM, Shi ZX, Li H, Zhao TB, et al. Tet3-mediated DNA demethylation contributes to the direct conversion of fibroblast to functional neuron. Cell Rep. 2016;17:2326–39. https://doi.org/10.1016/j.celrep.2016.10.081.

    Article  CAS  PubMed  Google Scholar 

  80. Li Z, Cai X, Cai CL, Wang J, Zhang W, Petersen BE, et al. Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies. Blood. 2011;118:4509–18. https://doi.org/10.1182/blood-2010-12-325241.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Delhommeau F, Dupont S, Della Valle V, James C, Trannoy S, Masse A, et al. Mutation in TET2 in myeloid cancers. N Engl J Med. 2009;360:2289–301. https://doi.org/10.1056/NEJMoa0810069.

    Article  PubMed  Google Scholar 

  82. Madzo J, Liu H, Rodriguez A, Vasanthakumar A, Sundaravel S, Caces DB, et al. Hydroxymethylation at gene regulatory regions directs stem/early progenitor cell commitment during erythropoiesis. Cell Rep. 2014;6:231–44. https://doi.org/10.1016/j.celrep.2013.11.044.

    Article  CAS  PubMed  Google Scholar 

  83. Yang R, Qu C, Zhou Y, Konkel JE, Shi S, Liu Y, et al. Hydrogen sulfide promotes Tet1- and Tet2-mediated Foxp3 demethylation to drive regulatory T cell differentiation and maintain immune homeostasis. Immunity. 2015;43:251–63. https://doi.org/10.1016/j.immuni.2015.07.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Yan HX, Wang YM, Qu XL, Li J, Hale J, Huang YM, et al. Distinct roles for TET family proteins in regulating human erythropoiesis. Blood. 2017;129:2002–12. https://doi.org/10.1182/blood-2016-08-736587.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Montagner S, Leoni C, Emming S, Della Chiara G, Balestrieri C, Barozzi I, et al. TET2 regulates mast cell differentiation and proliferation through catalytic and non-catalytic activities. Cell Rep. 2016;15:1566–79. https://doi.org/10.1016/j.celrep.2016.04.044.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Yue XJ, Trifari S, Aijo T, Tsagaratou A, Pastor WA, Zepeda-Martinez JA, et al. Control of Foxp3 stability through modulation of TET activity. J Exp Med. 2016;213:377–97. https://doi.org/10.1084/jem.20151438.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Kitagawa Y, Ohkura N, Kidani Y, Vandenbon A, Hirota K, Kawakami R, et al. Guidance of regulatory T cell development by Satb1-dependent super-enhancer establishment. Nat Immunol. 2017;18:173–83. https://doi.org/10.1038/ni.3646.

    Article  CAS  PubMed  Google Scholar 

  88. Zhong X, Wang QQ, Li JW, Zhang YM, An XR, Hou J. Ten-eleven translocation-2 (Tet2) is involved in myogenic differentiation of skeletal myoblast cells in vitro. Sci Rep UK. 2017. https://doi.org/10.1038/srep43539.

    Article  Google Scholar 

  89. Ancey PB, Ecsedi S, Lambert MP, Talukdar FR, Cros MP, Glaise D, et al. TET-catalyzed 5-hydroxymethylation precedes HNF4A promoter choice during differentiation of bipotent liver progenitors. Stem Cell Rep. 2017;9:264–78. https://doi.org/10.1016/j.stemcr.2017.05.023.

    Article  CAS  Google Scholar 

  90. Shimozaki K. Ten-eleven translocation 1 and 2 confer overlapping transcriptional programs for the proliferation of cultured adult neural stem cells. Cell Mol Neurobiol. 2017;37:995–1008. https://doi.org/10.1007/s10571-016-0432-6.

    Article  CAS  PubMed  Google Scholar 

  91. Kim H, Jang WY, Kang MC, Jeong J, Choi M, Sung Y, et al. TET1 contributes to neurogenesis onset time during fetal brain development in mice. Biochem Biophys Res Commun. 2016;471:437–43. https://doi.org/10.1016/j.bbrc.2016.02.060.

    Article  CAS  PubMed  Google Scholar 

  92. Seritrakul P, Gross JM. Tet-mediated DNA hydroxymethylation regulates retinal neurogenesis by modulating cell-extrinsic signaling pathways. PLoS Genet. 2017;13:e1006987. https://doi.org/10.1371/journal.pgen.1006987.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Mi YJ, Gao XC, Dai JX, Ma Y, Xu LX, Jin WL. A novel function of TET2 in CNS: sustaining neuronal survival. Int J Mol Sci. 2015;16:21846–57. https://doi.org/10.3390/ijms160921846.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Fong KSK, Hufnagel RB, Khadka VS, Corley MJ, Maunakea AK, Fogelgren B, et al. A mutation in the tuft mouse disrupts TET1 activity and alters the expression of genes that are crucial for neural tube closure. Dis Model Mech. 2016;9:585–96. https://doi.org/10.1242/dmm.024109.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Zhao X, Dai J, Ma Y, Mi Y, Cui D, Ju G, et al. Dynamics of ten-eleven translocation hydroxylase family proteins and 5-hydroxymethylcytosine in oligodendrocyte differentiation. Glia. 2014;62:914–26. https://doi.org/10.1002/glia.22649.

    Article  PubMed  Google Scholar 

  96. Rao LJ, Yi BC, Li QM, Xu Q. TET1 knockdown inhibits the odontogenic differentiation potential of human dental pulp cells. Int J Oral Sci. 2016;8:110–6. https://doi.org/10.1038/ijos.2016.4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Kunimoto H, Fukuchi Y, Sakurai M, Sadahira K, Ikeda Y, Okamoto S, et al. Tet2 disruption leads to enhanced self-renewal and altered differentiation of fetal liver hematopoietic stem cells. Sci Rep. 2012;2:273. https://doi.org/10.1038/srep00273.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Li C, Lan Y, Schwartz-Orbach L, Korol E, Tahiliani M, Evans T, et al. Overlapping requirements for Tet2 and Tet3 in normal development and hematopoietic stem cell emergence. Cell Rep. 2015;12:1133–43. https://doi.org/10.1016/j.celrep.2015.07.025.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Carty SA, Gohil M, Banks LB, Cotton RM, Johnson ME, Stelekati E, et al. The loss of TET2 promotes CD8(+) T cell memory differentiation. J Immunol. 2018;200:82–91. https://doi.org/10.4049/jimmunol.1700559.

    Article  CAS  PubMed  Google Scholar 

  100. Chapman CG, Mariani CJ, Wu F, Meckel K, Butun F, Chuang A, et al. TET-catalyzed 5-hydroxymethylcytosine regulates gene expression in differentiating colonocytes and colon cancer. Sci Rep. 2015;5:17568. https://doi.org/10.1038/srep17568.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Authors’ contributions

XW and RX wrote the manuscript and prepared the figure and table. GL contributed to critically revising the manuscript at the final stage. RX developed the idea, corrected, and revised the manuscript. All authors read and approved the final manuscript.

Acknowledgements

We thank Dr. Xiaohui Hu and Dr. Garry Wong for constructive comments and critical reading on the manuscript. We apologize to our colleagues whose references were omitted owing to space constraints. Research in the Xie laboratory is supported by grants from the National Natural Science Foundation of China (NSFC 31701276), Macau Science and Technology Development Fund (FDCT 170/2017/A3), and University of Macau Multi-Year Research Grant (MYRG 2016-00065-FHS).

Competing interests

The authors declare that they have no competing interests.

Availability of data and materials

Not applicable.

Consent for publication

Not applicable.

Ethics approval and consent to participate

Not applicable.

Funding

This work was supported by grants from the National Natural Science Foundation of China (NSFC 31701276), Macau Science and Technology Development Fund (FDCT 170/2017/A3), and University of Macau Multi-Year Research Grant (MYRG 2016-00065-FHS) to RX.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ruiyu Xie.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, X., Li, G. & Xie, R. Decoding the role of TET family dioxygenases in lineage specification. Epigenetics & Chromatin 11, 58 (2018). https://doi.org/10.1186/s13072-018-0228-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13072-018-0228-7

Keywords