Skip to main content
Advanced Genetics logoLink to Advanced Genetics
. 2020 Nov 27;2(1):e10033. doi: 10.1002/ggn2.10033

Active DNA demethylation—The epigenetic gatekeeper of development, immunity, and cancer

Rahul Prasad 1, Timothy J Yen 1,, Alfonso Bellacosa 1
PMCID: PMC9744510  PMID: 36618446

Abstract

DNA methylation is a critical process in the regulation of gene expression with dramatic effects in development and continually expanding roles in oncogenesis. 5‐Methylcytosine was once considered to be an inherited and stably repressive epigenetic mark, which can be only removed by passive dilution during multiple rounds of DNA replication. However, in the past two decades, physiologically controlled DNA demethylation and deamination processes have been identified, thereby revealing the function of cytosine methylation as a highly regulated and complex state—not simply a static, inherited signature or binary on‐off switch. Alongside these fundamental discoveries, clinical studies over the past decade have revealed the dramatic consequences of aberrant DNA demethylation. In this review we discuss DNA demethylation and deamination in the context of 5‐methylcytosine as critical processes for physiological and physiopathological transitions within three states—development, immune maturation, and oncogenic transformation; and we describe the expanding role of DNA demethylating drugs as therapeutic agents in cancer.

Keywords: AID, APOBEC, DNA deamination, DNA demethylation, DNA methylation, TDG, TET


Thymine DNA glycosylase (TDG) “clears the cytosine bases” through base excision repair. DNA methyltransferases “start the ball game” by methylating cytosine at the 5‐carbon position to produce 5‐methylcytosine (5mC), which then serves as a substrate for oxidation or deamination. Left panel: In the case of oxidation, ten‐eleven translocation (TET) enzymes sequentially oxidize 5mC, first to 5‐hydroxymethylcytosine (5hmC), then to 5‐formylcytosine and finally 5‐carboxylcytosine. Both of the latter oxidation products are substrates for base excision by TDG. Right panel: In the case of deamination, 5mC is deaminated by AID/APOBEC family enzymes to thymine (T), creating a T:G mismatch—the nominal substrate for TDG. Similarly, 5hmC could possibly be deaminated to 5‐hydroxymethyluracil (5hmU), another TDG substrate. In addition to 5mC, TET enzymes can also oxidize thymine to 5hmU. Intermediates formed by either process, in particular the abasic site (AP) can be enzymatically targeted by downstream base excision repair activities (AP endonuclease, DNA polymerase β, DNA ligase). Of note, methyl‐binding domain‐4 and SMUG1 are the only other enzymes with known T:G and hmU:G base excision function, respectively.

graphic file with name GGN2-2-e10033-g004.jpg

1. INTRODUCTION

DNA methylation—the addition of a methyl group at the five position of cytosine (C) in DNA to form 5‐methylcytosine (5mC)—is a critical process in the regulation of gene expression that affects multiple cellular functions and is ultimately responsible for significant organ‐level effects in mammals. DNA methylation was first discovered in the context of development, with the vast majority found at palindromic CpG dinucleotides. The presence or absence of these marks within gene promoters and enhancers have since been shown to influence progenitor and stem cells across the developmental spectrum. The hematopoietic system is one of the best examples of DNA methylation influence, as changes accompany every step of stem cell functional development, and ultimately affect the commitment, progression, and differentiation of the erythroid, myeloid, and lymphoid lineages. Given its significance in regulating cellular states and functions, defects in CpG methylation are responsible for a plethora of developmental problems.

It is tempting to consider genomic DNA methylation patterns, that is, the DNA methylome, as merely repressive marks, which are only detrimental when ectopically present or absent at particular genes of interest. However, disruptions of the DNA methylome in cancer suggest higher level functions preceding regulation of gene expression. 1 , 2 , 3 One such disruption is the genome‐wide hypomethylation of LINE‐1 in cancer, a phenomenon seen in multiple tissue types. 4 , 5 , 6 , 7 The function of DNA methylation as a regulator of cellular proliferation and differentiation naturally implies a role in oncogenesis, which has indeed been amply corroborated. Evidence continues to grow in identifying both developmentally relevant and cancer‐specific gene expression changes that are driven by aberrations in the DNA methylome.

Only recently has it become clear that the DNA methylome is tightly and dynamically regulated, with both “writers” and “erasers” modulating epigenetic marks in opposite directions via enzymatic activity. DNA methyltransferases (DNMTs) were identified as a class of enzymes, not only conserving or maintaining methylated sites following DNA replication (DNMT1), but also capable of de novo placement of methylation sites (DNMT3A and DNMT3B). The significant medical result of uncovering the process of maintenance and de novo methylation was the discovery of two cytidine analogs, 5‐azacytidine (azacitidine), and its deoxy derivative 5‐aza‐2′‐deoxycytidine (decitabine). These two analogs were originally developed as anticancer antimetabolites, and can irreversibly inhibit DNMTs through covalent binding, ultimately leading to hypomethylation.

While the inhibition of DNMT enzymes induces a gradual, accumulated reduction in methylation, a physiological system of removing cytosine methylation has long been suspected, but remained controversial until approximately 10 years ago. The identification of dioxygenases of the ten‐eleven translocation (TET) family and of the DNA repair enzyme thymine DNA glycosylase (TDG) as the main enzymes responsible for active DNA demethylation in mammals confirmed this suspicion. Over the past decade, much progress has been made in delineating the mechanism of DNA demethylation in both biochemical and clinical contexts. A comprehensive view of the current data suggests that DNA demethylation plays a critical role in physiological and physiopathological transitions, with expanding potential as a therapeutic target.

2. MECHANISMS OF DNA DEMETHYLATION

So far three mechanisms for demethylation have been identified—active demethylation (replication independent), passive demethylation mediated by TET (replication dependent), and 5mC deamination (Figure 1). Although far less established, and probably less prevalent than the two TET‐mediated pathways, evidence continues to grow suggesting the presence and importance of deamination in regulation of demethylation.

FIGURE 1.

FIGURE 1

Pathways of DNA demethylation through oxidation or deamination. Bases are signified by colored shapes (circles = cytosine variants, square = thymine/hydroxymethyluracil), the stem signifying the deoxyribose moiety. A, Replication‐independent DNA demethylation: ten‐eleven translocation (TET) enzymes oxidize 5mC (black) to 5hmC (orange) and then sequentially further oxidize to 5fC (red) and 5caC (burgundy). All products are stable, and the latter two are substrates for thymine DNA glycosylase (TDG). Excision by TDG results in an abasic site that is processed in a replication‐independent manner by AP endonuclease and downstream base excision repair activities to produce an unmethylated cytosine (white). B, Replication‐dependent DNA demethylation: 5hmC is the most prevalent product of TET enzyme activity and is largely found at CpG sites. DNA containing fully methylated CpGs is converted to strands of hemi‐methylated DNA during replication (right side of panel). The DNMT1/UHRF complex maintains DNA methylation during replication by targeting hemi‐methylated DNA and methylating the newly incorporated, unmethylated cytosine (white) in the complementary strand to 5mC (black). 5hmC (orange), however, is not/poorly recognized by the DNMT1/UHRF complex, as indicated (left side of panel). The “hemi‐hydroxymethylated” DNA is either further oxidized by TET enzymes during the remaining cell cycle or “diluted” by subsequent rounds of replication. Thus, in the absence of TET enzyme activity, methylated DNA remains undiluted by replication. C, DNA demethylation by deamination: When methylated cytosine is deaminated (as opposed to oxidized in A and B), thymine (blue) is produced. The presence of thymine opposite guanine, the natural base partner of cytosine, creates a DNA mismatch. TET enzymes are also capable of oxidizing thymine, resulting in 5hmU (fuchsia). Both thymine and 5hmU, when mismatched with G, are specific substrates for TDG; mismatched thymine and 5hmU are also specific substrates of MBD4 and SMUG1, respectively

2.1. 5‐Methylcytosine demethylation—replication independent

In the “replication‐independent” active demethylation, TET dioxygenases sequentially oxidize 5mC to 5‐hydroxymethylcytosine (5hmC), and further to 5‐formylcytosine (5fC), and finally to 5‐carboxylcytosine (5caC). 8 , 9 , 10 5fC and 5caC are then removed by base excision repair (BER), namely by DNA glycosylases (Figure 2). As a result, the loss of methylated cytosine is not due to the lack of binding of the DNMT1‐UHRF in a replication‐coupled process (see below), but direct excision through BER. The main DNA glycosylase capable of excising the TET‐generated 5fC and 5caC is TDG 9 , 10 , 11 ; in addition, Nei‐like 1 DNA glycosylase (NEIL1) can also remove 5caC, but not 5fC. 12 , 13 Of note, the partial redundancy of TDG and NEIL1 appears to be mechanistically relevant, as NEIL1 directly interacts and enhances TDG activity on 5caC. 13 Upon removal of 5fC and 5caC, the resulting apurinic/apyrimidinic/abasic (AP site) is cleaved by AP endonuclease (APE); the one residue gap is filled by DNA polymerase β and finally the nick is sealed by DNA ligase. 14

FIGURE 2.

FIGURE 2

Biochemical details of pathways of DNA demethylation. Schematic of DNA demethylation pathways. 5caC, 5‐carboxylcytosine; 5fC, 5‐formylcytosine; 5hmC, 5‐hydroxymethylcytosine; 5hmU, 5‐hydroxymethyluracil; 5mC, 5‐methylcytosine; AID, activation‐induced deaminase; AP site, apurinic/apyrimidinic site; APE, AP endonuclease; APOBEC3A, apolipoprotein B RNA‐editing catalytic component 3A; C, cytosine; DNMTs, DNA methyltransferases; MBD4, methyl‐binding domain 4; NEIL1, Nei‐like 1 glycosylase; SMUG1, single‐strand selective monofunctional uracil DNA glycosylase 1; T, thymine; TDG, thymine DNA glycosylase; TETs, ten‐eleven translocation dioxygenases; U, uracil; UDG, uracil DNA glycosylase

This pathway is very significant in development, likely due to the need for rapid programming of cellular state changes for pathway commitment, for example, myeloid vs lymphoid, neural vs neural crest. In another word, aggressive demethylation is needed since cells are rapidly proliferating but also malleable—a critical point). The clear evidence of the developmental significance of this pathway is the embryonic lethality phenotype of TDG null mice. 15 , 16

2.2. 5‐Methylcytosine demethylation—replication dependent

Characterization of TET enzyme function and recruitment has shown that the most prevalent process of demethylation is linked with replication and mediated by gene expression and reprogramming via transcriptional states. 17 , 18 , 19 The model is as follows: (a) frequent transcription of genes results in the recruitment of TET enzymes by transcription factors; (b) TET enzymes induce the oxidation of 5mC to 5hmC; (c) 5hmC bases are poorly recognized by the DNMT1‐UHRF complex, which targets and fully methylates the transiently hemimethylated strands during replication; (d) after oxidation, the strand becomes hemi‐methylated following one round of replication, and then fully demethylated after two rounds of replication. Thus, this mechanism of demethylation is likely a reflection of transcriptional activity, rather than induction of transcription per se.

This model is consistent with the fact that “CpG islands”—large intragenic CpG‐rich regions that act as cis‐elements at gene promoters—are largely unmethylated, that is, methylation is inversely proportional to gene expression. Accordingly, the role of TET enzymes in mediating passive demethylation is very significant during differentiation, when a lot of transcriptional activities occur. This is likely due to the need to maintain lineage and pathway commitment once established (ie, once the ball is rolling, keep the pace since the cells are proliferating quickly), as is the case for differentiation of immune progenitors to mature immune cells. 20 One potential problem with this model is that DNMT1 activity is affected by the oxidation state of the hemi‐methylated strand—its activity is only about a third on the Cs opposite to 5hmCs compared to those Cs opposite to 5mC. 21

2.3. Deamination of cytosine and 5‐methylcytosine

Deamination of cytosine is one of the most characterized processes of cytosine modification, but the least understood in terms of pathway involvement. While pyrimidines are known to be prone to spontaneous deamination, 22 , 23 enzymatic deamination of cytosine is driven by the AID/APOBEC family of deaminases, turning cytosine into uracil and resulting in the generation of a G:U mismatch. The resulting uracil in the G:U mismatch can be processed by any of four BER DNA glycosylases: uracil DNA glycosylase (UDG, encoded by UNG), TDG, methyl‐binding domain‐4 (MBD4), and single‐strand selective monofunctional uracil DNA glycosylase (SMUG). In addition to BER alone, cytosine deamination can be repaired in a concerted manner by both BER and mismatch repair (MMR) enzymes. The model proposed by Girelli Zubani et al describes two possible forms of G:U MMR—(a) DNA glycosylase‐mediated base excision and nick formation at the mismatch or (b) MLH1/PMS2‐mediated endonuclease activity away from the mismatch. In both cases, error prone repair is mediated by Polη and ExoI, ultimately leading to A/T mutagenesis. The authors convincingly argue that both processes contribute equally as the rate of A/T mutagenesis is only significantly decreased in the Pms2Ung double‐knockout mice, compared to respective single knockout. 24 This “error‐prone” system is utilized in a controlled fashion to diversify the antibody affinity during immunoglobulin switch and somatic hypermutation (SHM), and the nontrivial role of DNA glycosylase repair may allude to the involvement of DNA methylation. 24

Deamination has an inherently important role in the context of DNA methylation. In fact, deamination of 5mC produces thymine, generating a G:T mismatch in the CpG context. Only two enzymes are known to excise thymine in this unique context, TDG and MBD4. While TDG is known for its causing significant embryonic lethality phenotype when mutated, and its defined role in active demethylation, 15 , 16 Mbd4 mutant mice develop normally and do not show a dramatic increase in cancer susceptibility 25 , 26 , 27 ; however, an increased number of C to T transition mutations at CpG sites are noted, that is, CpG to CpA or CpG to TpG, as expected due to unrepaired G:T mismatches. 25 , 26 In addition, when Mbd4 mutant mice were crossed into mice mutant for the tumor suppressor gene Apc or the MMR gene Mlh1, an increased tumor burden and rate of tumor progression is noted. 25 , 26 , 28 This suggests that MBD4 may serve as a possible link between deamination and demethylation.

3. LINKS BETWEEN DEMETHYLATION AND DEAMINATION

Recently, a demethylation pathway involving TET1, AID, and MBD4 was uncovered in mouse embryonic cells; the loss of function of any of these genes resulted in hypermethylation of the transfected CpG island reporter, but not endogenous CpG islands. 29 The model suggested by the authors is that in contrast to maintenance of the overall bimodal methylation pattern, a specific pathway can be activated for “resetting” methylation patterns during somatic cell reprogramming. It is interesting that this putative pathway might be at the intersection between the TET‐mediated oxidation and AID/APOBEC‐mediated deamination of methylated CpGs.

Indeed, previous studies have correlated AID with DNA demethylation. For example, our group has shown in vivo the direct interactions between AID and TDG, 16 but has not functionally tested the involvement of AID or other deaminases in demethylation. MBD4 was shown to be involved in active DNA demethylation in zebrafish, acting in concert with AID and the adaptor protein GADD45, 30 but some of these data is considered controversial. 31 However, since many of these studies were conducted prior to the discovery of TET enzyme function and their establishment as prominent mediators of active demethylation, the functional significance of these interactions needs to be reexamined.

Most work investigating AID/APOBEC enzymes in the context of CpG methylation has focused on characterizing the substrate specificities for 5mC and TET‐oxidized derivatives. The results of in vitro biochemical studies demonstrated that most AID/APOBEC family enzymes have physiologically insignificant affinities for 5mC, largely dismissing the concept of demethylation through deamination. 32 However, studies over the past 5 years reintroduced the concept of deamination as affecting cytosine methylation, through either a direct or indirect biochemical process. APOBEC3A was recently shown as being capable of significantly deaminating 5mC, providing a direct molecular mechanism of generating T:G mismatches from 5mC. 33 Sabag et al, as mentioned previously, showed that the maintenance of the demethylated state in CpG islands in embryonic stem cells is disrupted in TET1 knockout cells, as expected, as well as in AID, MBD4, and GADD45a null lines. 29 As further studies showed the interactions of AID and GADD45a with TDG, 16 a link between demethylation and deamination has been reinforced.

A recent study further advanced the role of GADD45a in demethylation, and possibly deamination, by demonstrating a mechanistic link with TET1. The prevailing paradigm for TET enzyme targeting for demethylation is through its recruitment by transcription factors to transcriptionally active regions. However, details beyond protein‐protein interactions have been largely unknown. Recent work by Niehrs and collegues makes a significant progress in identifying a process of TET1 recruitment to CpG islands in promoters through DNA‐RNA hybrids, or R‐loops. 34 Through antisense transcription of the region, a long noncoding RNA (lncRNA) is produced resulting in R‐loop formation that is then bound by GADD45a, which subsequently recruits TET1 and induces demethylation. Whether this process is complementary or parallel to the role of transcription factor‐induced TET recruitment is an important question that may point to a role for deamination.

Arguably the most significant limitation in the initially proposed deamination‐demethylation pathways was the assumption that oxidized uracil compounds, 5‐hydroxymethyluracil (5hmU), 5‐formyluracil, and 5‐carboxymethyluracil, which are considered to be putative substrates for TDG, MBD4 and SMUG, were derived from deamination of the TET‐oxidized 5mC derivatives. However, a functional study investigating 5hmU showed that the very TET enzymes could oxidize thymine to produce 5hmU physiologically. 35

In vivo studies support the presence of a deamination‐mediated demethylation pathway, possibly intersecting with TET enzyme function. Guo et al demonstrated that overexpression of TET1 in HEK293 cells increased 5hmC, as expected. Intriguingly, the increase in 5hmC was reduced when the deaminase AID was co‐overexpressed with TET1. 36 To investigate whether deaminases were involved in 5hmC demethylation, the authors transfected HEK293 cells with PCR products containing 5hmC. Using bisulfite‐sequencing, a genomic DNA methylation analysis method, they found that, when co‐transfected with AID or APOBEC family expression constructs, the PCR products showed increased rates of demethylation. The sequence context of 5hmC demethylation showed a 2.4‐fold higher rate of demethylated residues in the WRC vs SYC context (where W is A or T, R is A or G, S is G or C and Y is C or T—IUPAC nucleotide codes), reflective of the AID sequence selectivity for WRC motifs. Finally, overall distribution of the 5hmC demethylation deviated significantly from the expected Poisson distribution, suggestive of a processive (ie, by scanning), and not distributive (ie, by random targeting), mechanism of demethylation. 36

The concept of deamination and processive demethylation has since been investigated, but remains significantly understudied overall. Franchini et al used a DNA targeting domain (GAL4) fused to AID and a Xenopus extract system to show that deamination in the vicinity of 5mC residues induced demethylation. 37 The group then applied the same targeting approach in vivo, breeding transgenic GAL4‐AID mice to transgenic mice bearing the GAL4 binding sites (UAS) into the first of four paternal (methylated) H19 differentially methylated regions (DMRs). The results showed demethylation throughout the region, irrespective of GAL4 binding motif context, that is, not just of the first DMR, making AID induced deamination of each CpG very unlikely and instead suggesting involvement of a processive DNA repair pathway (ie, long patch BER or MMR). 37

While the functional relations between deamination and demethylation are intriguing, how the respective reactions might co‐exist is an open question. Why do enzymes such as SMUG1 and TDG act on 5hmU and related residues, if it is unlikely that AID/APOBEC deaminases generate them in the first place? If TDG, an evolutionarily conserved enzyme with an embryonic lethal knock‐out phenotype, has a primary role in demethylation, then why does it have overlapping roles in excising both 5hmC oxidation products (5fC, 5caC) and potential deamination products (T, 5hmU)? One simple yet uninvestigated possibility is that the oxidation reactions by TET enzymes may exacerbate the effects of aberrant deamination. If aberrant deamination of 5mC occurs (ie, by APOBEC3A, thereby producing thymine) in regions of high TET activity, then 5hmU becomes an additional source of DNA damage. In short, AID/APOBEC enzymes may increase the need for a “guardian” of TET (over‐)activity; with this perspective, TDG might have a well‐designed role in maintaining the genomic code (Figure 3).

FIGURE 3.

FIGURE 3

Blurred boundaries between physiological DNA modifications and pathological DNA damage: the underlying central role of thymine DNA glycosylase (TDG) and base excision repair in active demethylation. The production of stable 5hmC oxidation derivatives (5fC and 5caC) by ten‐eleven translocation dioxygenase (TET) enzymes has been well established in vivo but their function remains unresolved. Conversely, 5hmU has long thought to be a form of oxidative DNA damage, yet the regulation of deaminating enzymes is understudied, despite evidence of significant effects on the global DNA methylation state. Enzymatic DNA oxidation and deamination are conceptually linked through the production of substrates specifically excised by TDG. 5hmU, 5‐hydroxymethyluracil; 5mC, 5‐methylcytosine; 5hmC,5‐hydroxymethylcytosine; 5fC, 5‐formylcytosine; 5caC, 5‐carboxylcytosine; T, thymine. DNA bases are color coded as in Figure 1

The most striking example of the relationship between (de‐)methylation and deamination is in their effects on genomic CpG mutagenesis. While the concept of demethylation driven solely by deamination has clearly not been established, the likely intersection between TET and AID/APOBEC enzyme function can clearly be seen in their mutagenic patterns (Table 1). Further studies will reveal whether TET‐mediated oxidation and AID/APOBEC‐mediated deamination function together as a synchronized pathway or independently to induce controlled CpG mutagenesis.

TABLE 1.

TET and AID/APOBEC mutational effects at CpG sites

Context SNV, motif Trend and genomic bias Experimental system References
5hmC C > T Decreased at CpG sites Human 38
5hmC C > G Increased at 5hmC sites Mouse, human 39
TET1 KO C > T Increased at CpG sites in coding genome Mouse, cancer 40
TET2 KO C > T, WRC motif Increased in coding genome Mouse, cancer 41
TET2 KO C > T Increased at CpG sites with 5hmC within coding genome Mouse, cancer 42
TET2/3 DKO C > T increased at CpG sites, euchromatin Mouse, cancer 43
TET2/3 DKO C > T Increased bias at VH186.2 Mouse, cancer 44
APOBEC3A activity C > T, YTCR motif Within DNA hairpins; putative cancer driver hotspots In vitro, human, cancer 45

High

AICDA expression

C > T, WRCG motif

Increased at highly methylated CpG sites

Negative correlation

mutation vs methylation

Human, cancer—lymphoma 46

High

APOBEC expression

C > T

Increased at CpG sites, unmethylated

Negative correlation methylation vs mutation

Human, cancer—TCGA 47

High

APOBEC3B expression

C > T Increased at CpG sites Human, cancer—TCGA 48
High APOBEC3 expression C > T, TCW motif Increased at non‐Ig loci no correlation to CpG sites Human, cancer—CLL 49

A significant element of complexity in uncovering the overlap between methylation and deamination in the genomic context is the reliance on the bisulfite sequencing method, currently the gold standard for identifying methylated cytosines. 50 , 51 Bisulfite sequence analysis identifies methylated cytosines by deaminating all unmethylated but not methylated cytosines. However, as a result of this in vitro deamination, any cell‐produced uracil is indistinguishable from bisulfite‐generated uracil. One approach to address this problem is to compare untreated to bisulfite‐treated sequences with single‐base resolution in “normal” vs AID/APOBEC contexts, allowing a complete assessment of the cytosine modification state.

4. THE SIGNIFICANCE OF 5mC OXIDATION DERIVATIVES

The biochemical pathway of sequential oxidation of 5mC clearly points to TDG as the terminal enzyme, due to its specific excision of 5fC and 5caC. 9 , 10 , 11 The presence of this system is reflected in vivooverexpression of TET enzymes increases 5fC and 5caC, while loss of TET activity leads to the reduction of 5hmCas well as 5fC and 5caC. 8 , 9 , 10 , 52 Accordingly, loss of TDG leads to an increase in 5fC and 5caC, in support of its function as the primary excision enzyme of these substrates.

Shortly after being identified as the terminal enzyme in 5mC oxidation, the possibility of TDG affecting DNA methylation was investigated by us as well as by Schär's group. Using knockout strategies, both groups identified TDG as being essential for embryonic development in mice, and showed that DNA methylation increases at developmentally relevant promoter and enhancer sites. 15 , 16 In addition, our group also described the demethylation defect of homozygous Tdg knock‐in mice bearing a point mutation (Tdg N151A ) that inactivated its glycosylase activity, thus providing compelling genetic evidence on the existence of an active, enzymatically driven DNA demethylation process. 16 , 53 Furthermore, our recent work shows evidence of increased de novo methylation following TDG knockdown in melanoma cell lines 54 and mouse adenomas bearing the Tdg N151A knock‐in mutation, 55 respectively. In summary, the studies using three genetic systems across two mammalian species strongly support the role of TDG in preventing aberrant DNA methylation in the form of elevated 5mC.

The mechanism of how TDG affects DNA methylation physiologically is an important but unanswered question. The simplest explanation is that the accumulation of 5fC and 5caC results in negative feedback producing 5mC. In this scenario, an increase in global 5fC and 5caC would precede a global increase in 5mC. While the first part of this possibility has been consistently demonstrated, the kinetics of 5mC accumulation has not, and thus warrants more in‐depth investigation. Additional possibilities are more complex and imply a direct effect of 5fC and 5caC. One study demonstrated that 5caC accumulation in TDG null mouse embryo fibroblasts induces ectopic CTCF binding sites across the genome. 56 CTCF is a multifunctional transcription factor that binds to DNA in a sequence specific manner influenced by CpG methylation, and has insulator activity, blocking enhancer‐promoter communications. The best characterized example of its function is in imprinting regulation of IGF‐2 expression: when the CTCF binding site is demethylated, CTCF binds and induces a methylation change downstream, thereby repressing IGF‐2 expression. Thus, the presence of ectopic CTCF sites can be predicted as having dramatic changes in methylation patterns that need not be reflected as a global increase in 5mC.

5fC and 5caC may also serve as functional epigenetic marks with effects beyond 5mC and 5hmC levels. Perhaps the most compelling evidence points to a function in chromatin. 5fC can form DNA protein crosslinks with nucleosomes. 57 , 58 Similar to CTCF, ectopic nucleosomes could induce reciprocal change in DNA methylation patterns without a global increase. Thus, the possibility that 5fC and 5caC induce ectopic repressive marks at a chromatin level, which then result in focal, but not global, DNA methylation increases, is an intriguing and important unanswered question. Raiber et al showed that 5fC enhanced nucleosome occupancy through stabilization via lysine‐Schiff base crosslinking. 59 This increased occupancy occurs at active enhancers, correlating with increased expression of associated genes, in a tissue specific manner. 59 Another recent study showed reduced TDG enzyme activity at nucleosome bound 5fC that contained histones H2A.Z and H3.3, two variants enriched at enhancer sequences. 60 Interestingly, incubation with transcription factor FOXA1 caused increased TDG activity at 5fC sites, consistent with a reported ability of FOXA1 to open compacted chromatin. 60 , 61 Examples of transcription factors preferentially binding to 5caC containing recognition sites have been reported, possibly functioning to promote open chromatin in a manner similar to FOXA1. 62 , 63 , 64 , 65

Perhaps the strongest evidence for 5fC and 5caC oxidation derivatives functioning as markers of active chromatin is in the context of p300, an essential histone acetyltransferase. Song et al used two methods of 5fC genome‐wide profiling to show the enrichment at enhancers as well as a correlation with acquired sites containing p300. 66 As the authors note, this may reflect active 5mC oxidation to facilitate p300 binding. In line with this concept, studies have shown TDG to enhance CBP/p300 mediated transcription through allosteric interactions. 67 , 68 , 69 Similarly, absence of TDG in vivo results in a reduction of histone acetylation, with in vitro studies confirming TDG as a co‐activator of p300 activity. 70

However, the possibility that 5fC and 5caC are inevitable oxidation byproducts of TET enzyme activity cannot be ruled out. Evidence for this lies in characteristics of these bases resembling oxidative damage. 71 , 72 , 73 , 74 , 75 , 76 , 77 , 78 , 79 A mutagenic effect of 5fC was demonstrated prior to the discovery of the TET enzyme mechanism of formation, whereas 5caC was also shown to induce MMR protein activity. 72 , 80 Furthermore, both 5fC and 5caC induce a reduction in transcriptional efficiency—a counterintuitive effect considering the link between TET enzyme mediated demethylation and transcriptional activation. 57 , 81 Thus, transcriptional regulation by DNA (de)methylation/deamination should be considered, in all respects, a repurposing of DNA damage during evolution, by which chemical “lesions” have been redeployed as physiological elements of the epigenetic syntax, which again required the involvement of DNA repair enzymes like TDG to “check and balance” dioxygenase/deaminase activity in order to maintain genomic and epigenomic stability (Figure 3).

While the question of 5fC and 5caC as functional epigenetic marks vs inevitable byproducts remains equivocal, demethylation overall plays a vital role in mammalian development that continues to be delineated. In the context of organogenesis, much progress has been made in neural development, the system most enriched for 5mC oxidation derivatives. Transient changes in 5fC/5caC have been observed in neural lineage specification and in hepatic organogenesis 82 , 83 , 84 (Figure 4). In the hematopoietic system, TET KO mouse studies continue to demonstrate the complex but important roles in the immune system development and maturation, as discussed in a recent review 132 and below in this article.

FIGURE 4.

FIGURE 4

(De)methylation underscores the transition points between oncogenesis and differentiation. The developmental progression of hematopoietic, neural, and hepatic cell lineages is summarized and depicted in “rolling marble” format. Colors of lineages annotated in Table 2 are conserved and bolded. Associated cancers from Table 2 are in red. Solid blue arrows represent transitions reported to involve changes in 5fC/5caC levels. 82 , 83 , 84 , 131 Solid purple arrows represent transitions reported to involve ten‐eleven translocation dioxygenase (TET)/thymine DNA glycosylase (TDG) enzymes, discussed in text. Dashed blue arrows represent transitions predicted to involve changes in 5fC/5caC levels

5. THE CLINICAL SIGNIFICANCE OF DNA (DE)METHYLATION REFLECTS INVOLVEMENT IN DEVELOPMENT AND CANCER

The spectrum of clinically significant mutations bridges the developmental functions of (de)methylating enzymes with their roles in oncogenesis. To illustrate this point, the incidences of the most commonly mutated DNA (de)methylating enzymes (DNMT3A, TET2, and isocitrate dehydrogenase [IDH] 1/2) are listed with the associated cancers, and graphically represented in the context of developmental lineages (Table 2 and Figure 4). IDH 1 and 2 mutations are included as a presumed physiological equivalent to reduced activities of all three TET enzymes. This assumption is based on the roles of IDH1 and IDH2 in the production of alpha‐ketoglutarate, an essential co‐factor for TET‐mediated 5mC oxidation, and the observation that the most frequent IDH1/2 mutations result in the production of 2‐hydroxyglutarate, a global TET enzyme inhibitor. Accordingly, these mutations result in a hypermethylation phenotype and loss of 5hmC, and are mutually exclusive with TET2 mutations in AML. 133

TABLE 2.

Pooled prevalence of DNMT3a, TET2, and IDH1/2 mutations in specific cancers

DNMT3a TET2 IDH1/2
Lineage Cancer type Mutated (or cancer present) Total Prevalence a (%) Mutated (or cancer present) Total Prevalence a (%) Mutated (or cancer present) Total Prevalence a (%) References
Myeloid Primary AML 197 840 23.5 172 945 18.2 42 208 20.2 85, 86, 87, 88, 89, 90, 91
CMML 22 367 6.0 247 592 41.7 4 165 2.4 85, 92, 93, 94, 95, 96, 97, 98, 99, 100
Lymphoid Mature, nodal T‐cell lymphomas 113 407 27.7 281 551 50.9 45 226 19.9 101, 102, 103, 104, 105, 106, 107, 108
Diffuse large B‐cell lymphoma 9 1711 0.5 119 2338 5.1 0 1049 0 101, 109, 110, 111, 112, 113, 114, 115, 116, 117, 118
Neural Low grade glioma 4 512 0.8 2 512 0.4 418 512 81.6 119
GBM 4 397 1.0 0 397 0 25 397 6.3 120
Endoderm HCC 4 366 1.1 1 366 0.3 6 366 1.6 121
Cholangiocarcinoma 0 36 0 0 36 0 16 98 16.3 122, 123
Germline Primary AML 3 16 18.8 1 14 7.1 2 56 3.6 124, 125, 126
CMML 1 14 7.1 127, 128
Lymphoma—T‐cell origin b 2 14 14.3 129, 130
Lymphoma—B‐cell origin c 5 14 35.7
Low grade glioma 6 56 10.7
GBM 0 56 0
HCC 0 56 0
Cholangiocarcinoma 2 56 3.6

Abbreviations: AML, acute myeloid leukemia; CMML, chronic myelomonocytic leukemia; DLBCL, diffuse large B‐cell lymphoma; GBM, glioblastoma multiforme; HCC, hepatocellular carcinoma; LGG, low grade glioma; NLPHL, nodular lymphocyte predominant Hodgkin lymphoma.

a

The number of individuals with somatic mutation were pooled from indicated references. Multiple mutations in the same individual were counted as a single case. Mutations in both genes were not included in this analysis (with the exception of IDH2 and TET2 in AITCL). 108 For germline mutations, prevalence reflects number of reported cancer cases among individuals with germline mutation.

b

Lymphoma of T‐cell origin (germline)—T‐cell origin: T‐cell lymphoma; nodal peripheral T‐cell lymphoma of T follicular helper.

c

Lymphoma of B‐cell origin (germline)—B‐cell origin: NLPHL; EBV‐positive Hodgkin‐like polymorphic B‐cell lymphoproliferative disorder; primary mediastinal large B‐cell lymphoma.

When analyzing the cumulative incidences (Table 2), few notable observations can be made. The incidence of mutations is by far most significant in low‐grade glioma, consistent with the brain having the highest enrichment of 5mC oxidation derivatives. TET2 mutations are more frequent in “mature” type hematopoietic cancers, consistent with its defined function in terminal differentiation/maturation within both myeloid and lymphoid lineages. 132 The incidence patterns of specific cancers in patients carrying germline mutations appear to resemble the somatic mutation patterns, likely in further support of respective developmental functions. One notable exception is the high incidence of TET2 somatic mutations in T‐cell lymphomas vs B‐cell lymphomas in germline mutation carriers, which may reflect differences in survival observed in Tet2 KO mouse models. 134 Finally, the abundance of IDH1/2 mutations in gliomas and cholangiocarcinomas, but lack of TET2 mutations, are likely a reflection of TET enzymes functioning in tissue specific patterns—TET1, for example, has very recently been shown to be an oncogenic driver in IDH‐wt cholangiocarcinoma. 135

6. CYTOSINE MODIFICATION AND IMMUNE FUNCTION

6.1. Methylation

The concept of dynamic epigenetic marks has been investigated in the hematopoietic system, with mapping of DMRs showing specific signatures in the differentiation of progenitor cells. 136 In addition to direct influence on gene expression (ie, methylation of promoter regions), methylation changes in specific transcription factor binding sites seem to maintain predefined “differentiation trajectories.” A clear example is in the commitment of progenitor cells to lymphoid vs myeloid differentiation, in which lymphoid precursors acquire global increase in methylation at DMRs throughout the genome. 136 Notably, this paradigm extends into lymphoid cell maturation: a whole genome analysis of B‐cell differentiation showed a progressive accumulation of methylation changes in each progenitor stage, affecting 30% of all autosomal CpG sites. 137

While the effects on gene expression are variable, the role of methylation is reinforced in physiological models: in mice with DNMT1 knockdown and subsequent CpG hypomethylation, myeloid cell development, and numbers are normal, while lymphoid development is deficient. Interestingly, in the same study, Broske et al noted that cell numbers of all hematopoietic stem cell (HSC) derivatives were normal in mice lacking de novo DNMTs DNMT3A and DNMT3B. This suggests that additional DNA methylation is dispensable in defining the fate of stem cells and further highlighting the nuanced but significant differences within methylation‐dependent effects. 138 In parallel to the global methylation changes noted with B‐cell differentiation, however, a direct functional role for methylation in immune function is apparent at the IgK light chain locus. Here the methylation status of these sites appears to be regulated by DNMT3A and DNMT3B, as conditional knockout in mice produces precocious Vk‐Jk recombination in B‐cell progenitor cells. 139

Finally, the significant role of DNMT enzymes in HSC development is clinically apparent in Immunodeficiency, centromeric instability and facial anomalies syndrome, a rare but notable developmental disease caused by loss‐of‐function mutations in DNMT3B. The clinical hallmark of this syndrome is agammaglobulinemia causing immunodeficiency, which most often leads to early age fatalities. Although patients exhibit levels of lymphoid cells in peripheral blood ranging from normal to low, the significant impact of this disease on immune function highlights the nuanced but potent relationship between methylation and hematopoietic system. 140

6.2. DNA (de)methylation and immune function

The significance of both methylation and demethylation in differentiation is not limited to embryonic development or lineage commitment—it is an essential and continual process in the immune system. As in HSC progression, the combination of noncoding factors (eg, lncRNAs) and transcription factors is linked with DNA methylation status, and as such, the differentiation and maturation of both erythroid and lymphoid cells are a model for the roles of all three pathways of DNA demethylation.

As mentioned previously, TET enzymes are primarily recruited to enhancer regions by transcription factors. This is consistent with the current paradigm that the majority of DNA demethylation is induced by transcriptional activation as a positive feedback to enhance chromatin accessibility. 141 , 142 , 143 , 144 , 145 , 146 , 147 Demethylation coupled to transcription factor binding sites is also seen in plasma cell differentiation, as well as in every type of thymocyte: differentiation of naïve CD4 T cells, memory cell differentiation in CD8 T cells, lineage modulation of invariant natural killer T cells, and Foxp3‐expressing T regulatory (T reg) cells. 148 , 149 , 150 Another example of TET functions is in IgK rearrangement and B‐cell differentiation, wherein, as mentioned above, the loss of de novo DNMTs leads to precocious rearrangement. 139 Accordingly, the TET proteins have been shown to regulate IgK rearrangement by oxidizing 5mC at enhancer regions. 144 For more detail, the reader is referred to a recent review describing the role of TET enzymes and passive demethylation in immune maturation and function. 132

The strongest example of active demethylation involvement in the hematopoietic system is in a recent study by Suzuki et al investigating the master transcription factor RUNX1. Overexpression studies showed the binding site‐directed DNA demethylation, and co‐immunoprecipitation demonstrated the interactions between TET enzymes and TDG. Furthermore, analysis in noncycling cells also showed the RUNX1‐mediated DNA demethylation, serving as convincing support of active demethylation as opposed to a passive process depending on replication. 151

6.3. Deamination

Cytosine mutagenesis mediated by deamination plays an important role in SHM of immunoglobulin genes. Uracil residues introduced from cytosine deamination by AID or APOBEC are processed by uracil glycosylases and repaired by MMR enzymes. Counterintuitively, the critical nicking complex of PMS2/MLH1 has been shown to be dispensable for the generation of mutations at A:T base pairs. However, a recent study showed that the PMS2/MLH1 pathway of mutagenesis acts in concert with uracil glycosylases in strand‐specific manners that expand the overall mutation spectrum. 24

In addition to SHM, DNA deamination plays a role in class switch recombination (CSR) as well. Similar to SHM, CSR relies on the generation of a DNA nick induced by uracil glycosylase/BER recognition and removal of uracil. In a study using B cells derived from hyper IgM patients with UDG mutations, the strand specificity of uracil excision from ssDNA was demonstrated by the lack of compensation by SMUG1. Thus, despite a mechanistic overlap between DNA glycosylases, their in vivo functions appear to be more specific and regulated, as evidenced by strand specificity. 152 Similarly, MBD4 interacts with MMR to facilitate CSR. 153 , 154

Notably, a less publicized but physiologically critical role for AID and deamination is in the suppression of autoreactivity. As AID was originally identified as the driver of antibody diversification through the previously mentioned processes (ie, SHM, CSR), expression and function of AID was thought to be limited to activated and mature B cells. Interestingly, low level of AID expression in immature B‐cells in human bone marrow is essential for suppressing autoantibody development through induction of apoptosis. 155 , 156 The consequence of increased autoreactive B‐cell clones following AID knockdown may very well allude to a phenotype‐modifying role in lymphoma; indeed, autoimmune paraneoplastic syndromes are a common yet highly variable occurrence in lymphomas. 157

Although the mechanisms of demethylation through deamination remain unresolved, recent studies continue to demonstrate their functional connections. AID expression has recently been shown to be directly regulated by the TET enzymes. 158 Recent work by Levine and colleagues argues that AID is an epigenetic regulator of myeloid/erythroid differentiation, through DNA methylation changes in HSC. Kunimoto et al showed that both TET2 and AID impact HSC differentiation through altered transcription via methylation changes at key promoters. Loss of function in both cases causes myeloid expansion and increased erythroid output. The roles of AID and TET2 do not seem to be redundant, however, as alterations in HSC self‐renewal and susceptibility to malignant transformation are seen with the loss of TET2, but not AID. 159

7. MECHANISTIC ROLE OF DNA (DE)METHYLATION IN CANCER

7.1. Ten‐eleven translocations

The connection between demethylation and cancer has been strong from the beginning, with the TET enzymes being the first of many links. TET1, the first identified TET dioxygenase enzyme, was discovered in the early 2000s as the fusion partner to MLL in two cases of AML containing t(10, 11) translocations, one being in an 8‐year old patient. 160 , 161 Discovery of additional family members, TET2 and TET3, has solidified this link, with TET2 being the most clinically defined.

Somatic TET2 mutations were first reported as present in ~15% of myeloid cancers, with the loss of function mutations in TET2 accounting for approximately 10% of newly onset AML, 50% of CMML, and approximately one‐third of all mutations in myelodysplastic syndromes. 92 , 93 Clinically, TET2 mutations appear to have an unfavorable impact on prognosis of AML. In a most recent meta‐analysis by Wang et al, TET2 mutations predicted inferior overall survival as well as event‐free survival for patients with AML, and specifically in patients under 65 years of age (the authors noted that a data deficiency prevented an analysis of patients over 65 years of age) and those cytogenetically normal and intermediate‐risk AML. 162 In the context of MDS, TET2 mutations do not appear to carry a significant prognostic value. 163 Two independent meta‐analysis by Guo et al and Lin et al showed an incidence of 18.34% and 19.19%, respectively. In both studies, no significant difference in overall survival of the pooled cohorts was observed when comparing the patients with TET2 mutations to those without them. 94 , 164

The role of TET2 as a tumor suppressor is not limited to myeloid derived malignancies. TET2 mutations are present in an estimated 56% of patients with angioimmunoblastic T‐cell lymphoma and 46% of peripheral T‐cell lymphomas, suggesting that the TET2 loss has a more dramatic malignant phenotype in the lymphoid derivatives compared to the myeloid line. 101 , 102 , 103 , 104 , 105 , 165 , 166 Notably, TET2 mutations are far less prevalent in diffuse B‐cell lymphomas at ~7% of reported cases. 109 , 167

The clonal expansion of leukemic‐driver mutations in clinically healthy patients has been defined as “clonal hematopoiesis of indeterminate potential” (CHIP), which was first proposed by Steensma et al. 168 In line with their developmental and oncogenic significance, TET2 and DNMT3a mutations represent 93% of driver mutations in clonal hematopoiesis. 169 Underscoring the systemic influence of the hematopoietic system, the significance of CHIP has been demonstrated not as an oncogenic precursor, but instead as an accelerator of atherosclerosis and chronic heart failure. 170 , 171 , 172

The mechanism by which the TET2 loss‐of‐function mutations induce clonal expansion remains in question and likely holds the answer to an apparent paradox: despite the clear association between TET2 mutations and hyperproliferative states, TET2 mutations are not always identified as the driving or founding mutation in AML. The order of mutation acquisition has been investigated in one study in myeloproliferative neoplasms, where a severe clinical phenotype of polycythemia vera was observed when JAK2 mutations preceded TET2 mutations. Notably, the “TET2‐first” mutation acquisition appeared to dampen the upregulating gene expression effects of the pathogenic JAK2 mutation in a cell‐type specific manner. 173 Further details regarding clonal hematopoiesis are discussed in cogent reviews by Bowman et al and Steensma. 168 , 174

While TET2 appears to have a clear role as a tumor suppressor, the function of TET1 in oncogenesis is equivocal as a tumor suppressor vs a driving oncogene. As mentioned above, the significance of TET1 was first noted as a fusion partner with the MLL gene in AML with the t(10, 11) translocation. 160 , 161 A subsequent study showed that TET1 was essential to the oncogenesis resulting from this fusion, primarily through regulation of the TET1‐MLL fusion protein with its significant target genes. 175 The upregulation of TET1 coinciding with a mechanistic role is strongly supported by the global increase in 5hmC levels within these cells. 176 In the case of solid tumors, a recent bioinformatic study reported elevated expression of TET1 in a subset of triple negative breast cancer cases in the Cancer Genome Atlas (TCGA; in comparison to normal breast tissue and hormone‐dependent tumors), which coincided with reduced methylation levels; in addition, TNBC cases with elevated TET1 expression levels and reduced methylation had worse prognosis. 177 As mentioned previously, TET1 has also been shown to be an oncogenic driver in IDH‐wild‐type cholangiocarcinoma. 135

In lymphoid lineage malignancy, however, TET1 appears to play a tumor suppressor role comparable to TET2. TET1 deletion in mouse models, both in isolation and in combination with TET2 deletion, produces B‐cell lymphomas. 40 , 178 Notably, TET1 as a tumor suppressor of B‐cell malignancy has also been supported by evidence of reduced TET1 expression in human non‐Hodgkin's lymphomas. 40

7.2. AID and APOBEC3A

As discussed earlier, the mechanistic role of AID/APOBEC family enzymes in demethylation is still unclear. However, growing evidence continues to show how these deaminating enzymes play a direct role in immune development and oncogenesis. As with the DNMT and TET enzymes, this effect appears to be in both myeloid and lymphoid lineage cancers.

In diffuse large B‐cell lymphoma (DLBCL), a recent study has shown that expression of the activation‐induced cytidine deaminase (AICDA) gene that encodes AID is a driver of global methylation heterogeneity and hypomethylation, the features that are correlated with inferior clinical outcome. 179 Another study investigated the methylation levels of the promoter region of the Fanconi anemia complementation group A (FANCA) gene, which is highly expressed in DLBCL. By correlating AID with recruitment of TET2, subsequent demethylation of the promoter, and increased FANCA gene expression, the complex but significant synergy between AID and TET2 is highlighted, mirroring the similar process discussed earlier that is seen in HSC. 159 , 180 Coupled with the findings of Lio et al showing TET enzymes as directly modulating the expression levels of AID through the “AICDA superenhancer,” the framework for an inherent feedback loop appears in place. 158

As the prognostic and therapeutic value of tumor mutational burden becomes much clearer in both solid and liquid tumor types, the mutagenic role of the deaminase class enzymes beyond AID remains an important question. 181 , 182 For example, APOBEC activity defines 3 of 9 mutational signatures identified from over 1000 tumor genomes. 183

APOBEC3A, significant for its ability to deaminate 5mC, appears to have an important role in oncogenesis. An enrichment of mutational motif specific to APOBEC3A compared to other APOBEC family members has been defined as an “A3A‐like” signature; notably, this signature can be distinguished from other APOBEC enzyme signatures in multiple cancer types with high levels of APOBEC‐driven mutations. 184 A recent study with potentially significant therapeutic relevance found that the increased APOBEC3A contributes to mutational burden and drives the development of “passenger hotspot mutations” or mutational red herrings that have been mischaracterized as driver mutations. 45 While many cancer types with APOBEC mutational signatures are solid tumor types, such as breast cancer, the similar role of APOBEC3A has also been shown in a liquid tumor, AML. 185 Furthermore, an apparently conserved group of AMLs has been identified with high levels of APOBEC3A expression. With subsequent in vitro cell culture studies demonstrating sensitization to DNA replication checkpoint inhibition, this AML outlier group defined by APOBEC3A expression may serve as an important link between clonal hematopoiesis, malignant transformation, and therapeutic approach. 186

7.3. Thymine DNA glycosylase

Conditional Tdg inactivation in Apc‐Min mice results in the increased frequency of small intestinal adenomas. Interestingly, this increase occurred in female mice, consistent with TCGA data showing that among colorectal cancer (CRC) cases, the lowest quartile of TDG and APC expression is associated with an excess of female cases. 187 As TDG has known interactions with an estrogen receptor, coupled with the fact that women with Lynch syndrome carry a higher lifetime risk of CRC, the connection between the estrogen hormone and colorectal oncogenesis may involve the demethylation machinery. 188 In the clinical literature, a recent case with a series of colorectal cancer patients identified a nearly fourfold risk increase in the patients carrying a SNP in TDG, rs4135113. 189 Notably, this G199S variant was shown to induce DNA double‐strand breaks and cellular transformation in cultured cells.

A case of pediatric rectal cancer was reported in the setting of constitutional PMS2 deficiency, one of the Lynch syndrome‐associated MMR genes. A heterozygous missense TDG mutation was noted in the tumor and subsequently confirmed by IHC to result in reduced TDG protein levels. Accordingly, the tumor DNA contained a number of C‐to‐A transition mutations, consistent with the mutagenesis pattern through the 5mC deamination. 190 Although pediatric colorectal cancer (CRC) is a very rare disease, with a reported incidence of 1 in 1 million, its biology has notable differences from adult CRC; as such, this case may be further evidence of a role for TDG and active demethylation as a cancer phenotype modifier, as opposed to a singular tumor suppressor role seen with TET2. To this point, another case series investigated TDG expression and gene methylation patterns in patients carrying a TP53 SNP that was associated with a form of Li‐Fraumeni syndrome. Referred to as the “Brazilian variant,” the pR337H mutation was discovered as a founder mutation in Brazil and has been associated with the increased incidence of adrenocortical carcinomas (ACC) in children. 191 Interestingly, increased TDG expression was observed in the ACC patients with the presence of the germline pR337H mutation, further suggesting a possible endocrine/hormonal relationship between TDG and cancer. 192

We have recently identified TDG as an anticancer target in melanoma. TDG knockdown induced cell cycle arrest and killing of melanoma cell lines 54 ; cells that escaped these two fates accumulated >4n DNA content and underwent senescence. 54 Concomitantly, gene set enrichment analysis of TDG knockdown melanoma cells revealed the enrichment of E2F target genes, which are known to be involved in cell cycle progression and DNA replication and repair. 54 Unexpectedly, upregulation of transcripts was also found, encoding immune regulators, secreted cytokines and proteases belonging to the inflammatory and immune response. 54 This suggests that the normal function of TDG suppresses inflammatory responses, presumably through modulation of DNA methylation levels.

7.4. MBD4

Similar to TET1, MBD4 was discovered to have a foundational role in linking cytosine methylation with oncogenesis. Early clinical reports showed a strong correlation between MMR deficiency and secondary MBD4 mutations in Lynch syndrome spectrum cancers (CRC, endometrial, and pancreatic cancer). 28 , 193 , 194 , 195 , 196 Subsequent research highlighted the epigenetic significance of MBD4 as protecting against 5mC transition mutations to thymine, specifically through the binding of MBD4 to 5mC deamination products as well as to MLH1 protein directly. 25 , 197 , 198 , 199 , 200 These studies are in hindsight the first to link high mutational burden tumor types with epigenetics, with CRC serving as the most popular model. Interestingly, a recent case was reported of early onset rectosigmoid cancer with a germline heterozygous MBD4 mutation in the absence of Lynch syndrome or other germline colorectal cancer predisposing mutation. 201 Compared to the aforementioned case of pediatric rectal cancer with heterozygous TDG mutation, these cases appear to highlight the similar but nonredundant features of the 5mC maintenance enzymes, which have been established at both biochemical and model system levels.

More recent clinical studies have presented the significance of MBD4 as a “cancer phenotype modifier,” including the hematopoietic system. A recent article by Sanders et al described how the accumulation of C‐to‐T transition mutations due to germline loss of MBD4 resulted in the specific progression of clonal hematopoiesis to the early onset of AML. 202 Perhaps most significantly, the presence of germline MBD4 mutations in uveal melanoma reintroduced the clinical significance of mutational burden in the dawning era of immunotherapy. In addition to two independent reports demonstrating significant “outlier” responses of uveal melanoma to pembrolizumab in the setting of germline MBD4 mutations, the direct mechanism of pathogenesis also has begun to be uncovered. 203 , 204 , 205 Indeed, a recent comprehensive analysis of MBD4 mutations in uveal melanoma confirmed the role of MBD4 as a tumor suppressor in line with Knudson's two‐hit hypothesis, and convincingly joined BAP1 as the only identified predisposing genes for uveal melanoma. 206 Future studies may reveal whether the mechanism of tumor suppression by MBD4 is through addressing spontaneous deamination of 5mC or through a more complex system involving AID/APOBEC enzymes.

8. CURRENT AND POTENTIAL ROLE FOR METHYLOME ALTERING THERAPY: DNA HYPOMETHYLATING AGENTS

With such a large physiological investment in maintaining DNA methylation homeostasis in normal cells, and the risk of oncogenesis when disrupted, a natural question would be: can modulating the DNA methylation status be a good therapeutic approach for cancer? The answer appears to be an unequivocal yes, as supported by the efficacy of hypomethylating (demethylating) drugs in specific contexts of cancer.

Decitabine (5‐aza‐deoxycytidine, tradename Dacogen) was first synthesized in 1964. By substituting a nitrogen for the five position carbon in cytosine, the so‐called “azanucleoside” incorporates into DNA and forms a complex with DNMT enzymes; the complex can prevent post‐enzymatic release causing degradation of the enzyme. 207 Decitabine entered the research consciousness after being demonstrated to have anticancer properties in the 1980s. After a series of trials over the 1990s, which demonstrated its therapeutic potential in AML, pivotal trials using lower doses, as opposed to maximally tolerated doses, showed the differentiation and antiproliferative effects of this drug through hypomethylating effects. By 2006, decitabine was approved for MDS and CMML. 208 , 209

The beginning of the past decade was marked by mixed opinions regarding decitabine as a treatment for AML. A phase III trial of decitabine vs supportive care or cytarabine (treatment choice) in patients >65 years of age reported an improved complete remission (including without platelet recovery) of 17.8% with decitabine vs 7.8% in controls, and a nonsignificant increase in the primary analysis of median overall survival to 7.7 months from 5.0 months. As elderly AML patients have poorer outcomes and are often ineligible for standard induction therapy, this study was a notable investigation into alternative approaches. The FDA ultimately did not approve decitabine for AML in elderly patients, whereas its counterpart in Europe approved it for de novo or secondary AML with poor or intermediate risk cytogenetics. 210

Over the past 5 years, the therapeutic role of hypomethylating agents has been developed in combination therapy. Landmark studies led by DiNardo et al showed that when hypomethylating agents were used in combination with venetoclax, a BCL‐2 inhibitor, a remarkable increase in complete remission (73%) and overall survival (17.5 months) was achieved. 211 , 212 Whereas these studies were nonrandomized and open label, a randomized double‐blind phase 3 trial comparing decitabine in combination with venetoclax vs decitabine with placebo is currently ongoing (NCT02993523). 213

Several reports have alluded to the potential for hypomethylating agents in combination therapy beyond de novo AML. In one report of therapy‐related AML in patients with advanced solid tumors, the complete remission rate of the combination of venetoclax and decitabine was comparable to the rates in de novo AML. While therapy‐related AML is uncommon, it is associated with aggressive cytogenetic features, and is frequently disqualified from clinical trials. 214

Hypomethylating agents have intriguing and theoretical potential for combination therapy. Hypomethylation has an established effect of endogenous retroviral activation and expression of other anti‐tumor genes, with an overall immune response that has the potential to convert immunotherapy resistant AML to sensitive one. 215 A nonrandomized open label study investigating azacitadine combined with nivolumab in relapsed/refractory AML reported improved remission rates (22%), compared to the reported rates of 10% to 16% for hypomethylating agents used as single‐agent therapy. 216 Currently, there are five ongoing clinical trials investigating combination immunotherapy with hypomethylating agents for AML. For more detail regarding immunotherapy for AML, with and without hypomethylating agent combination, readers are referred to recent reviews. 215 , 217 Combination hypomethylating agents and immunotherapy outside of AML is also being investigated, at both the in vitro and the clinical level. Patient‐derived xenograft studies on microsatellite stable colorectal cancer, which is less sensitive to immunotherapy than the microsatellite instability counterpart, showed an improved response to the programmed death‐1 (PD‐1) blockade due to the increased expression of antigen presentation‐related genes. 218 In a very recent clinical report, combining decitabine with camrelizumab, an anti‐PD‐1 antibody, demonstrated significantly higher complete remission rates in relapsed/refractory classical Hodgkin lymphoma compared to camrelizumab alone. 219

9. CONCLUSIONS

As summarized in this review, multiple epigenetic mechanisms modulate cytosine within the genome to produce significant effects on development, immune function, and oncogenesis. As studies continue to demonstrate the effects of cytosine methylation on each of these processes, the opportunities to apply hypomethylating therapies increase. Similarly, as the relationships between immune maturation and cancer are strengthened, the potential for aberrant deamination to play a significant role has begun to emerge. By combining the perspectives of physiological C/5mC modification and the processes of deamination and demethylation, the potential for novel therapies as well as the optimization of therapies that are currently available will likely be more effectively realized.

CONFLICT OF INTEREST

The authors declare no conflict of interest.

AUTHOR CONTRIBUTIONS

Rahul Prasad, Timothy J. Yen, and Alfonso Bellacosa: Conceived and wrote the review. Rahul Prasad and Alfonso Bellacosa: Prepared the figures. Rahul Prasad: Prepared the tables.

ACKNOWLEDGMENTS

The authors would like to thank Dr. Priscilla K. Cooper for the concept of the “baseball” figure of demethylation/deamination (cover photo) and Ms Karen Trush for drawing it. This study was supported by NIH grants CA191956 (to A. B. and T. J. Y.) and CA06927 (to the Fox Chase Cancer Center), DOD grant W81XWH‐17‐1‐0136 (to A. B. and T. J. Y.), Melanoma Research Alliance grant 693670 (to A. B.), PA Cure grant # SAP4100079728 (to T. J. Y.), a generous gift from Millie and Jules Roden Foundation (to T. J. Y.), John and Robin Spurlino, the Edwards Family (to A. B.), and an appropriation from the Commonwealth of Pennsylvania to the Fox Chase Cancer Center.

Prasad R, Yen TJ, Bellacosa A. Active DNA demethylation—The epigenetic gatekeeper of development, immunity, and cancer. Advanced Genetics. 2021;2:e10033. 10.1002/ggn2.10033

Funding information Commonwealth of Pennsylvania; DOD, Grant/Award Number: W81XWH‐17‐1‐0136; Edwards Family; John and Robin Spurlino; Melanoma Research Alliance, Grant/Award Number: 693670; Millie and Jules Roden Foundation; National Cancer Institute, Grant/Award Numbers: CA06927, CA191956; PA Cure, Grant/Award Number: SAP4100079728

Contributor Information

Timothy J. Yen, Email: timothy.yen@fccc.edu.

Alfonso Bellacosa, Email: alfonso.bellacosa@fccc.edu.

DATA AVAILABILITY STATEMENT

Data sharing not applicable to this article as no datasets were generated or analyzed during the current study.

REFERENCES

  • 1. Ehrlich M. DNA methylation in cancer: too much, but also too little. Oncogene. 2002;21(35):5400‐5413. [DOI] [PubMed] [Google Scholar]
  • 2. Mahmood N, Rabbani SA. DNA methylation readers and cancer: mechanistic and therapeutic applications. Front Oncol. 2019;9:489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Wajed SA, Laird PW, DeMeester TR. DNA methylation: an alternative pathway to cancer. Ann Surg. 2001;234(1):10‐20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Dante R, Dante‐Paire J, Rigal D, Roizes G. Methylation patterns of long interspersed repeated DNA and alphoid repetitive DNA from human cell lines and tumors. Anticancer Res. 1992;12(2):559‐563. [PubMed] [Google Scholar]
  • 5. Jurgens B, Schmitz‐Drager BJ, Schulz WA. Hypomethylation of L1 LINE sequences prevailing in human urothelial carcinoma. Cancer Res. 1996;56(24):5698‐5703. [PubMed] [Google Scholar]
  • 6. Santourlidis S, Florl A, Ackermann R, Wirtz HC, Schulz WA. High frequency of alterations in DNA methylation in adenocarcinoma of the prostate. Prostate. 1999;39(3):166‐174. [DOI] [PubMed] [Google Scholar]
  • 7. Takai D, Yagi Y, Habib N, Sugimura T, Ushijima T. Hypomethylation of LINE1 retrotransposon in human hepatocellular carcinomas, but not in surrounding liver cirrhosis. Jpn J Clin Oncol. 2000;30(7):306‐309. [DOI] [PubMed] [Google Scholar]
  • 8. Tahiliani M, Koh KP, Shen Y, et al. Conversion of 5‐methylcytosine to 5‐hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science. 2009;324(5929):930‐935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. He YF, Li BZ, Li Z, et al. Tet‐mediated formation of 5‐carboxylcytosine and its excision by TDG in mammalian DNA. Science. 2011;333(6047):1303‐1307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Ito S, Shen L, Dai Q, et al. Tet proteins can convert 5‐methylcytosine to 5‐formylcytosine and 5‐carboxylcytosine. Science. 2011;333(6047):1300‐1303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. 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(41):35334‐35338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Muller U, Bauer C, Siegl M, Rottach A, Leonhardt H. TET‐mediated oxidation of methylcytosine causes TDG or NEIL glycosylase dependent gene reactivation. Nucleic Acids Res. 2014;42(13):8592‐8604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Slyvka A, Mierzejewska K, Bochtler M. Nei‐like 1 (NEIL1) excises 5‐carboxylcytosine directly and stimulates TDG‐mediated 5‐formyl and 5‐carboxylcytosine excision. Sci Rep. 2017;7(1):9001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Hickson ID. General introduction to DNA base excision repair. In: Hickson ID, ed. Base Excision Repair of DNA Damage. Austin, TX: Landes Bioscience (Springer‐Verlag); 1997:1‐5. [Google Scholar]
  • 15. Cortazar D, Kunz C, Selfridge J, et al. Embryonic lethal phenotype reveals a function of TDG in maintaining epigenetic stability. Nature. 2011;470:419‐423. [DOI] [PubMed] [Google Scholar]
  • 16. Cortellino S, Xu J, Sannai M, et al. Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination‐base excision repair. Cell. 2011;146(1):67‐79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Hackett JA, Sengupta R, Zylicz JJ, et al. Germline DNA demethylation dynamics and imprint erasure through 5‐hydroxymethylcytosine. Science. 2013;339(6118):448‐452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Jin C, Qin T, Barton MC, Jelinek J, Issa JP. Minimal role of base excision repair in TET‐induced global DNA demethylation in HEK293T cells. Epigenetics. 2015;10(11):1006‐1013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Otani J, Kimura H, Sharif J, et al. Cell cycle‐dependent turnover of 5‐hydroxymethyl cytosine in mouse embryonic stem cells. PLoS One. 2013;8(12):e82961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Tsagaratou A, Lio CJ, Yue X, Rao A. TET methylcytosine oxidases in T cell and B cell development and function. Front Immunol. 2017;8:220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Seiler CL, Fernandez J, Koerperich Z, et al. Maintenance DNA methyltransferase activity in the presence of oxidized forms of 5‐methylcytosine: structural basis for ten eleven translocation‐mediated DNA demethylation. Biochemistry. 2018;57(42):6061‐6069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Pearl LH, Savva R. The problem with pyrimidines. Nat Struct Biol. 1996;3(6):485‐487. [DOI] [PubMed] [Google Scholar]
  • 23. Shen JC, Rideout WM, Jones PA. The rate of hydrolytic deamination of 5‐methylcytosine in double‐stranded DNA. Nucleic Acids Res. 1994;22(6):972‐976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Girelli Zubani G, Zivojnovic M, de Smet A, et al. Pms2 and uracil‐DNA glycosylases act jointly in the mismatch repair pathway to generate Ig gene mutations at A‐T base pairs. J Exp Med. 2017;214(4):1169‐1180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Millar CB, Guy J, Sansom OJ, et al. Enhanced CpG mutability and tumorigenesis in MBD4‐deficient mice. Science. 2002;297(5580):403‐405. [DOI] [PubMed] [Google Scholar]
  • 26. Wong E, Yang K, Kuraguchi M, et al. Mbd4 inactivation increases C→T transition mutations and promotes gastrointestinal tumor formation. Proc Natl Acad Sci U S A. 2002;99(23):14937‐14942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Cortellino S, Turner D, Masciullo V, et al. The base excision repair enzyme MED1 mediates DNA damage response to antitumor drugs and is associated with mismatch repair system integrity. Proc Natl Acad Sci U S A. 2003;100(25):15071‐15076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Tricarico R, Cortellino S, Riccio A, et al. Involvement of MBD4 inactivation in mismatch repair‐deficient tumorigenesis. Oncotarget. 2015;6(40):42892‐42904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Sabag O, Zamir A, Keshet I, et al. Establishment of methylation patterns in ES cells. Nat Struct Mol Biol. 2014;21(1):110‐112. [DOI] [PubMed] [Google Scholar]
  • 30. Rai K, Huggins IJ, James SR, Karpf AR, Jones DA, Cairns BR. DNA demethylation in zebrafish involves the coupling of a deaminase, a glycosylase, and gadd45. Cell. 2008;135(7):1201‐1212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Shimoda N, Hirose K, Kaneto R, et al. No evidence for AID/MBD4‐coupled DNA demethylation in zebrafish embryos. PLoS One. 2014;9(12):e114816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Nabel CS, Jia H, Ye Y, et al. AID/APOBEC deaminases disfavor modified cytosines implicated in DNA demethylation. Nat Chem Biol. 2012;8(9):751‐758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Schutsky EK, Nabel CS, Davis AKF, DeNizio JE, Kohli RM. APOBEC3A efficiently deaminates methylated, but not TET‐oxidized, cytosine bases in DNA. Nucleic Acids Res. 2017;45(13):7655‐7665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Arab K, Karaulanov E, Musheev M, et al. GADD45A binds R‐loops and recruits TET1 to CpG Island promoters. Nat Genet. 2019;51(2):217‐223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Pfaffeneder T, Spada F, Wagner M, et al. Tet oxidizes thymine to 5‐hydroxymethyluracil in mouse embryonic stem cell DNA. Nat Chem Biol. 2014;10(7):574‐581. [DOI] [PubMed] [Google Scholar]
  • 36. Guo JU, Su Y, Zhong C, Ming G‐L, Song H. Hydroxylation of 5‐methylcytosine by TET1 promotes active DNA demethylation in the adult brain. Cell. 2011;145(3):423‐434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Franchini DM, Chan CF, Morgan H, et al. Processive DNA demethylation via DNA deaminase‐induced lesion resolution. PLoS One. 2014;9(7):e97754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Tomkova M, McClellan M, Kriaucionis S, Schuster‐Boeckler B. 5‐hydroxymethylcytosine marks regions with reduced mutation frequency in human DNA. Elife. 2016;5:e17082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Supek F, Lehner B, Hajkova P, Warnecke T. Hydroxymethylated cytosines are associated with elevated C to G transversion rates. PLoS Genet. 2014;10(9):e1004585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Cimmino L, Dawlaty MM, Ndiaye‐Lobry D, et al. TET1 is a tumor suppressor of hematopoietic malignancy. Nat Immunol. 2015;16(6):653‐662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Mouly E, Ghamlouch H, Della‐Valle V, et al. B‐cell tumor development in Tet2‐deficient mice. Blood Adv. 2018;2(6):703‐714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Pan F, Wingo TS, Zhao Z, et al. Tet2 loss leads to hypermutagenicity in haematopoietic stem/progenitor cells. Nat Commun. 2017;8:15102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Lopez‐Moyado IF, Tsagaratou A, Yuita H, et al. Paradoxical association of TET loss of function with genome‐wide DNA hypomethylation. Proc Natl Acad Sci U S A. 2019;116(34):16933‐16942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Schoeler K, Aufschnaiter A, Messner S, et al. TET enzymes control antibody production and shape the mutational landscape in germinal Centre B cells. FEBS J. 2019;286(18):3566‐3581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Buisson R, Langenbucher A, Bowen D, et al. Passenger hotspot mutations in cancer driven by APOBEC3A and mesoscale genomic features. Science. 2019;364(6447):eaaw2872. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Rogozin IB, Lada AG, Goncearenco A, et al. Activation induced deaminase mutational signature overlaps with CpG methylation sites in follicular lymphoma and other cancers. Sci Rep. 2016;6:38133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Seplyarskiy VB, Soldatov RA, Popadin KY, Antonarakis SE, Bazykin GA, Nikolaev SI. APOBEC‐induced mutations in human cancers are strongly enriched on the lagging DNA strand during replication. Genome Res. 2016;26(2):174‐182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Burns MB, Temiz NA, Harris RS. Evidence for APOBEC3B mutagenesis in multiple human cancers. Nat Genet. 2013;45(9):977‐983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Rebhandl S, Huemer M, Gassner FJ, et al. APOBEC3 signature mutations in chronic lymphocytic leukemia. Leukemia. 2014;28(9):1929‐1932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Frommer M, McDonald LE, Millar DS, et al. A genomic sequencing protocol that yields a positive display of 5‐methylcytosine residues in individual DNA strands. Proc Natl Acad Sci U S A. 1992;89(5):1827‐1831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Clark SJ, Harrison J, Paul CL, Frommer M. High sensitivity mapping of methylated cytosines. Nucleic Acids Res. 1994;22(15):2990‐2997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. 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(7310):1129‐1133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Dalton SR, Bellacosa A. DNA demethylation by TDG. Epigenomics. 2012;4(4):459‐467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Mancuso P, Tricarico R, Bhattacharjee V, et al. Thymine DNA glycosylase as a novel target for melanoma. Oncogene. 2019;38(19):3710‐3728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Tricarico R, Madzo J, Scher G, et al. TET1 and TDG suppress intestinal tumorigenesis by down‐regulating the inflammatory and immune response pathways. bioRxiv. 2019. 10.1101/676445. [DOI] [Google Scholar]
  • 56. Nanan KK, Sturgill DM, Prigge MF, et al. TET‐catalyzed 5‐carboxylcytosine promotes CTCF binding to suboptimal sequences genome‐wide. iScience. 2019;19:326‐339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Ji S, Park D, Kropachev K, et al. 5‐Formylcytosine‐induced DNA‐peptide cross‐links reduce transcription efficiency, but do not cause transcription errors in human cells. J Biol Chem. 2019;294(48):18387‐18397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. Li F, Zhang Y, Bai J, Greenberg MM, Xi Z, Zhou C. 5‐Formylcytosine yields DNA‐protein cross‐links in nucleosome core particles. J Am Chem Soc. 2017;139(31):10617‐10620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. Raiber EA, Portella G, Martinez Cuesta S, et al. 5‐Formylcytosine organizes nucleosomes and forms Schiff base interactions with histones in mouse embryonic stem cells. Nat Chem. 2018;10(12):1258‐1266. [DOI] [PubMed] [Google Scholar]
  • 60. Deckard CE III, Banerjee DR, Sczepanski JT. Chromatin structure and the pioneering transcription factor FOXA1 regulate TDG‐mediated removal of 5‐formylcytosine from DNA. J Am Chem Soc. 2019;141(36):14110‐14114. [DOI] [PubMed] [Google Scholar]
  • 61. Cirillo LA, Lin FR, Cuesta I, Friedman D, Jarnik M, Zaret KS. Opening of compacted chromatin by early developmental transcription factors HNF3 (FoxA) and GATA‐4. Mol Cell. 2002;9(2):279‐289. [DOI] [PubMed] [Google Scholar]
  • 62. Hashimoto H, Olanrewaju YO, Zheng Y, Wilson GG, Zhang X, Cheng X. Wilms tumor protein recognizes 5‐carboxylcytosine within a specific DNA sequence. Genes Dev. 2014;28(20):2304‐2313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63. Sayeed SK, Zhao J, Sathyanarayana BK, Golla JP, Vinson C. C/EBPbeta (CEBPB) protein binding to the C/EBP|CRE DNA 8‐mer TTGC|GTCA is inhibited by 5hmC and enhanced by 5mC, 5fC, and 5caC in the CG dinucleotide. Biochim Biophys Acta. 2015;1849(6):583‐589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Wang D, Hashimoto H, Zhang X, et al. MAX is an epigenetic sensor of 5‐carboxylcytosine and is altered in multiple myeloma. Nucleic Acids Res. 2017;45(5):2396‐2407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Yang J, Horton JR, Li J, et al. Structural basis for preferential binding of human TCF4 to DNA containing 5‐carboxylcytosine. Nucleic Acids Res. 2019;47(16):8375‐8387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66. Song CX, Szulwach KE, Dai Q, et al. Genome‐wide profiling of 5‐formylcytosine reveals its roles in epigenetic priming. Cell. 2013;153(3):678‐691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Mohan RD, Litchfield DW, Torchia J, Tini M. Opposing regulatory roles of phosphorylation and acetylation in DNA mispair processing by thymine DNA glycosylase. Nucleic Acids Res. 2010;38(4):1135‐1148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68. Mohan RD, Rao A, Gagliardi J, Tini M. SUMO‐1‐dependent allosteric regulation of thymine DNA glycosylase alters subnuclear localization and CBP/p300 recruitment. J Mol Cell Biol. 2007;27(1):229‐243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69. Tini M, Benecke A, Um SJ, Torchia J, Evans RM, Chambon P. Association of CBP/p300 acetylase and thymine DNA glycosylase links DNA repair and transcription. Mol Cell. 2002;9(2):265‐277. [DOI] [PubMed] [Google Scholar]
  • 70. Henry RA, Mancuso P, Kuo YM, et al. Interaction with the DNA repair protein thymine DNA glycosylase regulates histone acetylation by p300. Biochemistry. 2016;55(49):6766‐6775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Ito S, Kuraoka I. Epigenetic modifications in DNA could mimic oxidative DNA damage: a double‐edged sword. DNA Repair. 2015;32:52‐57. [DOI] [PubMed] [Google Scholar]
  • 72. Kamiya H, Tsuchiya H, Karino N, Ueno Y, Matsuda A, Harashima H. Mutagenicity of 5‐formylcytosine, an oxidation product of 5‐methylcytosine, in DNA in mammalian cells. Int J Biochem. 2002;132(4):551‐555. [DOI] [PubMed] [Google Scholar]
  • 73. Karino N, Ueno Y, Matsuda A. Synthesis and properties of oligonucleotides containing 5‐formyl‐2′‐deoxycytidine: in vitro DNA polymerase reactions on DNA templates containing 5‐formyl‐2′‐deoxycytidine. Nucleic Acids Res. 2001;29(12):2456‐2463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74. Munzel M, Lischke U, Stathis D, et al. Improved synthesis and mutagenicity of oligonucleotides containing 5‐hydroxymethylcytosine, 5‐formylcytosine and 5‐carboxylcytosine. Chemistry. 2011;17(49):13782‐13788. [DOI] [PubMed] [Google Scholar]
  • 75. Thomson JP, Moggs JG, Wolf CR, Meehan RR. Epigenetic profiles as defined signatures of xenobiotic exposure. Mutat Res Genet Toxicol Environ Mutagen. 2014;764‐765:3‐9. [DOI] [PubMed] [Google Scholar]
  • 76. Zarakowska E, Gackowski D, Foksinski M, Olinski R. Are 8‐oxoguanine (8‐oxoGua) and 5‐hydroxymethyluracil (5‐hmUra) oxidatively damaged DNA bases or transcription (epigenetic) marks? Mutat Res Genet Toxicol Environ Mutagen. 2014;764‐765:58‐63. [DOI] [PubMed] [Google Scholar]
  • 77. Purmal AA, Kow YW, Wallace SS. Major oxidative products of cytosine, 5‐hydroxycytosine and 5‐hydroxyuracil, exhibit sequence context‐dependent mispairing in vitro. Nucleic Acids Res. 1994;22(1):72‐78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Wallace SS. DNA damages processed by base excision repair: biological consequences. Int J Radiat Biol. 1994;66(5):579‐589. [DOI] [PubMed] [Google Scholar]
  • 79. Wallace SS. Biological consequences of free radical‐damaged DNA bases. Free Radic Biol Med. 2002;33(1):1‐14. [DOI] [PubMed] [Google Scholar]
  • 80. Shibutani T, Ito S, Toda M, et al. Guanine‐ 5‐carboxylcytosine base pairs mimic mismatches during DNA replication. Sci Rep. 2014;4:5220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81. Kellinger MW, Song CX, Chong J, Lu XY, He C, Wang D. 5‐formylcytosine and 5‐carboxylcytosine reduce the rate and substrate specificity of RNA polymerase II transcription. Nat Struct Mol Biol. 2012;19(8):831‐833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82. Lewis LC, Lo PC, Foster JM, et al. Dynamics of 5‐carboxylcytosine during hepatic differentiation: potential general role for active demethylation by DNA repair in lineage specification. Epigenetics. 2017;12(4):277‐286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83. Iurlaro M, McInroy GR, Burgess HE, et al. In vivo genome‐wide profiling reveals a tissue‐specific role for 5‐formylcytosine. Genome Biol. 2016;17(1):141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84. Wheldon LM, Abakir A, Ferjentsik Z, et al. Transient accumulation of 5‐carboxylcytosine indicates involvement of active demethylation in lineage specification of neural stem cells. Cell Rep. 2014;7(5):1353‐1361. [DOI] [PubMed] [Google Scholar]
  • 85. Abdel‐Wahab O, Mullally A, Hedvat C, et al. Genetic characterization of TET1, TET2, and TET3 alterations in myeloid malignancies. Blood. 2009;114(1):144‐147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Cancer Genome Atlas Research Network , Ley TJ, Miller C, et al. Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia. N Engl J Med. 2013;368(22):2059‐2074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87. Klco JM, Miller CA, Griffith M, et al. Association between mutation clearance after induction therapy and outcomes in acute myeloid leukemia. JAMA. 2015;314(8):811‐822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88. Ley TJ, Ding L, Walter MJ, et al. DNMT3A mutations in acute myeloid leukemia. N Engl J Med. 2010;363(25):2424‐2433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Nibourel O, Kosmider O, Cheok M, et al. Incidence and prognostic value of TET2 alterations in de novo acute myeloid leukemia achieving complete remission. Blood. 2010;116(7):1132‐1135. [DOI] [PubMed] [Google Scholar]
  • 90. Tyner JW, Tognon CE, Bottomly D, et al. Functional genomic landscape of acute myeloid leukaemia. Nature. 2018;562(7728):526‐531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Weissmann S, Alpermann T, Grossmann V, et al. Landscape of TET2 mutations in acute myeloid leukemia. Leukemia. 2012;26(5):934‐942. [DOI] [PubMed] [Google Scholar]
  • 92. Delhommeau F, Dupont S, Della Valle V, et al. Mutation in TET2 in myeloid cancers. N Engl J Med. 2009;360(22):2289‐2301. [DOI] [PubMed] [Google Scholar]
  • 93. Kosmider O, Gelsi‐Boyer V, Ciudad M, et al. TET2 gene mutation is a frequent and adverse event in chronic myelomonocytic leukemia. Haematologica. 2009;94(12):1676‐1681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94. Lin Y, Lin Z, Cheng K, et al. Prognostic role of TET2 deficiency in myelodysplastic syndromes: a meta‐analysis. Oncotarget. 2017;8(26):43295‐43305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Grossmann V, Kohlmann A, Eder C, et al. Molecular profiling of chronic myelomonocytic leukemia reveals diverse mutations in >80% of patients with TET2 and EZH2 being of high prognostic relevance. Leukemia. 2011;25(5):877‐879. [DOI] [PubMed] [Google Scholar]
  • 96. Kohlmann A, Grossmann V, Klein HU, et al. Next‐generation sequencing technology reveals a characteristic pattern of molecular mutations in 72.8% of chronic myelomonocytic leukemia by detecting frequent alterations in TET2, CBL, RAS, and RUNX1. J Clin Oncol. 2010;28(24):3858‐3865. [DOI] [PubMed] [Google Scholar]
  • 97. Makishima H, Jankowska AM, McDevitt MA, et al. CBL, CBLB, TET2, ASXL1, and IDH1/2 mutations and additional chromosomal aberrations constitute molecular events in chronic myelogenous leukemia. Blood. 2011;117(21):e198‐e206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Patnaik MM, Barraco D, Lasho TL, et al. DNMT3A mutations are associated with inferior overall and leukemia‐free survival in chronic myelomonocytic leukemia. Am J Hematol. 2017;92(1):56‐61. [DOI] [PubMed] [Google Scholar]
  • 99. Patnaik MM, Zahid MF, Lasho TL, et al. Number and type of TET2 mutations in chronic myelomonocytic leukemia and their clinical relevance. Blood Cancer J. 2016;6(9):e472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Yamazaki J, Taby R, Vasanthakumar A, et al. Effects of TET2 mutations on DNA methylation in chronic myelomonocytic leukemia. Epigenetics. 2012;7(2):201‐207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Quivoron C, Couronne L, Della Valle V, et al. TET2 inactivation results in pleiotropic hematopoietic abnormalities in mouse and is a recurrent event during human lymphomagenesis. Cancer Cell. 2011;20(1):25‐38. [DOI] [PubMed] [Google Scholar]
  • 102. Couronne L, Bastard C, Bernard OA. TET2 and DNMT3A mutations in human T‐cell lymphoma. N Engl J Med. 2012;366(1):95‐96. [DOI] [PubMed] [Google Scholar]
  • 103. Lemonnier F, Couronne L, Parrens M, et al. Recurrent TET2 mutations in peripheral T‐cell lymphomas correlate with TFH‐like features and adverse clinical parameters. Blood. 2012;120(7):1466‐1469. [DOI] [PubMed] [Google Scholar]
  • 104. Palomero T, Couronne L, Khiabanian H, et al. Recurrent mutations in epigenetic regulators, RHOA and FYN kinase in peripheral T cell lymphomas. Nat Genet. 2014;46(2):166‐170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105. Sakata‐Yanagimoto M, Enami T, Yoshida K, et al. Somatic RHOA mutation in angioimmunoblastic T cell lymphoma. Nat Genet. 2014;46(2):171‐175. [DOI] [PubMed] [Google Scholar]
  • 106. Cairns RA, Iqbal J, Lemonnier F, et al. IDH2 mutations are frequent in angioimmunoblastic T‐cell lymphoma. Blood. 2012;119(8):1901‐1903. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107. Dobay MP, Lemonnier F, Missiaglia E, et al. Integrative clinicopathological and molecular analyses of angioimmunoblastic T‐cell lymphoma and other nodal lymphomas of follicular helper T‐cell origin. Haematologica. 2017;102(4):e148‐e151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Wang C, McKeithan TW, Gong Q, et al. IDH2R172 mutations define a unique subgroup of patients with angioimmunoblastic T‐cell lymphoma. Blood. 2015;126(15):1741‐1752. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109. Kubuki Y, Yamaji T, Hidaka T, et al. TET2 mutation in diffuse large B‐cell lymphoma. J Clin Exp Hematop. 2017;56(3):145‐149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Asmar F, Punj V, Christensen J, et al. Genome‐wide profiling identifies a DNA methylation signature that associates with TET2 mutations in diffuse large B‐cell lymphoma. Haematologica. 2013;98(12):1912‐1920. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Chapuy B, Stewart C, Dunford AJ, et al. Molecular subtypes of diffuse large B cell lymphoma are associated with distinct pathogenic mechanisms and outcomes. Nat Med. 2018;24(5):679‐690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Dominguez PM, Ghamlouch H, Rosikiewicz W, et al. TET2 deficiency causes germinal center hyperplasia, impairs plasma cell differentiation, and promotes B‐cell lymphomagenesis. Cancer Discov. 2018;8(12):1632‐1653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Hoadley KA, Yau C, Hinoue T, et al. Cell‐of‐origin patterns dominate the molecular classification of 10,000 tumors from 33 types of cancer. Cell. 2018;173(2):291‐304.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114. Karube K, Enjuanes A, Dlouhy I, et al. Integrating genomic alterations in diffuse large B‐cell lymphoma identifies new relevant pathways and potential therapeutic targets. Leukemia. 2018;32(3):675‐684. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115. Lohr JG, Stojanov P, Lawrence MS, et al. Discovery and prioritization of somatic mutations in diffuse large B‐cell lymphoma (DLBCL) by whole‐exome sequencing. Proc Natl Acad Sci U S A. 2012;109(10):3879‐3884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Morin RD, Mungall K, Pleasance E, et al. Mutational and structural analysis of diffuse large B‐cell lymphoma using whole‐genome sequencing. Blood. 2013;122(7):1256‐1265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Reddy A, Zhang J, Davis NS, et al. Genetic and functional drivers of diffuse large B cell lymphoma. Cell. 2017;171(2):481‐94.e15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118. Schmitz R, Wright GW, Huang DW, et al. Genetics and pathogenesis of diffuse large B‐cell lymphoma. N Engl J Med. 2018;378(15):1396‐1407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Cancer Genome Atlas Research Network , Brat DJ, Verhaak RG, et al. Comprehensive, integrative genomic analysis of diffuse lower‐grade gliomas. N Engl J Med. 2015;372(26):2481‐2498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120. Brennan CW, Verhaak RG, McKenna A, et al. The somatic genomic landscape of glioblastoma. Cell. 2013;155(2):462‐477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121. Cancer Genome Atlas Research Network . Comprehensive and integrative genomic characterization of hepatocellular carcinoma. Cell. 2017;169(7):1327‐41.e23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Borger DR, Tanabe KK, Fan KC, et al. Frequent mutation of isocitrate dehydrogenase (IDH)1 and IDH2 in cholangiocarcinoma identified through broad‐based tumor genotyping. Oncologist. 2012;17(1):72‐79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123. Farshidfar F, Zheng S, Gingras MC, et al. Integrative genomic analysis of cholangiocarcinoma identifies distinct IDH‐mutant molecular profiles. Cell Rep. 2017;19(13):2878‐2880. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. DiNardo CD, Beird HC, Estecio M, et al. Germline DNMT3A mutation in familial acute myeloid leukaemia. Epigenetics. 2020;1‐10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Shen W, Heeley JM, Carlston CM, et al. The spectrum of DNMT3A variants in Tatton‐Brown‐Rahman syndrome overlaps with that in hematologic malignancies. Am J Med Genet A. 2017;173(11):3022‐3028. [DOI] [PubMed] [Google Scholar]
  • 126. Xin B, Cruz Marino T, Szekely J, et al. Novel DNMT3A germline mutations are associated with inherited Tatton‐Brown‐Rahman syndrome. Clin Genet. 2017;91(4):623‐628. [DOI] [PubMed] [Google Scholar]
  • 127. Duployez N, Goursaud L, Fenwarth L, et al. Familial myeloid malignancies with germline TET2 mutation. Leukemia. 2020;34(5):1450‐1453. [DOI] [PubMed] [Google Scholar]
  • 128. Kaasinen E, Kuismin O, Rajamaki K, et al. Impact of constitutional TET2 haploinsufficiency on molecular and clinical phenotype in humans. Nat Commun. 2019;10(1):1252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129. Akiyama M, Yamaoka M, Mikami‐Terao Y, et al. Somatic mosaic mutations of IDH1 and NPM1 associated with cup‐like acute myeloid leukemia in a patient with Maffucci syndrome. Int J Hematol. 2015;102(6):723‐728. [DOI] [PubMed] [Google Scholar]
  • 130. Prokopchuk O, Andres S, Becker K, Holzapfel K, Hartmann D, Friess H. Maffucci syndrome and neoplasms: a case report and review of the literature. BMC Res Notes. 2016;9:126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131. Inoue A, Shen L, Dai Q, He C, Zhang Y. Generation and replication‐dependent dilution of 5fC and 5caC during mouse preimplantation development. Cell Res. 2011;21(12):1670‐1676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132. Lio CJ, Rao A. TET enzymes and 5hmC in adaptive and innate immune systems. Front Immunol. 2019;10:210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133. Figueroa ME, Abdel‐Wahab O, Lu C, et al. Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell. 2010;18(6):553‐567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134. Lio CJ, Yuita H, Rao A. Dysregulation of the TET family of epigenetic regulators in lymphoid and myeloid malignancies. Blood. 2019;134(18):1487‐1497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135. Bai X, Zhang H, Zhou Y, et al. TET1 promotes malignant progression of cholangiocarcinoma with IDH1 wild‐type. Hepatology. 2020. 10.1002/hep.31486 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136. Ji H, Ehrlich LI, Seita J, et al. Comprehensive methylome map of lineage commitment from haematopoietic progenitors. Nature. 2010;467(7313):338‐342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137. Kulis M, Merkel A, Heath S, et al. Whole‐genome fingerprint of the DNA methylome during human B cell differentiation. Nat Genet. 2015;47(7):746‐756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138. Broske AM, Vockentanz L, Kharazi S, et al. DNA methylation protects hematopoietic stem cell multipotency from myeloerythroid restriction. Nat Genet. 2009;41(11):1207‐1215. [DOI] [PubMed] [Google Scholar]
  • 139. Manoharan A, du Roure C, Rolink AG, Matthias P. De novo DNA methyltransferases Dnmt3a and Dnmt3b regulate the onset of Igkappa light chain rearrangement during early B‐cell development. Eur J Immunol. 2015;45(8):2343‐2355. [DOI] [PubMed] [Google Scholar]
  • 140. Ehrlich M. The ICF syndrome, a DNA methyltransferase 3B deficiency and immunodeficiency disease. Clin Immunol. 2003;109(1):17‐28. [DOI] [PubMed] [Google Scholar]
  • 141. Costa Y, Ding J, Theunissen TW, et al. NANOG‐dependent function of TET1 and TET2 in establishment of pluripotency. Nature. 2013;495(7441):370‐374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142. de la Rica L, Rodriguez‐Ubreva J, Garcia M, et al. PU.1 target genes undergo Tet2‐coupled demethylation and DNMT3b‐mediated methylation in monocyte‐to‐osteoclast differentiation. Genome Biol. 2013;14(9):R99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143. Guilhamon P, Eskandarpour M, Halai D, et al. Meta‐analysis of IDH‐mutant cancers identifies EBF1 as an interaction partner for TET2. Nat Commun. 2013;4:2166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144. 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;5:e18290. 10.7554/eLife.18290 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145. Sardina JL, Collombet S, Tian TV, et al. Transcription factors drive Tet2‐mediated enhancer demethylation to reprogram cell fate. Cell Stem Cell. 2018;23(6):905‐906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146. Wang Y, Xiao M, Chen X, et al. WT1 recruits TET2 to regulate its target gene expression and suppress leukemia cell proliferation. Mol Cell. 2015;57(4):662‐673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147. Xiong J, Zhang Z, Chen J, et al. Cooperative action between SALL4A and TET proteins in stepwise oxidation of 5‐methylcytosine. Mol Cell. 2016;64(5):913‐925. [DOI] [PubMed] [Google Scholar]
  • 148. Montagner S, Leoni C, Emming S, et al. TET2 regulates mast cell differentiation and proliferation through catalytic and non‐catalytic activities. Cell Rep. 2016;15(7):1566‐1579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Nestor CE, Lentini A, Hagg Nilsson C, et al. 5‐Hydroxymethylcytosine remodeling precedes lineage specification during differentiation of human CD4(+) T cells. Cell Rep. 2016;16(2):559‐570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150. Yue X, Trifari S, Aijo T, et al. Control of Foxp3 stability through modulation of TET activity. J Exp Med. 2016;213(3):377‐397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151. Suzuki T, Shimizu Y, Furuhata E, et al. RUNX1 regulates site specificity of DNA demethylation by recruitment of DNA demethylation machineries in hematopoietic cells. Blood Adv. 2017;1(20):1699‐1711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152. Kavli B, Andersen S, Otterlei M, et al. B cells from hyper‐IgM patients carrying UNG mutations lack ability to remove uracil from ssDNA and have elevated genomic uracil. J Exp Med. 2005;201(12):2011‐2021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153. Grigera F, Bellacosa A, Kenter AL. Complex relationship between mismatch repair proteins and MBD4 during immunoglobulin class switch recombination. PLoS One. 2013;8(10):e78370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154. Grigera F, Wuerffel R, Kenter AL. MBD4 facilitates immunoglobulin class switch recombination. J Mol Cell Biol. 2017;37(2):e00316‐16. 10.1128/MCB.00316-16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155. Cantaert T, Schickel JN, Bannock JM, et al. Activation‐induced cytidine deaminase expression in human B cell precursors is essential for central B cell tolerance. Immunity. 2015;43(5):884‐895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156. Kuraoka M, Holl TM, Liao D, et al. Activation‐induced cytidine deaminase mediates central tolerance in B cells. Proc Natl Acad Sci U S A. 2011;108(28):11560‐11565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157. Hagler KT, Lynch JW Jr. Paraneoplastic manifestations of lymphoma. Clin Lymphoma. 2004;5(1):29‐36. [DOI] [PubMed] [Google Scholar]
  • 158. Lio CJ, Shukla V, Samaniego‐Castruita D, et al. TET enzymes augment activation‐induced deaminase (AID) expression via 5‐hydroxymethylcytosine modifications at the Aicda superenhancer. Sci Immunol. 2019;4(34):eaau7523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159. Kunimoto H, McKenney AS, Meydan C, et al. Aid is a key regulator of myeloid/erythroid differentiation and DNA methylation in hematopoietic stem/progenitor cells. Blood. 2017;129(13):1779‐1790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160. Lorsbach RB, Moore J, Mathew S, Raimondi SC, Mukatira ST, Downing JR. TET1, a member of a novel protein family, is fused to MLL in acute myeloid leukemia containing the t(10;11)(q22;q23). Leukemia. 2003;17(3):637‐641. [DOI] [PubMed] [Google Scholar]
  • 161. Ono R, Taki T, Taketani T, Taniwaki M, Kobayashi H, Hayashi Y. LCX, leukemia‐associated protein with a CXXC domain, is fused to MLL in acute myeloid leukemia with trilineage dysplasia having t(10;11)(q22;q23). Cancer Res. 2002;62(14):4075‐4080. [PubMed] [Google Scholar]
  • 162. Wang R, Gao X, Yu L. The prognostic impact of tet oncogene family member 2 mutations in patients with acute myeloid leukemia: a systematic‐review and meta‐analysis. BMC Cancer. 2019;19(1):389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163. Kosmider O, Gelsi‐Boyer V, Cheok M, et al. TET2 mutation is an independent favorable prognostic factor in myelodysplastic syndromes (MDSs). Blood. 2009;114(15):3285‐3291. [DOI] [PubMed] [Google Scholar]
  • 164. Guo Z, Zhang SK, Zou Z, Fan RH, Lyu XD. Prognostic significance of TET2 mutations in myelodysplastic syndromes: a meta‐analysis. Leuk Res. 2017;58:102‐107. [DOI] [PubMed] [Google Scholar]
  • 165. Yoo HY, Sung MK, Lee SH, et al. A recurrent inactivating mutation in RHOA GTPase in angioimmunoblastic T cell lymphoma. Nat Genet. 2014;46(4):371‐375. [DOI] [PubMed] [Google Scholar]
  • 166. Shimoda K, Shide K, Kameda T, et al. TET2 mutation in adult T‐cell leukemia/lymphoma. J Clin Exp Hematop. 2015;55(3):145‐149. [DOI] [PubMed] [Google Scholar]
  • 167. Shingleton JR, Dave SS. TET2 deficiency sets the stage for B‐cell lymphoma. Cancer Discov. 2018;8(12):1515‐1517. [DOI] [PubMed] [Google Scholar]
  • 168. Steensma DP. Clinical consequences of clonal hematopoiesis of indeterminate potential. Blood Adv. 2018;2(22):3404‐3410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169. Buscarlet M, Provost S, Zada YF, et al. DNMT3A and TET2 dominate clonal hematopoiesis and demonstrate benign phenotypes and different genetic predispositions. Blood. 2017;130(6):753‐762. [DOI] [PubMed] [Google Scholar]
  • 170. Jaiswal S, Natarajan P, Silver AJ, et al. Clonal hematopoiesis and risk of atherosclerotic cardiovascular disease. N Engl J Med. 2017;377(2):111‐121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171. Fuster JJ, MacLauchlan S, Zuriaga MA, et al. Clonal hematopoiesis associated with TET2 deficiency accelerates atherosclerosis development in mice. Science. 2017;355(6327):842‐847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172. Dorsheimer L, Assmus B, Rasper T, et al. Association of mutations contributing to clonal hematopoiesis with prognosis in chronic ischemic heart failure. JAMA Cardiol. 2019;4(1):25‐33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173. Ortmann CA, Kent DG, Nangalia J, et al. Effect of mutation order on myeloproliferative neoplasms. N Engl J Med. 2015;372(7):601‐612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174. Bowman RL, Levine RL. TET2 in Normal and malignant hematopoiesis. Cold Spring Harb Perspect Med. 2017;7(8):a026518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175. Huang H, Jiang X, Li Z, et al. TET1 plays an essential oncogenic role in MLL‐rearranged leukemia. Proc Natl Acad Sci U S A. 2013;110(29):11994‐11999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176. Ko M, Huang Y, Jankowska AM, et al. Impaired hydroxylation of 5‐methylcytosine in myeloid cancers with mutant TET2. Nature. 2010;468(7325):839‐843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Good CR, Panjarian S, Kelly AD, et al. TET1‐mediated hypomethylation activates oncogenic signaling in triple‐negative breast cancer. Cancer Res. 2018;78(15):4126‐4137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178. Zhao Z, Chen L, Dawlaty MM, et al. Combined loss of Tet1 and Tet2 promotes B cell, but not myeloid malignancies, in mice. Cell Rep. 2015;13(8):1692‐1704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179. Teater M, Dominguez PM, Redmond D, et al. AICDA drives epigenetic heterogeneity and accelerates germinal center‐derived lymphomagenesis. Nat Commun. 2018;9(1):222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180. Jiao J, Jin Y, Zheng M, et al. AID and TET2 co‐operation modulates FANCA expression by active demethylation in diffuse large B cell lymphoma. Clin Exp Immunol. 2019;195(2):190‐201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181. Samstein RM, Lee CH, Shoushtari AN, et al. Tumor mutational load predicts survival after immunotherapy across multiple cancer types. Nat Genet. 2019;51(2):202‐206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182. Hamilton BK, Rybicki L, Hirsch C, et al. Mutation clonal burden and allogeneic hematopoietic cell transplantation outcomes in acute myeloid leukemia and myelodysplastic syndromes. Bone Marrow Transplant. 2019;54(8):1281‐1286. [DOI] [PubMed] [Google Scholar]
  • 183. Supek F, Lehner B. Clustered mutation signatures reveal that error‐prone DNA repair targets mutations to active genes. Cell. 2017;170(3):534‐47.e23. [DOI] [PubMed] [Google Scholar]
  • 184. Chan K, Roberts SA, Klimczak LJ, et al. An APOBEC3A hypermutation signature is distinguishable from the signature of background mutagenesis by APOBEC3B in human cancers. Nat Genet. 2015;47(9):1067‐1072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185. Nik‐Zainal S, Wedge DC, Alexandrov LB, et al. Association of a germline copy number polymorphism of APOBEC3A and APOBEC3B with burden of putative APOBEC‐dependent mutations in breast cancer. Nat Genet. 2014;46(5):487‐491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186. Green AM, Budagyan K, Hayer KE, et al. Cytosine deaminase APOBEC3A sensitizes leukemia cells to inhibition of the DNA replication checkpoint. Cancer Res. 2017;77(17):4579‐4588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187. Xu J, Cortellino S, Tricarico R, et al. Thymine DNA glycosylase (TDG) is involved in the pathogenesis of intestinal tumors with reduced APC expression. Oncotarget. 2017;8(52):89988‐89997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188. Liu Y, Duong W, Krawczyk C, et al. Oestrogen receptor β regulates epigenetic patterns at specific genomic loci through interaction with thymine DNA glycosylase. Epigenetics Chromatin. 2016;9:7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189. Reddy Parine N, Alanazi IO, Shaik JP, et al. TDG gene polymorphisms and their possible association with colorectal cancer: a case control study. J Oncol. 2019;2019:7091815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190. Vasovcak P, Krepelova A, Menigatti M, et al. Unique mutational profile associated with a loss of TDG expression in the rectal cancer of a patient with a constitutional PMS2 deficiency. DNA Repair. 2012;11(7):616‐623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191. Giacomazzi J, Selistre SG, Rossi C, et al. Li‐Fraumeni and Li‐Fraumeni‐like syndrome among children diagnosed with pediatric cancer in Southern Brazil. Cancer. 2013;119(24):4341‐4349. [DOI] [PubMed] [Google Scholar]
  • 192. Fortes FP, Kuasne H, Marchi FA, Miranda PM, Rogatto SR, Achatz MI. DNA methylation patterns of candidate genes regulated by thymine DNA glycosylase in patients with TP53 germline mutations. Braz J Med Biol Res. 2015;48(7):610‐615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193. Bader S, Walker M, Hendrich B, et al. Somatic frameshift mutations in the MBD4 gene of sporadic colon cancers with mismatch repair deficiency. Oncogene. 1999;18(56):8044‐8047. [DOI] [PubMed] [Google Scholar]
  • 194. Riccio A, Aaltonen LA, Godwin AK, et al. The DNA repair gene MBD4 (MED1) is mutated in human carcinomas with microsatellite instability. Nat Genet. 1999;23(3):266‐268. [DOI] [PubMed] [Google Scholar]
  • 195. Bader S, Walker M, Harrison D. Most microsatellite unstable sporadic colorectal carcinomas carry MBD4 mutations. Br J Cancer. 2000;83(12):1646‐1649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196. Chow E, Lipton L, Lynch E, et al. Hyperplastic polyposis syndrome: phenotypic presentations and the role of MBD4 and MYH. Gastroenterology. 2006;131(1):30‐39. [DOI] [PubMed] [Google Scholar]
  • 197. Hendrich B, Hardeland U, Ng HH, Jiricny J, Bird A. The thymine glycosylase MBD4 can bind to the product of deamination at methylated CpG sites. Nature. 1999;401(6750):301‐304. [DOI] [PubMed] [Google Scholar]
  • 198. Bellacosa A, Cicchillitti L, Schepis F, et al. MED1, a novel human methyl‐CpG‐binding endonuclease, interacts with DNA mismatch repair protein MLH1. Proc Natl Acad Sci U S A. 1999;96(7):3969‐3974. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199. Petronzelli F, Riccio A, Markham GD, et al. Investigation of the substrate spectrum of the human mismatch‐specific DNA N‐glycosylase MED1 (MBD4): fundamental role of the catalytic domain. J Cell Physiol. 2000;185:473‐480. [DOI] [PubMed] [Google Scholar]
  • 200. Petronzelli F, Riccio A, Markham GD, et al. Biphasic kinetics of the human DNA repair protein MED1 (MBD4), a mismatch‐specific DNA N‐glycosylase. J Biol Chem. 2000;275:32422‐32429. [DOI] [PubMed] [Google Scholar]
  • 201. Tanakaya K, Kumamoto K, Tada Y, et al. A germline MBD4 mutation was identified in a patient with colorectal oligopolyposis and early onset cancer: a case report. Oncol Rep. 2019;42(3):1133‐1140. [DOI] [PubMed] [Google Scholar]
  • 202. Sanders MA, Chew E, Flensburg C, et al. MBD4 guards against methylation damage and germline deficiency predisposes to clonal hematopoiesis and early‐onset AML. Blood. 2018;132:1526‐1534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203. Rodrigues M, Mobuchon L, Houy A, et al. Outlier response to anti‐PD1 in uveal melanoma reveals germline MBD4 mutations in hypermutated tumors. Nat Commun. 2018;9(1):1866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204. Johansson PA, Stark A, Palmer JM, et al. Prolonged stable disease in a uveal melanoma patient with germline MBD4 nonsense mutation treated with pembrolizumab and ipilimumab. Immunogenetics. 2019;71(5–6):433‐436. [DOI] [PubMed] [Google Scholar]
  • 205. Rodrigues M, Mobuchon L, Houy A, et al. Evolutionary routes in metastatic uveal melanomas depend on MBD4 alterations. Clin Cancer Res. 2019;25:5513‐5524. [DOI] [PubMed] [Google Scholar]
  • 206. Derrien AC, Rodrigues M, Eeckhoutte A, et al. Germline MBD4 mutations and predisposition to uveal melanoma. J Natl Cancer Inst. 2020;djaa047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207. Stresemann C, Lyko F. Modes of action of the DNA methyltransferase inhibitors azacytidine and decitabine. Int J Cancer. 2008;123(1):8‐13. [DOI] [PubMed] [Google Scholar]
  • 208. Jabbour E, Issa JP, Garcia‐Manero G, Kantarjian H. Evolution of decitabine development: accomplishments, ongoing investigations, and future strategies. Cancer. 2008;112(11):2341‐2351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209. de Vos D, van Overveld W. Decitabine: a historical review of the development of an epigenetic drug. Ann Hematol. 2005;84(suppl 1):3‐8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210. Kantarjian HM, Thomas XG, Dmoszynska A, et al. Multicenter, randomized, open‐label, phase III trial of decitabine versus patient choice, with physician advice, of either supportive care or low‐dose cytarabine for the treatment of older patients with newly diagnosed acute myeloid leukemia. J Clin Oncol. 2012;30(21):2670‐2677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211. DiNardo CD, Pratz KW, Letai A, et al. Safety and preliminary efficacy of venetoclax with decitabine or azacitidine in elderly patients with previously untreated acute myeloid leukaemia: a non‐randomised, open‐label, phase 1b study. Lancet Oncol. 2018;19(2):216‐228. [DOI] [PubMed] [Google Scholar]
  • 212. DiNardo CD, Pratz K, Pullarkat V, et al. Venetoclax combined with decitabine or azacitidine in treatment‐naive, elderly patients with acute myeloid leukemia. Blood. 2018;133(1):7–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213. Mei M, Aldoss I, Marcucci G, Pullarkat V. Hypomethylating agents in combination with venetoclax for acute myeloid leukemia: update on clinical trial data and practical considerations for use. Am J Hematol. 2019;94(3):358‐362. [DOI] [PubMed] [Google Scholar]
  • 214. Otoukesh S, Salhotra A, Marcucci G, Forman SJ, Pullarkat V, Aldoss I. The feasibility of venetoclax and decitabine in therapy‐related acute myeloid leukemia with concurrent advanced non‐hematological malignancies. Leuk Res. 2019;84:106196. [DOI] [PubMed] [Google Scholar]
  • 215. Daver N, Boddu P, Garcia‐Manero G, et al. Hypomethylating agents in combination with immune checkpoint inhibitors in acute myeloid leukemia and myelodysplastic syndromes. Leukemia. 2018;32(5):1094‐1105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216. Daver N, Garcia‐Manero G, Basu S, et al. Efficacy, safety, and biomarkers of response to azacitidine and nivolumab in relapsed/refractory acute myeloid leukemia: a nonrandomized, open‐label, phase II study. Cancer Discov. 2019;9(3):370‐383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217. Stahl M, Goldberg AD. Immune checkpoint inhibitors in acute myeloid leukemia: novel combinations and therapeutic targets. Curr Oncol Rep. 2019;21(4):37. [DOI] [PubMed] [Google Scholar]
  • 218. Yu G, Wu Y, Wang W, et al. Low‐dose decitabine enhances the effect of PD‐1 blockade in colorectal cancer with microsatellite stability by re‐modulating the tumor microenvironment. Cell Mol Immunol. 2019;16(4):401‐409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219. Nie J, Wang C, Liu Y, et al. Addition of low‐dose decitabine to anti‐PD‐1 antibody camrelizumab in relapsed/refractory classical Hodgkin lymphoma. J Clin Oncol. 2019;37(17):1479‐1489. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing not applicable to this article as no datasets were generated or analyzed during the current study.


Articles from Advanced Genetics are provided here courtesy of Wiley

RESOURCES