Sie sind auf Seite 1von 20

Received: 14 August 2019 | Accepted: 31 January 2020

DOI: 10.1002/jcp.29642

REVIEW ARTICLE

Altering chromatin methylation patterns and the


transcriptional network involved in regulation of
hematopoietic stem cell fate

Mohammad Shokouhian1 | Marziye Bagheri2 | Behzad Poopak3 |


Rouzbeh Chegeni4 | Nader Davari2 | Najmaldin Saki2

1
Department of Hematology and Blood
Transfusion, School of Allied Medical Sciences, Abstract
Iran University of Medical Sciences,
Tehran, Iran Hematopoietic stem cells (HSCs) are quiescent cells with self‐renewal capacity and
2
Thalassemia and Hemoglobinopathy Research potential multilineage development. Various molecular regulatory mechanisms such
Center, Health Research Institute, Ahvaz
as epigenetic modifications and transcription factor (TF) networks play crucial roles
Jundishapur University of Medical Sciences,
Ahvaz, Iran in establishing a balance between self‐renewal and differentiation of HSCs. Histone/
3
Department of Hematology, Faculty of DNA methylations are important epigenetic modifications involved in transcriptional
Paramedical Sciences, Tehran Medical
Sciences, Islamic Azad University, Tehran, Iran regulation of specific lineage HSCs via controlling chromatin structure and
4
Michener Institute of Education at University accessibility of DNA. Also, TFs contribute to either facilitation or inhibition of gene
Health Network, Toronto, Canada
expression through binding to enhancer or promoter regions of DNA. As a result,
Correspondence epigenetic factors and TFs regulate the activation or repression of HSCs genes,
Najmaldin Saki, Research Center of
playing a central role in normal hematopoiesis. Given the importance of
Thalassemia and Hemoglobinopathy, Health
Research Institute, Ahvaz Jundishapur histone/DNA methylation and TFs in gene expression regulation, their aberrations,
University of Medical Sciences, Ahvaz, Iran.
including changes in HSCs‐related methylation of histone/DNA and TFs (e.g.,
Email: najmaldinsaki@gmail.com
CCAAT‐enhancer‐binding protein α, phosphatase and tensin homolog deleted on the
chromosome 10, Runt‐related transcription factor 1, signal transducers and
activators of transcription, and RAS family proteins) could disrupt HSCs fate.
Herewith, we summarize how dysregulations in the expression of genes related to
self‐renewal, proliferation, and differentiation of HSCs caused by changes in
epigenetic modifications and transcriptional networks lead to clonal expansion and
leukemic transformation.

KEYWORDS

epigenetic modifications, gene expression, hematopoietic stem cells, leukemia, transcriptional


regulation

1 | INTRODUCTION equilibrium between self‐renewal and differentiation (Mayani, 2016;


Z. Wang & Bunting, 2013).
Hematopoietic stem cells (HSCs) are characterized by their quies- HSCs fate is governed by a combination of cell‐intrinsic networks
cent state despite preserving self‐renewal and multipotency cap- and extrinsic stimuli (Mayani, 2016). Some intrinsic regulators, in-
abilities. HSCs undergo a hierarchical cascade of differentiation, cluding epigenetic modifiers, transcription factors (TFs), and signaling
which gives rise to different mature cells of myeloid and lymphoid pathways are effective in stem cell function. Indeed, these intrinsic
lineages (Dorritie, Redner, & Johnson, 2014). To ensure their lifelong regulators play a vital role in controlling the balance between self‐
maintenance in the bone marrow (BM), HSCs rely on a dynamic renewal and differentiation of HSCs (Mayani, 2016; Sharma &

J Cell Physiol. 2020;1–20. wileyonlinelibrary.com/journal/jcp © 2020 Wiley Periodicals, Inc. | 1


2 | SHOKOUHIAN ET AL.

Gurudutta, 2016). Modifications in components of chromatin (his- 2.1 | Histone methylation


tones and DNA) cause changes in the structure of nucleosomes,
and make important contributions to gene expression and have a Histone modification, a type of posttranslational modification, is
central role in the determination of HSCs' fate because a decrease in implicated in different biological processes such as cell fate de-
the expression of self‐renewal genes is correlated to the activation of termination via modulating genomic accessibility and gene expres-
genes promoting HSCs differentiation (Mayani, 2016; Ohlsson, sion caused by methylation of lysine or arginine residues. Histone
Schuster, Hasemann, & Porse, 2016). Interaction between chromatin methylation marks have a central role in epigenetic regulation
modifications (e.g., histone acetylation/deacetylation and histone/ through activating or repressing the transcription of genes involved
DNA methylation/demethylation) and TFs plays an essential role in in either differentiation or proliferation (Asada & Kitamura, 2018;
the modulation of gene expression in HSCs. As processes effective in Katoh, 2013; Madan et al., 2018; P. Zhang, Chen et al., 2018). Epi-
chromatin modifications, histone, and DNA methylations are capable genetic modifiers, which are effective in histone methylation, can
of making genomic DNA accessible for cellular processes such as help maintain the stability of specific genomic expression patterns in
transcription (Sharma & Gurudutta, 2016). On the other hand, once HSCs or terminally differentiated cells (Cui et al., 2009).
TFs bind to the enhancer or promoter regions of DNA, they facilitate A family of chromatin‐binding proteins named the additional sex
or inhibit gene expression. When TFs bind to a specific DNA combs‐like (ASXL) family plays a vital role in histone modifications.
sequence of target genes, RNA polymerase II is recruited to the The ASXL1 gene is located on chromosome 20q11 and encodes the
promoter and contributes to transcriptional regulation of HSCs and ASXL1 protein as a tumor suppressor, which is crucial for hemato-
lineage‐specific gene expression (Ohlsson et al., 2016; Safaei poiesis such as regulation of self‐renewal and differentiation of bone
et al., 2018). marrow stromal cells and hematopoietic stem and progenitor cells
Aberrant changes or mutations in epigenetic regulators can re- (HSPCs; Asada, Fujino, Goyama, & Kitamura, 2019; Oak &
sult in perturbation of normal hematopoiesis, aberrant clonal hema- Ohgami, 2017; P. Zhang, Chen et al., 2018). The ASXL genes (ASXL1,
topoietic expansion, and leukemic transformation via altering histone ASXL2, and ASXL3) are human homologs of Drosophila Asx recruiting
modifications and gene expression (Carbuccia et al., 2009; Chan & polycomb group (PcG) repressor complexes (PRC; PRC1 and PRC2)
Majeti, 2013; Kunimoto & Nakajima, 2017). It seems that dysregu- and trithorax group (TrxG) activator complexes to genes and mod-
lations in histone/DNA methylation and TFs activity in HSCs lead to ifying chromatin structures. PRC2 recruits PRC1 and catalyzes mono
aberrant gene expression, creating an imbalance between cell self‐ ‐ubiquitination of histone H2A at lysine 119 (H2AK119ub), resulting
renewal and lineage commitment decisions, leading to abnormal in the recruitment of specific genomic loci, trimethylation catalysis of
hematopoiesis and leukemogenesis (myeloid and lymphoid malig- lysine residue 27 (H3K27; H3K27me3) and consequent transcrip-
nancies). Hence, here we discuss how histone/DNA methylation and tional repression of PcG target genes (Figure 1; Katoh, 2013). TrxG
TFs modulate gene expression of HSCs in a way that results in the and PcG are critical proteins of the chromatin‐modification process
initiation or progression of leukemia. with opposing activities in the regulation of gene expression, so that
TrxG is associated with positive regulation of gene expression via
methylation of histone H3 on lysine residue 4 (H3K4), while PcG
2 | H ISTON E A ND DN A M E THY LA TION represses gene expression through methylation of H3K27 (Krivtsov
I N V O L V E D I N G E N E E X P RE S S I O N O F & Armstrong, 2007). ASXL1 acts as a mediator for the function of
NORMA L H EMATOP O IESIS PRC2 and a cofactor for BRCA‐1 associated protein 1 (BAP1), which
is the deubiquitinase and essential component of the polycomb re-
Epigenetic modifications are vital biological processes for chromatin pressive deubiquitinase complex (P. Zhang, Chen et al., 2018). ASXL1
remodeling, DNA accessibility and gene expression regulation directly binds to BAP1 and then removes the ubiquitination of
(Herviou, Cavalli, Cartron, Klein, & Moreaux, 2016). Histone and DNA H2AK119ub, which results in transcriptional activation (Asada &
methylations are two types of epigenetic modifications cooperating to Kitamura, 2018; Asada et al., 2018; Katoh, 2013). On the other hand,
control DNA functions of various cell types such as HSCs via modifying the BAP1/ASXL1 complex causes transcriptional repression by direct
chromatin and modulating transcription initiation, elongation, splicing, binding to scaffold subunits (EZH2, EED, and SUZ12) of PRC2 (Asada
and termination. These processes play a central role in regulating self‐ et al., 2018; Madan et al., 2018) and increasing H3K27me3, which
renewal, maintenance, and lineage differentiation of HSCs through finally represses further recruitment of PRC1 (Asada et al., 2019;
specific gene expression patterns (Herviou et al., 2016; Safaei Katoh, 2013). Histone‐lysine N‐methyltransferase EZH2 is an epige-
et al., 2018; Sharma & Gurudutta, 2016). According to Figure 1, pro- netic regulator for catalytic subunit of PRC2, which is effective in
cesses such as histone methylation at lysine 4 of histone H3 and DNA controlling the proliferation and differentiation potential of em-
demethylation are associated with gene transcriptional activation, bryonic stem cells (ESCs) by repressing the INK4B‐ARF‐INK4A locus
while methylation at lysine 27 of histone H3 and DNA methylation and preventing the recruitment of AP1 transcriptional activator from
leads to gene transcriptional silencing or stable long‐term repression. binding and activating late differentiation genes in ESCs (Ezhkova
These processes are dependent on one another and their cross‐talk is et al., 2009). Furthermore, the ASXL1/BAP1 complex has a role in
essential for controlling gene expression in HSCs. controlling the expression of INK4B‐ARF‐INK4A locus and is also
SHOKOUHIAN ET AL. | 3

F I G U R E 1 Mechanisms of epigenetic modifications and transcription factor (TF) networks in gene expression of HSCs. Epigenetic modifiers
and TFs as regulators of gene expression which play a central role in balancing HSCs fate in normal hematopoiesis. Interaction between
epigenetic modifications and TFs network regulates the activation and repression of genes in HSCs and their aberrant interaction could lead to
leukemogenesis. α‐KG, α‐ketoglutarate; 5mC, 5‐methylcytosine; 5hmC, 5‐hydroxymethylcytosine; ASXL1, additional sex combs‐like; BAP1,
BRCA‐1 associated protein 1; CpG, cytosine‐phosphoguanine; DNMT3A, DNA methyltransferase; H2AK119ub, histone H2A at lysine 119;
H3K27me3, trimethylation catalysis of H3K27; H3K4me3, trimethylated H3K4; SAM, s‐adenosylmethionine; HAT, histone acetyltransferase;
HDAC, histone deacetylases; HSCs, hematopoietic stem cells; IDH1/2, isocitrate dehydrogenase; MLL, mixed‐lineage leukemia; PRC1/2,
polycomb group repressor complexes 1/2; SAH, s‐adenosylhomocysteine; TFs, transcription factors; TET2, ten‐eleven translocation 2

effective in cellular defense against tumorigenesis. It was shown that A. Yokoyama, 2017). In fact, MLL is a key factor in controlling the
high expression of INK4B increases both the recruitment of ASXL1/ transcriptional network of HSCs‐specific genes involved in self‐
BAP1 complex to INK4B locus and deubiquitylation of H2AK119ub, renewal, proliferation and lineage‐specific differentiation (Artinger
resulting in cell growth arrest or differentiation (Wu et al., 2015). In et al., 2013). Previous studies have shown that histone modification
contrast, mutations in ASXL1 contribute to the expansion of hema- following EZH2 is essential in the cell‐cycle status of HSCs and cell
topoietic progenitors in the BM as these mutations decrease the differentiation by modulating gene expression profiles (Mochizuki‐
deubiquitylation of H2AK119ub and increase DNA methylation in Kashio et al., 2011). Given the high level of EZH2 expression in HSCs
the INK4B promoter followed by a decrease in the expression levels and its decrease following differentiation, it has been reported that
of p15INK4B (Wu et al., 2015). Interaction of the ASXL1/BAP1 EZH2 overexpression helps preserve the long‐term repopulation
complex with BAP1‐binding partners such as HCFC1 results in the potential of HSCs and prevents HSCs exhaustion under replicative
recruitment of SET1‐like or mixed‐lineage leukemia (MLL) complexes stress conditions (Kamminga et al., 2006). EZH2 has a cooperative
to specific genomic loci and catalysis of trimethylated H3K4 role in managing the balance between self‐renewal and differentia-
(H3K4me3), which contributes to transcriptional activation, which is tion of HSCs by inhibiting the expression of HSCs differentiation
shown in Figure 1 (Katoh, 2013). The MLL gene is a human homolog genes such as the inhibitor of differentiation 2 (ID2) and SOX7 as
of Drosophila melanogaster TrxG encoding a DNA‐binding protein that well as normal B‐cell differentiation (Herviou et al., 2016; Safaei
methylates H3K4 and plays an important role in the development et al., 2018). Moreover, EZH2 represses proapoptotic genes such as
and expression of the Hox gene (Krivtsov & Armstrong, 2007). NOXA, P21, and WIG1, thereby playing an essential role in protecting
Through upregulating the expression of posterior Hox genes in the HSCs from cell death (Herviou et al., 2016). EZH2 can increase the
hematopoietic cell lineage, MLL triggers the proliferation of pool of quiescent HSCs and prevent stem cell senescence via in-
immature hematopoietic progenitors and blocks hematopoietic hibiting apoptosis and proliferation of HSCs and also regulating
differentiation by maintaining the expression of self‐renewal‐related HSCs differentiation (Herviou et al., 2016). In fact,
genes at immature progenitor stages (Akihiko Yokoyama, 2015; EZH2‐mediated H3K27me3 leads to PRC1 binding to chromatin,
4 | SHOKOUHIAN ET AL.

mono‐ubiquitinylated H2AK119ub and eventual transcription self‐renewal genes (e.g., NR4A2) as well as the genes involved in
repression of target genes (Herviou et al., 2016; Safaei et al., 2018). HSCs differentiation (e.g., RUNX1 and GATA3), which contributes to
EZH2 mutation increases the activity of HSCs, which means a rise in the maintenance of self‐renewal and normal HSCs development
HSCs proliferation that could result in heterogeneous hematopoietic (Chan & Majeti, 2013; Trowbridge & Orkin, 2011). Two mechanisms
malignancies (Herviou et al., 2016; Mochizuki‐Kashio et al., 2015). drive differentiation to lineage commitment, one of which is the no
Research has provided evidence for the compensatory role of EZH1 silence gene in HSCs owing to the lack of methylation in DNMT3A‐
for EZH2 (Mochizuki‐Kashio et al., 2011), so that EZH1 may be in- null HSCs and the other is elevated H3K4me3 due to receiving
fluential in the maintenance of hematopoiesis in the absence of EZH2 differentiation signals (Challen et al., 2011). Therefore, DNA me-
(Mochizuki‐Kashio et al., 2015). EZH1 is similar to EZH2 in main- thylation is an essential director of normal HSCs function.
taining primitive hematopoietic cells and balancing HSCs preserva- The ten‐eleven translocation (TET) family (TET1, TET2, and
tion against senescence and physiological HSCs activation so that a TET3) of proteins regulates DNA demethylation, making them pivotal
reduction in its expression is accompanied by the impairment of in the normal development of HSCs. Although TET2 and TET3 show
HSCs self‐renewal capacity (Hidalgo & Gonzalez, 2013; Hidalgo high expression in hematopoietic cells, TET1 has a higher level of
et al., 2012). Several studies suggest a connection between changes expression in ESCs and serves as a transcriptional repressor in these
in messenger RNA (mRNA) expression of PcG family genes with the cells (Pronier & Delhommeau, 2012). TET1 indirectly mediates the
age of HSPCs; for instance, the expression of EZH1 and Bmi1 is binding of EZH2 to gene promoters, followed by the recruitment of a
reduced in older HSPCs relative to younger ones, which implies that repressive histone mark (H3K27me3) resulting in the repression of
the PcG family delays HSPCs senescence (Y. Dong et al., 2018). gene expression and ESCs maintenance (Sui et al., 2012). However,
According to the above, it can be deduced that PcG and TrxG family TET1 reduction resulting from PRC2 complex binding to TET1 pro-
genes and their related chromatin‐binding proteins have indis- motor contributes to cell proliferation (Neri et al., 2015). TET2 is a
pensable roles in maintaining the mechanism of HSCs cellular tumor suppressor, whose gene is located on chromosome 4q24 and is
memory and preventing the alteration of HSCs fate determination essential for DNA hypermethylation prevention and demethylation
through chromatin modifications via promoting or repressing the promotion (L. Cimmino, Abdel‐Wahab, Levine, & Aifantis, 2011;
expression of target genes. Therefore, defects in these processes Mercher et al., 2012). DNA demethylation involves the conversion of
contribute to the senescence, aberrant proliferation, or differentia- 5mC to 5‐hydroxymethylcytosine (5hmC), 5‐formylcytosine (5fC) and
tion of HSCs (defect in hematopoiesis), and development of 5‐carboxylcytosine (5caC), in which Fe (II) and a‐ketoglutarate (a‐KG)
leukemogenesis. binding to both the double‐stranded β‐helix fold domain located at
C‐terminus of TET2 protein and oxygen is helpful (L. Cimmino
et al., 2011; Feng, Li, Cassady, Zou, & Zhang, 2019; Mercher
2.2 | DNA methylation et al., 2012). Indeed, methylated DNA recognition by TET2 in the
CpG island and 5mC insertion into the catalytic cavity facilitate their
DNA methylation/demethylation is another type of epigenetic mod- reaction with catalytic Fe (II), leading to 5mC oxidation (Hu
ification, which is crucial in cellular processes, such as regulation of et al., 2013). Then, 5hmC becomes deaminated by the activation‐
gene expression, genomic stability, and cell lineage commitment. induced cytosine deaminase/apolipoprotein B mRNA‐editing enzyme
Dysregulation of these modifications could result in aberrant stem complex protein family, producing 5hmU, after which unmethylated
cell function and tumorigenesis (Brunetti, Gundry, & Goodell, 2017; cytosine is replaced with thymine DNA glycosylase (TDG) via the
Nakajima & Kunimoto, 2014; Sui, Price, Li, & Chen, 2012). In fact, base exchange repair (BER) mechanism (L. Cimmino et al., 2011;
changes in genomic DNA methylation induce distinctions in different Mercher et al., 2012; Nakajima & Kunimoto, 2014). Similar to 5hmC,
stages of hematopoietic differentiation as there is a descending 5fC and 5caC are detected by TDG and unmethylated cytosine is
pattern in promoter methylation throughout HSPCs’ differentiation replaced with BER (Nakajima & Kunimoto, 2014). On the other hand,
and aging (Bocker et al., 2011). It was demonstrated that un- TET2 can have an impact on the regulation of DNA methylation
methylated HSCs‐specific genes induce normal hematopoiesis, sup- through passive DNA demethylation, so that this process is per-
porting the role of DNA methylation in silencing the genes affecting formed by replacing 5hmC with unmethylated cytosine (Bowman &
HSCs differentiation (Sharma & Gurudutta, 2016). DNA methyl- Levine, 2017; L. Cimmino et al., 2011). Moreover, TET protein in-
transferase (DNMT) family (DNMT1 and DNMT3A/B) members teraction with some proteins involved in epigenetic mechanisms, in-
cooperate in transcriptional silencing of downstream genes by adding cluding O‐linked β‐D‐N‐acetylglucosamine (O‐GlcNAc) transferase
a methyl group of S‐adenosyl‐methionine, to carbon‐5 (C5) of (OGT), may be important in gene transcription regulation (Q. Chen,
cytosine‐phosphoguanine (CpG) dinucleotides in DNA and generating Chen, Bian, Fujiki, & Yu, 2013; Nakajima & Kunimoto, 2014) as it was
5‐methylcytosine (5mC; Figure 1; Kunimoto & Nakajima, 2017). revealed that TET–OGT interaction contributes to the increase in
DNMT1 intervenes in methylation of the newly synthesized CpG binding of SET1/COMPASS H3K4 methyltransferase to chromatin
island during DNA replication, whereas DNMT3A and DNMT3B in- and subsequent promotion of transcriptional activation via increasing
tervene in de novo methylation activity (Brunetti et al., 2017). DNA histone H3K4me3 (Deplus et al., 2013; Feng et al., 2019; Nakajima &
methylation mediated by DNMT3A represses the expression of Kunimoto, 2014). Prior studies suggest that the catalytic activity of
SHOKOUHIAN ET AL. | 5

TET2 is related to the proteins encoded by other genes such as SCL/RUNX1 transcription (C. Li et al., 2015). Given the involvement
metabolic enzymes isocitrate dehydrogenase (IDH) 1 and 2 (Ko & of TET2, DNMT3A, IDH1, and IDH2 in the regulation of gene tran-
Rao, 2011; Pronier & Delhommeau, 2012). IDH1/IDH2 enzymes scription during the development, cellular fate, and function of HSCs,
convert isocitrate to a‐KG through oxidative decarboxylation it appears that dysregulation of these epigenetic modifiers results in
(L. Cimmino et al., 2011; Nakajima & Kunimoto, 2014). As shown in the generation of 5hmC and failure of hematopoiesis.
Figure 1, α‐KG is required for the functioning of the TET family and
some other enzymes such as histone demethylases (JmjC family) and
DNA methyltransferases (ALKBH2/3 and MGMT). It seems that 3 | TRANSCRIPT ION FACTOR N ETWO RKS
α‐KG‐dependent reactions are effective in epigenetics and DNA re- AND GEN E EXP RE SSION I N N ORMAL
pair of HSCs (Inoue, Lemonnier, & Mak, 2016). IDH1/IDH2 mutant HEMAT O POIESIS
proteins generate an oncometabolite (i.e., 2‐hydroxyglutarate) that
competes with a‐KG because a‐KG is capable of inhibiting the cata- As primary components of transcriptional regulatory networks, TFs
lytic activity of TET2 (L. Cimmino et al., 2011; Holmfeldt & Mulligh- play an important role in the regulation of gene expression by binding
an, 2011; Mercher et al., 2012). The IDH1 mutant promotes clonal to specific DNA sequences. Furthermore, the interaction between
expansion by enhancing self‐renewal in HSCs and expanding lineage‐ transcriptional and epigenetic modifiers is responsible for the control
specific Sca1+ c‐Kit+ (LSK) cells (Chan & Majeti, 2013). Furthermore, of gene expression (Gottgens, 2015). Notably, transcriptional
it has been shown that posttranslational modification (phosphoryla- mechanisms play key roles in HSCs fate such as self‐renewal, com-
tion) of the TET2 enzyme is effective in epigenetic changes during mitment, and differentiation of specific cell types (Avellino &
hematopoiesis. In this context, J. J. Jeong et al. (2019) reported that Delwel, 2017). In fact, TFs cooperate in managing the balance be-
Janus kinase 2 (JAK2) results in the interaction of phosphorylated tween self‐renewal and differentiation of HSCs (Suzuki et al., 2017),
TET2 with KLF1 (erythroid transcription factor) by phosphorylating so that lineage‐specific TFs (such as CCAAT‐enhancer‐binding pro-
tyrosine's 1939 and 1964 in the catalytic core of TET2. In addition, tein α [C/EBPα], phosphatase and tensin homolog deleted on the
increased JAK2 activation caused by erythropoietin (EPO) has a chromosome 10 [PTEN], Runt‐related transcription factor 1
positive feedback effect on the interaction between TET2 and KLF1, (RUNX1), and signal transducers and activators of transcription
increasing TET2 activation followed by elevated levels of 5hmC in [STAT] family proteins) induce HSCs differentiation through reducing
the genome and resulting in an increase of erythroid differentiation HSCs self‐renewal and activating lineage‐specific genes (Koschmie-
(J. J. Jeong et al., 2019). However, the increase of 5hmC in specific der, Halmos, Levantini, & Tenen, 2009; Ohlsson et al., 2016). Given
promoter regions of HSPCs caused by TET2 upregulation is accom- the role of TFs in gene expression of HSCs, it is not surprising that
panied with normal hematopoiesis, the decrease in transcriptional aberrant expression of HSCs‐related TFs is involved in the initiation
repression of these cells following a reduction in TET2 and 5hmC is of leukemia transformation by shifting HSCs toward self‐renewal or
associated with differentiation and increased immature LSK cells differentiation (Suzuki et al., 2017).
(L. Cimmino et al., 2011; Feng et al., 2019; Ko & Rao, 2011; Pan,
Weeks, Yang, & Xu, 2015). Activating histone marks such as H3K9
acetylation, H3K4me1, and H3K4me3 and the increase in 5hmC on 3.1 | CCAAT‐enhancer‐binding protein α
genomic loci of CD34+ stem/early progenitor cells commitment
during erythropoiesis following opening of condensed chromatin is CCAAT‐enhancer‐binding protein α (C/EBPα) is a member of the
accompanied with binding of specific erythroid TFs (e.g., GATA1, family of basic region leucine zipper transcription factors (C/EBP‐α,
GATA2, and KLF1) to these locations and erythroid differentiation ‐β, ‐γ, ‐δ, ‐ε, and ‐ζ) promoting gene expression via binding to certain
(Madzo et al., 2014). Epigenetic modifiers DNMT3A and TET2 create genes (Avellino & Delwel, 2017; Ohlsson et al., 2016). All members of
5mC and 5hmC, respectively, subsequently modifying cytosine at the CEBP have C‐terminal and N‐terminal domains that are essential to
CpG island that could help activate and repress HSCs genes; there- protein interactions (DNA binding and dimerization) and transcrip-
fore, the reduction in 5hmC level of HSCs lacking TET2 and DNMT3A tion control (Avellino & Delwel, 2017). CEBPA is an intron‐less gene
results in differentiation inhibition and self‐renewal promotion of located on chromosome 19q13.1 that encodes two isoforms (42‐kD
HSCs through increasing expression of EPOR and KLF1 genes and 30‐kD) of a DNA‐binding protein (Avellino & Delwel, 2017;
(X. Zhang et al., 2016). Thus, the balance between 5mC and 5hmC Ohlsson et al., 2016). C/EBPα is a key factor in differentiation pro-
has an essential role in gene transcription regulation and cell fate cesses, such as liver development, adipogenesis, and hematopoiesis
decisions. Increase of 5hmC, 5caC, and 5fC following overexpression (Ohlsson et al., 2016). There is a high expression of C/EBPα in
of TET2 can result in the cell‐cycle defect (G1‐S and/or intra‐S myeloid lineage and although its expression is low in HSCs, C/EBPα
transition) by activating P53 and inducing genetic instability and has a key role in controlling the expression of genes that keep HSCs
mutagenesis in the absence of TDG, which increases C > T muta- quiescent and maintain long‐term HSCs (LT‐HSCs; Avellino &
genesis in the CpG island (Mahfoudhi et al., 2016). Also, the increase Delwel, 2017; Koschmieder et al., 2009; Quintana‐Bustamante
of 5hmC following TET2/3 overlapping plays a role in normal HSCs et al., 2012). This transcription factor regulates differentiation, pro-
development by signaling and downstream expression of GATA2B/ liferation and cell death of HSCs through preventing their apoptosis
6 | SHOKOUHIAN ET AL.

and maintaining them in a quiescent state (Hasemann et al., 2014; of HSCs in BM via inhibiting both the PI3K/AKT signaling pathway
Ohlsson et al., 2016). Similarly, PU.1, which is an important tran- and cell contact‐related SRC/FAK/P38 signaling, whereas the former
scription factor for early myeloid and B‐cells differentiation, has low is related with the promotion of proliferation and differentiation of
expression in HSCs but acts as a regulator for HSC function. It has HSCs through inhibiting the PI3K/AKT signaling pathway and
been reported that overexpression of C/EBPα and PU.1 can impair enhancing cell contact‐related FAK/P38 signaling activity (J. Li
homeostasis of HSCs by reducing self‐renewal capacity, suppressing et al., 2016). As a result, phosphorylation of PTEN could have an
proliferation, decreasing mixed colony formation, increasing apop- important role in deciding hematopoietic cell fate. It was shown that
tosis, and facilitating differentiation (Fukuchi, Ito, Shibata, Kitamura, PTEN deletion reduced self‐renewal of HSCs, enhancing the level of
& Nakajima, 2008). On the other hand, C/EBPα results in granulocyte HSCs activation and generation of leukemia‐initiating cells, sup-
differentiation via interacting with the DNA‐binding domain of PU.1 porting the role of PTEN in the maintenance of quiescence, re-
and blocking PU.1 function (Reddy et al., 2002). C/EBPα also controls stricting the activation of HSCs and preventing leukemogenesis
the survival of HSPCs and maintains their repopulating ability, cell‐ (Yilmaz et al., 2006; J. Zhang et al., 2006). Therefore, PTEN is con-
cycle progression, and granulocytic differentiation by affecting the sidered invaluable in regulating the balance between proliferation
downstream target of the ecotropic viral integration site 2B and differentiation of blood cells and its disruption increases HSCs
(Zjablovskaja et al., 2017). It was demonstrated that the inhibition of cycling that might give rise to leukemia.
N‐Myc as a direct target of C/EBPα keeps HSCs in a quiescent state,
while its elevated expression in the complete absence of C/EBPα
results in an increase of proliferation and number of LT‐HSCs in BM 3.3 | Runt‐related transcription factor 1
(Ye et al., 2013). C/EBPα converts megakaryocyte/erythroid pro-
genitors and common lymphoid progenitors (CLPs) into the myeloid Runt‐related transcription factor 1 (RUNX1), known as acute mye-
lineage and induces myeloid differentiation of HSCs (Fukuchi logenous leukemia‐1 (AML1), is a member of the core‐binding factor
et al., 2006). Indeed, the expression of myeloid genes is activated by (CBF) family of TFs and a key factor in normal development (D. Hong
C/EBPα binding to promoters or enhancers of myeloid‐related genes et al., 2019). It is helpful in the generation of HSCs from the
(Avellino & Delwel, 2017). Furthermore, this transcription factor endothelium of great vessels including the vitelline and umbilical
inhibits the E2A/SCL heterodimer (essential to erythropoiesis) by arteries as well as from the aorta–gonad–mesonephros region
inducing the expression of ID1 and suppressing the erythropoietin (Kumano & Kurokawa, 2010; Kurokawa, 2006). RUNX1 expression is
receptor (EPOR), resulting in the inhibition of erythroid genes and high in myeloid and lymphoid lineage cells, while it has a low ex-
induction of myeloid genes (Suh et al., 2006). In addition to the role of pression in erythroid lineage (North, Stacy, Matheny, Speck, & de
C/EBPα in the initiation of myelopoiesis pathway, its activity is vital Bruijn, 2004), so that RUNX1 has a central role in the regulation of
for the functional evolution of myeloid cells such as highly active HSPCs biology and the establishment of a balance between pro-
migratory behavior (Y. Chen et al., 2009). According to the fact that liferation and differentiation of myeloid and lymphoid lineages
the number and function of HSCs and myeloid differentiation are (D. Hong et al., 2019; North et al., 2004). RUNX1 produces three
affected by C/EBPα, we can infer that the deregulation of C/EBPα alternatively spliced variants, namely RUNX1a, RUNX1b, and
expression results in the enhancement of predisposing preleukemic RUNX1c. Researchers have provided evidence for the role of RUN-
conditions, especially myeloid leukemogenesis. X1a in enhancing HSCs self‐renewal and driving the proliferation of
primitive progenitors that enable the expansion of immature pro-
genitors' population (Tsuzuki et al., 2007; Tsuzuki & Seto, 2012).
3.2 | Phosphatase and tensin homolog deleted on However, RUNX1b promotes granulocytic differentiation of HSCs
the chromosome 10 (Tsuzuki et al., 2007). RUNX1 forms a heterodimer complex by
binding to core‐binding factor subunit beta (CBF‐β) and interacting
The phosphatase and tensin homolog deleted on the chromosome 10 with the DNA sequence at promoters of target genes, which serve as
(PTEN) gene is located on chromosome 10q23 and encodes the PTEN either a transcription activator or repressor of target genes involved
protein that acts as a tumor suppressor and phosphatase. PTEN has a in hematopoietic differentiation, cell‐cycle regulation, ribosome bio-
vital role in multiple cellular mechanisms such as proliferation, sur- genesis, P53, and transforming growth factor β (TGFβ) pathways
vival, growth, migration, and differentiation by dephosphorylating (D. Hong et al., 2019; Sood, Kamikubo, & Liu, 2017). Actually, the
phosphatidylinositol 3, 4, 5‐trisphosphate (PIP3) and inactivating the RUNX1/CBF‐β complex represses the expression of a target gene
phosphatidylinositol 3 kinase (PI3K)/AKT signaling pathway (Fragoso through binding to the consensus DNA sequence (the PEBP2 site)
& Barata, 2014; Morotti et al., 2015). Previous studies have em- and interacting with corepressor mSin3A, which as a regulatory
phasized the key role of PTEN in the regulation of quiescence, cell‐ protein plays a role in stabilization of transcriptional repressor
cycle entry, proliferation, and differentiation of HSCs (Fragoso & complex by protecting RUNX1 from proteolytic degradation and
Barata, 2014; Morotti et al., 2015). PTEN has two forms: phos- subsequent formation of the transcriptional complex. Phosphoryla-
phorylated (p‐PTEN) and non‐phosphorylated (non‐p‐PTEN). The tion of RUNX1 on Ser249 and Ser266 by an extracellular signal‐
latter is effective in maintaining the quiescent/undifferentiated state regulated kinase (ERK) in response to growth factor stimuli
SHOKOUHIAN ET AL. | 7

contributes to the release of RUNX1 form mSin3A, leading to re- et al., 2007). It was also seen that decreasing expression of
cruitment of transcriptional coactivators P300 and CREB‐binding quiescence‐associated genes such as P57 and Tie2 increases cell‐
protein to RUNX1 and conversion of the RUNX1 complex into cycle and apoptosis, which brings about the loss of HSCs quiescence
an activator of gene transcription (Kurokawa, 2006). In addition, the and reduction of LT‐ HSC numbers (Z. Wang, Li, Tse, & Bunting, 2009;
interaction between RUNX1/CBF‐β complex and core components of Z. Wang et al., 2015). In addition, mutation of STAT5 in primitive LSK
PRC1 such as Ring1b and Bmi1 recruits PRC1 at site‐specific loci cells promotes self‐renewal of HSCs, expansion of multipotential
of target genes in megakaryocytic and lymphocytic cells (Yu progenitors and initiation or progress of leukemia such as myelo-
et al., 2012). Furthermore, the increase in the recruitment of TET and proliferative disorder (MPD; Dorritie et al., 2014; Kato et al., 2005). It
other demethylation‐related enzymes to the CpG island induces DNA was reported that down‐modulation of C/EBPα in CD34+ cells fol-
demethylation by RUNX1, which can be effective in unlocking si- lowing activated mutant STAT5A1*6 causes an enhancement of
lenced hematopoietic genes and hematopoietic development self‐renewal, expansion of LT‐HSCs, blocked myeloid differentiation,
(Suzuki et al., 2017). RUNX1 negatively regulates the number of and a bias toward erythropoiesis (Wierenga, Schepers, Moore,
quiescent HSCs, as there seems to be an inverse correlation between Vellenga, & Schuringa, 2006). Indeed, the activity of STAT5 on HSCs
LT‐HSCs activity and RUNX1 dosage. Overall, the disruption of self‐renewal and differentiation is dosage‐dependent as low to in-
RUNX1 function results in the expansion of immature hematopoietic termediate STAT5 activity is associated with optimal self‐renewal
cells, which develop into a variety of hematological malignancies and LT‐HSCs expansion, impaired myelopoiesis, and induction of
(Ichikawa et al., 2008). erythropoiesis (Wierenga, Vellenga, & Schuringa, 2008). In general, it
seems that given the association between the activity rate of STAT3
and STAT5 proteins with the regulation of stem cell fate and lineage
3.4 | Signal transducers and activators of commitment, changes in the activity of these STATs may result in
transcriptions hematologic malignancies.

The signal transducers and activators of transcriptions (STATs) family


is composed of seven members of cytoplasmic TFs with a crucial role 4 | DYSREGULATION OF HEMATOPOIETIC
in the regulation of proliferation, apoptosis suppression, differentia- STE M CE LLS ‐I N DU C E D LE U K E M I C
tion, and survival of hematopoietic cells (Dorritie et al., 2014). Acti- TRANSFORMATION BY IMPAIRED
vation of STAT3, by granulocyte‐macrophage‐colony‐stimulating EXPRESSION OF GE NES
factor (GM‐CSF) in primary CD34+ BM cells, results in its binding to
DNA sequence of the apoptosis inhibitor (survivin) core promoter Chromatin accessibility and gene expression patterns are controlled
and contributes to the inhibition of apoptosis and stimulation of by epigenetic modifications and transcriptional networks, which have
CD34+ cell proliferation by increasing the level of survivin important effects on the regulation of HSCs self‐renewal and their
(Gu, Chiang, Zhu, Findley, & Zhou, 2007). An overlap between the differentiation into mature cells of myeloid and lymphoid lineages
activation of STAT3 and the upregulation of HoxB4 induces gene‐ (Gottgens, 2015; Sharma & Gurudutta, 2016). Any aberration derived
specific TFs that are responsible for self‐renewal of HSCs such as from changes in epigenetic regulators and TFs in HSCs gene ex-
Oct‐4 and Nanog, resulting in the promotion of self‐renewal and pression results in impaired HSCs fate such as self‐renewal and
repopulating activity of HSCs (S. H. Hong et al., 2014). Notably, lineage choice, as well as an acceleration of leukemic transformation.
activation of STAT5 by cytokines including stem cell factor, throm-
bopoietin (TPO), granulocyte‐colony‐stimulating factor (G‐CSF),
GM‐CSF, interleukin‐3 (IL‐3), and EPO cooperates in keeping HSCs 4.1 | Additional sex combs‐like 1
quiescent, maintaining their self‐renewal capacity, and promoting
CD34+ cell growth (Dorritie et al., 2014; Z. Wang & Bunting, 2013). Wild‐type (WT) ASXL1 in cooperation with PRC2 inhibits target
Reduction of STAT5 results in the arrest of B‐cells development in genes such as the posterior HoxA gene. In other words, loss of
the pro‐ to pre‐B‐cell stage, supporting its role in the regulation of functional features of this tumor suppressor in ASXL1 truncation
lymphoid and myeloid development (Dorritie et al., 2014; Z. Wang & mutation (MT) reduce the inhibition of target genes, which can be
Bunting, 2013). A study by Z. Wang, Medrzycki, Bunting, and Bunting followed by progression and initiation of myeloid malignancies such
(2015) concluded that deficiency of STAT5 causes the upregulation as myelodysplastic syndromes (MDS), acute myeloid leukemia (AML),
of proto‐oncogene Myc and apoptosis inhibitor BCL‐2, which con- chronic myelomonocytic leukemia (CMML), and MPD (Asada &
tribute to the expansion of preleukemic lymphoid and induce trans- Kitamura, 2018; Asada et al., 2019; Oak & Ohgami, 2017). Alteration
formation to acute B lymphoblastic leukemia (B‐ALL). It was shown in ASXL1 might induce biased differentiation toward granulocytic/
that the downregulation of this transcription factor could effectively monocytic lineages, supporting the essential role of ASXL1 in main-
reduce the number of HSCs and the common myeloid progenitor taining HSPCs' functions and differentiation bias (P. Zhang, Chen
(CMP), decreasing their expansion due to the function of STAT5 in et al., 2018). Increased catalytic function and nuclear localization of
the maintenance and expansion of primitive HSCs (Schepers BAP1 in ASXL1‐MT results in mono‐ubiquitination at lysine 351 of
8 | SHOKOUHIAN ET AL.

mutant ASXL1. Monoubiquitinated ASXL1‐MT, in turn, increases the 1, C‐X‐C motif ligand 1 (CXCL1), and CXCL2 (P. Zhang, Chen
catalytic function of BAP1 and establishes the hyperactive et al., 2018). In addition, similar to ASXL1, OGT methylates H3K4/
ASXL1‐MT/BAP1 complex, leading to the decrease in the level of H3K27 through a direct relationship with HCFC1 and MLL5, thereby
H2AK119ub in HSCs, aberrant differentiation of these cells and exerting an influence on transcription activation and regulation of
myeloid progenitor cells (Asada & Kitamura, 2018; Asada et al., 2018; myeloid differentiation. This enzyme directly stabilizes ASXL1
Nagase et al., 2018). Furthermore, H2AK119Ub at gene promoters through O‐GlcNAcylation of ASXL1‐S199. Notably, knockdown of
has a repressive effect on gene expression; consequently, its reduc- ASXL1, OGT, or MLL5 as suppressors of myeloid malignancies re-
tion following the effect of the ASXL1‐MT/BAP1 complex disturbs duces H3K4 methylation and blocks myeloid differentiation, which
gene repression (Balasubramani et al., 2015). Besides this, reduction leads to myeloid malignancies (Inoue et al., 2018). Furthermore,
of H2AK119Ub mediated by the hyperactive ASXL1‐MT/BAP1 ASXL1‐MT in human CD34+ progenitor cells decreases the genera-
complex could finally reduce the level of H3K27me3 tion of CD11b+ and CD15+ cells (markers of granulocytes and/or
(Balasubramani et al., 2015) and lead to the promotion of myeloid monocytes or macrophages), subsequently causing a decline in the
transformation via hindering PRC2 recruitment to specific oncogenic quantity of granulocyte, monocyte‐colony‐forming unit (GM‐CFU)
target loci (Abdel‐Wahab et al., 2012; Asada et al., 2019). On the and granulocyte‐colony‐forming unit (G‐CFU) colonies. Eventually,
other hand, through direct binding to the promoter of the HoxA gene this mutation both disrupts human granulomonocytic differentiation
as well as IRF8 and H2AK119ub removal, the ASXL1‐MT/BAP1 and increases the number of multipotent mixed‐lineage colony‐
complex contributes to the upregulation of posterior HoxA gene and forming units (GEMM‐CFU; Davies et al., 2013). A recent study by
IRF8. Owing to the role of the HoxA gene and IRF8 in leukemic H. Yang et al. (2018) on an ASXL1Y588XTg mouse model concluded
transformation and monopoiesis, respectively, their dysregulation is that the increase in transgenic expression of a truncated form of
effective in leukemogenesis (Abdel‐Wahab et al., 2012; Asada & ASXL1 in the hematopoietic compartment of mice can be effective in
Kitamura, 2018; Asada et al., 2018). Also, the upregulation of pos- myeloid differentiation and facilitation of leukemogenesis because
terior HoxA genes like HoxA9 inhibits PRC2‐mediated methylation this mutation triggers the increase of LSK and granulocyte‐
of histone H3K27 in ASXL1‐MT, increasing the expression levels of macrophage progenitor cells (GMPs) populations with skewed dif-
microRNA‐125a (miR‐125a) and subsequently reducing the expres- ferentiation favoring the monocytic/granulocytic lineage in BM. It
sion levels of C‐type lectin domain family 5‐member a (Clec5a). was also shown that increased percentage of blast cells in peripheral
Clec5a has a role in myeloid differentiation, and its reduction influ- blood (PB), BM, and spleen as well as decreased lifespan in this
ences the pathogenesis of myeloid malignancies such as MDS de- mouse might reflect the development of myeloid malignancies, in-
velopment (Inoue et al., 2013). The balance between ASXL1 and cluding MDS, MPD, and AML (H. Yang et al., 2018), which supports
BAP1 activity is vital in normal hematopoiesis as increased BAP1 the role of ASXL1‐MT in pathogenesis of myeloid malignancies. A
deubiquitinase activity in the ASXL1aa1‐587 truncation mutant in number of authors have recognized that dysregulation of regulatory
ASXL1Y588XTg mice promotes myeloid malignancies (Guo factors for erythropoiesis because of a decline in histone H3K27me3
et al., 2018). Recently, a study by Uni et al. (2019) on ASXL1G643fs and H3K4me3 during knockdown of ASXL1 in cord blood CD34+
knockout in mice revealed that the level of H2AK119ub at the pro- cells increases apoptosis, induces dysplastic features in the erythroid
moter region of p16Ink4a is reduced in HSCs ASXL1‐MT. p16Ink4a is lineage, and impairs erythroid commitment and terminal differ-
a major cyclin‐dependent kinase inhibitor, whose function is regu- entiation (Shi et al., 2016). In fact, knockdown of ASXL1 reduces
lated by PRC1 and has a pivotal role in preventing cell‐cycle pro- H3K27me3 on the HoxA and P21 locus in erythroid progenitors,
gression by decelerating G1‐S transition and maintaining cellular leading to the apoptosis of erythroid progenitor cells and aberrant
senescence (Serrano, Hannon, & Beach, 1993). As a result, p16Ink4a differentiation of the erythroid lineage (Hilgendorf, Folkerts, Schur-
upregulation, which results from a reduction in PRC1 recruitment to inga, & Vellenga, 2016). ASXL1‐MT in HSCs is mainly responsible for
the promoter region of p16Ink4a in ASXL1‐MT, can accelerate cel- acquiring additional mutations such as in RUNX1. Consequently, the
lular senescence by decreasing the HSCs pool, increasing apoptosis, simultaneous presence of these mutations could contribute to the
cell‐cycle arrest in HSCs fraction, and causing aberrant myeloid dif- susceptibility of HSCs to leukemic transformation, increased
ferentiation (Uni et al., 2019). In another study, it was reported that LT‐HSCs repopulation, reduced apoptosis, and blockage of erythroid
decreased levels of H3K27me3 and H3K4me3 in ASXL1+/− and differentiation (Nagase et al., 2018). In addition, coexistence of
ASXL1−/− Lin– c‐Kit+ cells, upregulation of apoptotic facilitator ASXL1+/– and NF1+/– (a negative regulator of the RAS signaling
(BCL2l13), and downregulation of antiapoptotic genes (BCL‐2 and pathway) in HSPCs increases the activity of multiple oncogenic
BCL2l12) would decrease HSCs repopulating capacity along with pathways such as Myc, NRAS, and BRD4, causing aggressive pro-
skewed cell differentiation favoring granulocytic lineage and redu- gression of myeloid malignancies (P. Zhang, He et al., 2018). Prior
cing HSCs pool (Abdel‐Wahab et al., 2013; J. Wang et al., 2014). studies have suggested that the reduction in acetylation of several
Furthermore, by reducing H3K27me3 in the absence of ASXL1, de- lysine residues on histone H3 and H4 at promoter regions of multiple
regulation of the transcriptional function of the RNA polymerase II TGFβ pathway genes, which is mediated by ASXL1‐MT and SET
complex contributes to the downregulation of genes that are critical binding protein 1 (SETBP1) mutations, results in the repression of
for HSPCs maintenance, including vascular cell adhesion molecule multiple TGFβ pathway genes (Saika et al., 2018). In a study, it was
SHOKOUHIAN ET AL. | 9

revealed that as the dysregulation mechanism of the TGFβ pathway TET2 plays a role in both the early and late stages of myeloid dif-
promotes leukemogenesis (M. Dong & Blobe, 2006), deregulation of ferentiation so that promoter hydroxymethylation along with TET2 is
TGFβ signaling in ASXL1‐MT and SETBP1 mutations is effective necessary for gene activation during C/EBPα‐induced trans‐
for the induction of myeloid neoplasms. This study also addresses the differentiation. C/EBPα binds to upstream regions of TET2 and
fact that histone deacetylase inhibitors increase histone acetylation activates TET2, which in turn binds to methylated promoters of
at promoter regions of TGFβ pathway genes and inhibit leukemic myeloid genes leading to the hydroxylation of these genes and trans‐
proliferation of ASXL1‐MT/SETBP1‐mutated cells, which may be differentiation of pre‐B cells into macrophages as TET2 knockdown
useful as therapeutic drugs for MDS and AML cases affected by these inhibits the generation of macrophages (Kallin et al., 2012). Homo-
mutations (Saika et al., 2018). In addition, the recognition of ASXL1 zygous TET2 mutations have a role in increasing self‐renewal and
binding partners facilitates the perception of tumor‐suppressive capacity of HSCs, expansion of LSK and CMP in BM and the fetal
functions of ASXL1 and its role in myeloid malignancies. A more liver (FL; Nakajima & Kunimoto, 2014). TET2 increases the conver-
comprehensive description can be found in an in‐vitro study by sion of 5mC to 5hmC at target HoxA cluster regions, which increases
Z. Li et al. (2017), in which it was demonstrated that ASXL1 inter- the activation of HoxA transcription and plays an important role in
action with the cohesin complex, including RAD21, SMC1A, SMC3, the maintenance of genome stability (M. T. Bocker et al., 2012). In
and STAG1/STAG2, plays a vital role in maintaining normal sister contrast, increased phosphorylation of histone H2AX (as a DNA da-
chromatid separation during cell division and regulation of gene ex- mage marker), increased DNA methylation and decreased HoxA
pression in hematopoietic cells. Furthermore, loss of ASXL1 de- transcription in the loss of TET2/3 function cause genome instability,
creases occupancy of RAD21 and SMC1A on target genes, including uncontrolled cell proliferation, and rapid progression of myeloid
STAT3, CBFB, and FUS and impairs telophase chromatin separation leukemogenesis (An et al., 2015). Moreover, enhancing long‐term
in hematopoietic cells, resulting in the increase of dysplastic myeloid repopulating and self‐renewal capacity of LSK cells can result in the
cells with disrupted chromatin separation and the development of expansion of common myeloid progenitors in FL by inactivating TET2
myeloid malignancies (Z. Li et al., 2017). Therefore, it can be con- in FL (Kunimoto et al., 2012). TET2 is not only effective in BM but has
cluded that ASXL1 promotes susceptibility to myeloid transformation an impact on restricting the expansion and function of HSCs in FL.
via changing histone modifications and disturbing hematopoiesis. Besides this, TET2 is vitally important in erythropoiesis because
upregulation of c‐Kit as well as AXL and ERK tyrosine kinases re-
sulting from knockdown of TET2 cause clonal expansion of the ery-
4.2 | Ten‐eleven translocation 2 throid progenitor (Qu et al., 2018). In addition, knockdown studies of
TET2 or TET3 have distinct consequences, so that TET2 knockdown
Expression and activity of TET2 have a vital role in normal home- causes defects in early stages of erythropoiesis, resulting in hyper-
ostasis through controlling the function, self‐renewal, proliferation, proliferation and impaired differentiation of erythroid progenitors,
and differentiation of HSCs. Low levels of 5hmC and 5mC accumu- while defects in the late stages of erythropoiesis caused by TET3
lation at certain genomic locations in the absence of mutations of knockdown lead to rising apoptosis, generation of bi/multinucleated
TET2 cause deregulation of DNA methylation (Feng et al., 2019; Ko & polychromatic/orthochromatic erythroblasts, and impaired enuclea-
Rao, 2011), and result in a progressive defect of hematopoiesis, in- tion (Yan et al., 2017). Cooperation of TET2 mutations with other
cluding increase in HSCs self‐renewal and BM cellularity, abnormal gene mutations promotes the development/transformation of var-
HSCs proliferation or differentiation, and myeloid (promoted mono- ious types of hematological malignancies. This fact was successfully
cyte/macrophage differentiation) or lymphoid neoplasms (Holmfeldt described by Jin et al. (2018) when they revealed that NRASG12D/+
& Mullighan, 2011; M. Ko et al., 2011; Mercher et al., 2012; Pronier and TET2−/+ cause expansion of HSCs pool and multipotent pro-
et al., 2011). In fact, disruption of TET2 function and expression genitors with promoting cytokine hypersensitivity, highly preserved
contributes to the alteration of B‐ and T‐lymphoid differentiation and self‐renewal capacity, and competitiveness, which contribute to the
expansion of immature progenitors endowed with myeloid colony‐ acceleration of preleukemia (e.g., CMML‐like disease; Jin et al., 2018).
forming potential (Mercher et al., 2012; Quivoron et al., 2011). TET2 It was also reported that the expression of lineage‐specific TFs in-
is expressed in low levels in ESCs, but its level increases in hema- creases as HSCs differentiate, including upregulation of erythroid
topoietic cells along with differentiation. It has been seen that TET2 regulators (e.g., KLF1 and EPOR) if there is a double knockout of
knockdown in ESCs reduces the number and cloning capacities of TET2 and DNMT3A (X. Zhang et al., 2016). In addition, given the
hematopoietic progenitors, whereas its knockdown in HSCs increases protective role of TET2 in HSPCs genomes (susceptibility to the ac-
hematopoietic development (Langlois et al., 2014). Although TET2 is quisition of DNA mutagenicity), increased accumulation of numerous
expressed in high levels among myeloid progenitors, the loss of TET2 mutations, including APC, NF1, FLT3, CBL, Notch1, and MLL2 has
catalytic activity could result in myeloid leukemogenesis through been observed when there is TET2−/− (Pan et al., 2017). There is a
restricting self‐renewal capacity and increasing the proliferation of lower frequency of TET1 mutation in hematopoietic malignancies
premalignant cells (M. Ko et al., 2010; Kunimoto et al., 2014). As a than TET2/3 mutation. Overexpression of BCL‐2 decreases 5hmC
result, TET2 activity is indispensable for the suppression of myeloid and increases 5mC in loss of TET1, increasing self‐renewal of pro-
malignancy (Z. Zhao et al., 2016). Also, it has been observed that genitor B‐cells and the development of B‐cell lymphoma (L. Cimmino
10 | SHOKOUHIAN ET AL.

et al., 2015). Considering that TET2 mutations play a role in myeloid blocks differentiation (Challen et al., 2014). As primitive mutations in
and lymphoid neoplasms, restoring TET2 function is likely to be HSPCs, DNMT3A, TET2, and ASXL1 mutations can be involved in
useful in therapeutic strategies and supporting evidence in this clonal hematopoiesis and generation of clonal preleukemic stem cells
respect was provided by Cimmino et al. (2017), who showed that (pre‐LSCs; Kunimoto & Nakajima, 2017; Sato, Wheat, Steidl, &
restoration of TET2 could inhibit aberrant self‐renewal of pre- Ito, 2016). Pre‐LSCs are similar to HSCs in their differentiation and
leukemic stem cells because treatment with vitamin C as a cofactor self‐renewal capacity and can produce LSCs following the second
of Fe (II) and a‐KG is accompanied with increased restoration of TET mutations (Chan & Majeti, 2013; Tan et al., 2015). Clonal hemato-
activity in leukemia cells (Cimmino et al., 2017; Pan et al., 2015). poiesis caused by DNMT3A mutations and FLT3‐ITD may lead to
Vitamin C acts as a pharmacologic mimic of TET2 restoration and is leukemogenesis in multiple lineages, as it was shown that hetero-
beneficial in promoting DNA demethylation, reversing aberrant zygous and homozygous DNMT3A mutations with FLT3‐ITD could be
HSPCs self‐renewal and enhancing the sensitivity of leukemia cells to effective in myeloid and lymphoid leukemias, respectively (L. Yang
PARP inhibition (Cimmino et al., 2017). et al., 2016). Also, coexistence of DNMT3A mutations with the ac-
quisition of mutant nucleophosmin plays a role in the development of
MPD and final transformation to AML (Loberg et al., 2019). DNMT1,
4.3 | DNA methyltransferase 3A which is a member of DNMT, has a crucial role in HSCs' self‐renewal
and differentiation of HSCs to myeloid progenitors. In contrast, loss
DNA methylation of HSCs by DNMT is vital in protecting the self‐ of DNMT1 can increase cycling and differentiation of myeloid‐
renewal of HSCs and multipotent progenitors by silencing pre- restricted progenitors (Trowbridge, Snow, Kim, & Orkin, 2009). Dif-
dominant differentiation programs. Lack of methylation, elevated ferentiation of myeloid‐erythroid progenitors increases when the
H3K4me3, and incomplete repression of HSCs‐specific genes in the expression level of myeloid‐erythroid genes is increased following
mutation of DNMT3A activate the self‐renewal pathway and expand hypomethylation of GATA1, ID2, CD48, and C/EBPα transcription
the number of HSCs in BM (Challen et al., 2011). Indeed, upregula- factors in HSCs lacking DNMT1 (Broske et al., 2009). Also, increased
tion of the Hox gene and myeloid ecotropic viral integration 1 (Meis1) expression of C/EBPα through hypomethylation of CpG islands in
due to DNMT3A mutations could be followed by enhanced HSCs self DNMT1 mutant HSPCs results in defective HSPCs maintenance and
‐renewal (Tan, Liu, & Chen, 2015). These mutations are effective both induced HSPCs proliferation (X. Liu et al., 2015). Therefore, the
on LT‐HSCs expansion and on increasing differentiation blockage of DNMT family is critical in controlling differentiation and self‐renewal
HSCs into all hematopoietic lineages (Brunetti et al., 2017; Chan & of the hematopoietic system where their mutations can initiate and
Majeti, 2013). A strong relationship was reported between loss of maintain leukemogenesis.
DNMT3A and initiation of the premalignant condition of clonal
hematopoiesis with indeterminate potential by M. Jeong et al. (2018),
in which it was shown that loss of DNMT3A significantly increased 4.4 | CCAAT‐enhancer‐binding protein α
the self‐renewal capacity of HSCs and produced HSCs with indefinite
longevity and immortality. It was also demonstrated that loss of C/EBPα is responsible for the maintenance of HSCs self‐renewal and
DNMT3A causes BM failure with MDS features such as BM hy- proliferation of myeloid progenitors and its deficiency increases
percellularity, the arrest of erythroid progenitors' development, HSCs self‐renewal, blocks differentiation of the common myeloid
proliferation enhancement of myeloid progenitors, and an increase in progenitor, and increases proliferation of myeloid blasts (Avellino &
apoptosis of mature myeloid cells (Celik et al., 2015). Loss of Delwel, 2017; Koschmieder et al., 2009; Ohlsson et al., 2016).
DNMT3A not only increases myeloproliferation in BM but expands Moreover, increasing expression of Bmi1 as a direct target of C/EBPα
LSK cells and multipotent progenitors in the liver (Guryanova in C/EBPα deficiency contributes to the enhanced competitive re-
et al., 2016). On the other hand, given the influence of interaction populating activity of HSCs (P. Zhang et al., 2004). It was found that
between DNMT3A and PcG on expression regulation of their target mutations in C‐terminal C/EBPα cause premalignant HSCs expansion
genes, it has been observed that mutation of DNMT3A causes and increase the proliferation of LT‐HSCs (Bereshchenko
aberrant recruitment of PRC1 to specific differentiation‐associated et al., 2009). Indeed, mutations in C‐terminal are effective in blocking
gene loci and plays a role in blocking myeloid differentiation of HSCs terminal differentiation (Quintana‐Bustamante et al., 2012). If
and promoting monoblastic transformation (Koya et al., 2016). In mutations in C‐terminal are homozygous, there is a decrease in
fact, mutation of this methyltransferase leads to a transformation of myeloid gene expression and an increase in erythroid‐specific gene
HSCs to myeloid (e.g., MDS, AML, and primary myelofibrosis) and expression (Bereshchenko et al., 2009). However, mutations in
lymphoid malignancies (e.g., T‐ and B‐ALL; Mayle et al., 2015). N‐terminal C/EBPα are associated with the silencing of HSCs ex-
Furthermore, the combined loss of DNMT3A and other mutations pansion and the formation of committed myeloid progenitors
such as c‐Kit, Notch1, NRASG12D, and KRAS in HSCs is effective in (Quintana‐Bustamante et al., 2012). In fact, mutations in the
leukemic transformation (Celik et al., 2015; Mayle et al., 2015). N‐terminal are effective in maintaining immature progenitors. The
Activation of the Wnt/β‐catenin signaling pathway in simultaneous combination of N‐ and C‐terminal C/EBPα mutations expands
deletion of DNMT3A and DNMT3B enhances HSCs self‐renewal and premalignant stem cells and accelerates disease development
SHOKOUHIAN ET AL. | 11

(Bereshchenko et al., 2009). As a transcription factor involved in results in leukemogenesis (Huang et al., 2013). It was reported that
blocking terminal differentiation of granulocytes, C/EBPα represses deficiency of MLL in HSCs has a role in exiting HSCs from a quiescent
E2F and has an important role in controlling the proliferative capa- state and increasing HSCs populations that undergo cell‐cycle entry.
city of early myeloid progenitors. Mutation of C/EBPα impairs E2F However, deficiency of MLL in myeloid‐erythroid progenitors re-
repression, which results in BM transformation with blocking duces the proliferation and frequency of these progenitors (Jude
granulocyte differentiation and expansion of myeloid progenitors et al., 2007). Actually, deficiency of MLL reduces the expression of
(Porse et al., 2005). In addition, lack of inhibition of E2F activity and Hox gene and antiapoptotic BCL‐2, which plays a role in blockage of
downregulation of c‐Myc following deletions of C/EBPα amino frequency and growth of hematopoietic colonies (Ernst, Mabon,
terminus and basic region result in granulocytic development Davidson, Zon, & Korsmeyer, 2004). Deletion of MLL results in a
(D'Alo et al., 2003). Given the role of C/EBPα in balancing differ- defect in the hematopoietic stem and progenitor pool, including re-
entiation and self‐renewal, dysregulation of this transcription factor duced numbers of quiescent HSCs, LT‐HSC, and ST‐HSC (McMahon
can play a causative role in the neoplastic process. et al., 2007). Therefore, dysregulation of HSCs self‐renewal and
expansion of progenitors following fusion or translocations of MLL
might contribute to leukemogenesis.
4.5 | Mixed‐lineage leukemia

MLL has a vital role in controlling HSCs' gene expression such as self‐ 4.6 | Phosphatase and tensin homolog deleted on
renewal, proliferation, and lineage‐specific differentiation, and the chromosome 10
translocations or fusion of this transcription coactivator are asso-
ciated with leukemogenesis (Yokoyama, 2015). The SET domain of Phosphatase PTEN has a role in the regulation of quiescence, pro-
MLL in MLL translocations increases the expression of Hox and other liferation, and differentiation of HSCs. PIP3 accumulation and hy-
genes, which transform hematopoietic cells into LSC (Krivtsov & peractivation of AKT mediated by PTEN deletion promote HSCs
Armstrong, 2007). In addition, MLL fusion proteins activate the Hox proliferation, increase levels of cycling HSCs and HSCs exhaustion,
gene and promote the self‐renewal of HSCs (Yokoyama, 2015). MLL reducing LT‐HSCs, and leading to myeloid leukemogenesis (Fragoso
fusion proteins bind to CpG islands within the Hox locus and facil- & Barata, 2014; Morotti et al., 2015). Activation of mTOR in HSCs via
itate TF binding, upregulating HoxA7, A9, and Meis1 and causing interaction of the P110B isoform of PI3K with RAC in PTEN defi-
immortalization of hematopoietic cells, blocking myeloid differ- ciency can result in the expansion of myeloid cells (Yuzugullu
entiation and inducing hematopoietic progenitors to leukemia et al., 2015). In addition, increase in PI3K signaling with mutation of
(Martin et al., 2003; Yokoyama, 2015). For instance, MLL‐AF4 fusion PTEN arrests HSPCs differentiation into mature cells, enhances
protein activates transcriptional programs that play a role in leuke- HSPCs proliferation and homing of HSPCs from BM to the caudal
mia through occupying regions of the genome affecting the devel- hematopoietic tissue and increases the numbers of mitotic cells in
opment and self‐renewal of HSCs and changing chromatin structure this tissue (Choorapoikayil, Kers, Herbomel, Kissa, & den
and histone modifications in these regions (Guenther et al., 2008). Hertog, 2014). Moreover, hyperactivation of the PI3K/AKT pathway
ENL binding to CBX8 (a component of PRC1) in MLL‐ENL fusion and in Gr‐1+CD11b+ BM cells (myeloid cells) with PTEN deletion pro-
neutralization of PRC1 activate gene loci necessary for transforma- duces high levels of G‐CSF, IL‐8, and CXCL12, which promote HSCs
tion to leukemia (Maethner et al., 2013). Furthermore, MLL‐PTD mobilization from BM to the spleen, splenomegaly, and leukemo-
fusion enhances HSCs self‐renewal, reducing the number of HSCs genesis (Tesio et al., 2013). It has been observed that both G‐CSF and
because of increased apoptosis, clonal expansion with lineage bias, inflammatory cytokines (e.g., interferon‐α) trigger HSCs mobilization
and blockage of myeloid differentiation (Y. Zhang et al., 2012). As the and expansion in the spleen through hyperactivation of the
interaction between MLL and RUNX1c/EBFβ prevents RUNX1 from PI3K/mTOR pathway in PTEN‐deficient HSCs (Porter et al., 2016).
ubiquitin‐mediated degradation, recruits MLL to regulate down- The PTEN mutant with activated β‐catenin promotes self‐renewal
stream target genes and causes HSCs differentiation (X. Zhao and expansion of HSCs with long‐term functional capacity without
et al., 2014), it was observed that downregulation of RUNX1c/EBFβ differentiation. In fact, activation of β‐catenin blocks differentiation
as a result of CXXC domain of MLL fusion proteins promotes through upregulation of ID2 and reduces multiple differentiation‐
self‐renewal, differentiation blockade, and leukemogenesis (X. Zhao inducing TFs, and the PTEN mutant enhances HSCs self‐renewal and
et al., 2014). TET1 as a fusion partner of MLL has been detected in t expansion through upregulation of antiapoptotic MCL1 (Perry
(10;11) (p12;q23) translocation (L. Cimmino et al., 2011; Mercher et al., 2011). It seems that the activation of Wnt/β‐catenin and PTEN/
et al., 2012). MLL fusion proteins directly bind to the promoter region PI3k/AKT pathways is effective in inducing proliferation and LT‐HSC
of TET1, increase the expression of TET1 and subsequently raise expansion with inhibiting apoptosis and blocking differentiation.
5hmC level. Then, TET1 along with MLL fusion proteins increases Therefore, this combination of findings provides some support for
transcriptional activation in critical cotargets including HoxA9, the conceptual promise that the inhibition of PI3K signaling in PTEN
Meis1, and pre‐B‐cell leukemia homeobox 3 to inhibit apoptosis and deficiency would be efficacious in therapeutic practices by decreas-
cell differentiation and promote cell proliferation, which finally ing hyperproliferation and inducing differentiation of cells.
12 | SHOKOUHIAN ET AL.

4.7 | RAS proteins family the expression of CD34 and megakaryocyte antigens (CD41/CD61)
on erythroblasts by promoting the expression and activation of
RAS proteins belongs to GTPase family and transfer signals from protein kinase C, which leads to dyserythropoiesis in preleukemia
various receptors (e.g., receptor tyrosine kinases, G protein‐coupled (Darley et al., 2002). On the other hand, hyperactivation of MAPK,
receptors, and cytokine receptors) to activate multiple downstream AKT, and STAT5 pathways and increasing monocyte‐colony‐forming
signaling pathways (including RAF/MEK/ERK and PI3K/AKT), there- unit (M‐CFU) colony formation in NRASG12D plays a role in the dif-
by playing a central role in the regulation of cellular growth and ferentiation of HSCs to myelomonocytic lineage cells and initiation of
proliferation, survival, and gene expression (Khan et al., 2019). Mu- MPD (T. Wang et al., 2015). Hyperactivity of MEK/ERK1/2 signaling
tation or aberrant activation of RAS proteins in humans (e.g., KRAS in NRASG12D/+ causes HSCs hyperproliferation along with enhanced
and NRAS) might lead to leukemogenesis via uncontrolled tran- self‐renewal capability, increases the number of LT‐HSCs and ex-
scriptional gene expression (Khan et al., 2019). Loss of KRAS‐WT pansion of myeloid progenitors so that downregulation of MEK/ERK
causes HSC depletion and increases the proliferation of myeloid signaling by pharmacologic and genetic approaches can be beneficial
precursors by triggering the activation of all RAS isoforms and hy- in decreasing these processes (J. Wang et al., 2013). It was reported
peractivation of cytokine signaling (such as GM‐CSF in myeloid cells), by J. Wang et al. (2011) that tumor transformation through en-
eventually inducing the mechanisms of leukemogenic activity (Kong dogenous oncogenic NRAS is related to the dose of NRAS and cell
et al., 2016). Also, loss of KRAS reduces TPO‐evoked ERK1/2 acti- type. NRASG12D/G12D in myeloid progenitors and T‐cell lineage pro-
vation in HSCs, resulting in a decline of HSCs self‐renewal, LT‐HSCs motes the proliferation of myeloid progenitors by hyperactivating
compartments and bias toward myeloid differentiation that could ERK signaling and increasing the development of acute T‐ALL, re-
reduce myeloid progenitors and neutrophil survival by reducing GM‐ spectively, leading to acute MPD. However, NRASG12D/+ results in
CSF‐evoked ERK1/2 activation in myeloid cells (Damnernsawad the development of CMML (95% of BM cells) and T‐ALL (8% of BM
et al., 2016). However, increased activity of KRAS‐ERK signaling in cells) (J. Wang et al., 2011). Furthermore, the co‐occurrence of
KRAS mutant induced pluripotent stem cells (iPSCs) as a result of NRASG12D and loss of TET2 in HSCs promotes TPO hypersensitivity
reduced basic fibroblast growth factor retains self‐renewal of un- in HSCs through activation of JAK/STAT signaling and preserves
G12D
differentiated cells in human iPSCs (Kubara et al., 2018). KRAS competitiveness and self‐renewal of HSCs, accelerating HSCs pro-
elevates the level of phosphorylated STAT5, ERK, and S6 in HSPCs liferation and promoting leukemogenesis (Jin et al., 2018). Moreover,
and sensitizes these cells to GM‐CSF stimulation, thereby increasing the cooperation of NRASG12D with the ASXL1 mutant can promote
the proliferation of HSPCs (Van Meter et al., 2007). Moreover, myeloid leukemogenesis (Abdel‐Wahab et al., 2012). Therefore, RAS
G12D
KRAS induces the initiation of leukemias via reducing the num- mutation in HSCs dysregulates the expression and function of sig-
ber of HSCs, increasing proliferation of primitive HSCs with en- naling cascade mechanisms involved in HSCs maintenance, thereby
hanced repopulating ability, and increasing the entry of HSCs to the playing a causative role in leukemia development.
cell cycle by upregulating cyclin D1 expression (Sabnis et al., 2009).
As the first genetic hit in HSCs, KRAS mutations have a role in
preserving the extensive proliferation capacity and survival of HSCs 4.8 | Runt‐related transcription factor 1
and may induce leukemic transformation in cooperation with sec-
ondary genetic hits (J. Zhang et al., 2009). For instance, a combina- RUNX plays a central role in regulating the interaction between
G12D
tion of KRAS with secondary Notch1 mutations in thymocytes is HSCs and the BM niche, so that the loss of RUNX1 leads to hy-
effective in initiating acute T‐lineage leukemia/lymphoma (Sabnis persensitivity of HSCs to G‐CSF, contributing to the enhanced mo-
et al., 2009). On the other hand, the coexistence of KRASG12D/+ and bilization of HSCs and LSK cells into the spleen and PB (Lam
DNMT3A mutations could promote myeloid malignancies. As the et al., 2016). G‐CSF hypersensitivity of RUNX1+/− cells reduces the
first genetic hit, DNMT3A drives clonal hematopoietic expansion and expression of C‐X‐C chemokine receptor type 4 and increases the
leukemogenesis after a prolonged latency, while it depletes HSCs and activation of STAT3 signaling by reducing the expression of protein
increases self‐renewal of myeloid progenitors in combination with inhibitor of activated STAT3, which leads to the expansion and mo-
KRASG12D/+, which finally results in the initiation and promotion of bilization of HSCs as well as blockage of myeloid differentiation (Chin
myeloid malignancies such as juvenile myelomonocytic leukemia, et al., 2016). It has been shown that the loss of RUNX1 is effective in
CMML, and AML (Chang et al., 2015). It has also been shown that increasing the number of HSCs having high self‐renewal capacity and
mutations of NRAS increase the proliferation of HSPCs and the that the clonal maintenance of leukemia‐initiating cells suppresses
quantity of myeloid lineage cells (S. W. Shen et al., 2004). Besides cellular fail‐safe mechanisms induced by oncogenic NRAS such as
this, the upregulation of cell‐cycle inhibitors, namely P21CIP1/WAF1 apoptosis via upregulating BCL‐2 expression, senescence by elevat-
and p16INK4a in NRAS mutations is associated with t growth sup- ing Bmi1 expression, and differentiation (Motoda et al., 2007). Loss of
pression of HSPCs, an increase in myelomonocytic colonies, and RUNX1 downmodulates apoptosis and elevates the percentage of
decrease in erythroid differentiation (S. Shen, Passioura, Symonds, & cells in the G1 cell cycle and increases LT‐HSCs (Cai et al., 2011).
Dolnikov, 2007). In fact, NRAS mutation increases early erythro- Selective deletion of RUNX1 results in increased activation and
blasts, inhibits differentiation in early erythroid cells, and upregulates quiescence of HSCs, rising number of LT‐HSCs, CMP, GMP, LSK cells
SHOKOUHIAN ET AL. | 13

expansion, and decreasing number of CLP, pro‐B, pre‐B, and mature RUNX1–ETO‐induced myeloid leukemogenesis (Asada & Kita-
B‐cells, leading to abnormalities in the development of T‐cells and mura, 2018; Asada et al., 2018). On the other hand, given the over-
thrombocytopenia (Kumano & Kurokawa, 2010). Similarly, a study by lapping of ASXL2 target genes with transcriptional targets of RUNX1
Behrens et al. (2016) showed that RUNX1 knockout could increase and RUNX1–ETO, it has been observed that ASXL2 deletion co‐
granulocytic/monocytic commitment and myeloid progenitors with operates with RUNX1–ETO in the promotion of leukemogenesis
skewing toward megakaryocyte progenitors. RUNX1 mutations in- (Micol et al., 2017). Taken together, regarding the vital role of
crease self‐renewal, enhance LT‐HSCs, and deregulate the expression RUNX1 in the regulation of hematopoietic development, disruption
of critical genes involved in granulocytic differentiation on specific of RUNX1 may lead to the initiation or progression of
loci (e.g., C/EBPα), which drive the inhibition of GMP differentiation leukemogenesis.
and leukemogenesis, including myeloid and lymphoid leukemias
(Gerritsen et al., 2019). In addition, RUNX1 mutations might lead to
the creation and expansion of preleukemic stem cells in BM by 5 | D IS C U S S I O N
lowering growth and biosynthesis, shrinking cell phenotype, reducing
ribosome biogenesis, and reducing apoptosis (Cai et al., 2015). Normal hematopoiesis is controlled by gene expression patterns
Therefore, RUNX1 deficiency may drive the progression of leukemia that guard the balance between self‐renewal and differentiation of
via dysregulating apoptosis and proliferation. As P53 is effective in HSCs; so that several epigenetic modifiers and TFs are involved in
steady‐state hematopoiesis, it regulates HSCs quiescence and self‐ the regulation of gene expression and fate in these cells by
renewal by blocking cell‐cycle entry (Asai, Liu, Bae, & Nimer, 2011; Y. activation or repression of specific genes. The activation of a
Liu et al., 2009). P53 reduction (P53+/− and P53−/−) downmodulates specific gene locus depends on the ability of the accessibility of
apoptosis and increases the quantity of proliferating HSCs, leading to DNA to TFs. TFs facilitate gene expression through binding to the
the initiation and progression of leukemic cells (Asai et al., 2011; enhancer or promoter regions of DNA via modifying the chromatin
Dumble et al., 2007). Also, P53 mutation enhances HSCs' self‐ structure and making genes available for transcription. These
renewal and promotes HSCs' expansion, leading to clonal expansion processes occur following modification of chromatin structures via
of HSCs and leukemic transformation (S. Chen et al., 2018; Dumble histone methylation, enrichment of specific histone marks
et al., 2007; Nabinger et al., 2019). Besides this, P53 mutation in (H3K4me3) by histone modifiers such as the ASXL1/BAP1 complex
combination with second hits (e.g., FLT3‐ITD) could be effective in and MLL, acetylation of histone proteins by transferring the acetyl
the expansion of myeloid progenitor cells (Nabinger et al., 2019). group from acetyl‐CoA to specific lysine residues, as well as DNA
Interestingly, the oncogenic fusion protein of RUNX1 has a causative demethylation via producing 5hmC by the TET family. Further-
role in leukemogenesis. RUNX1‐TEL fusion increases the number of more, transcription silencing in specific genes of HSCs can occur
HSCs while retaining their self‐renewal capacity, blocking differ- via chromatin compaction and repressive epigenetic modifications
entiation of B‐cells, and enhancing self‐renewal of B‐cell precursors, such as the H3K27me3 histone marker, histone deacetylation, and
resulting in the induction of B‐lineage‐committed leukemic cells DNA methylation at CpG islands via producing 5mC by DNMTs.
(Morrow, Horton, Kioussis, Brady, & Williams, 2004; Schindler On the other hand, aberrant interaction in these regulators might
et al., 2009). Another RUNX1 fusion, namely, RUNX1–ETO, enhances lead to leukemogenesis via altering the expression of functional
the self‐renewal of HSPCs, inhibits colony formation, and blocks the genes that manage self‐renewal, lineage commitment, differ-
proliferation of committed progenitors, leading to leukemogenesis, entiation, and cell‐cycle progression of HSCs (Figure 1). For
and the accumulation of both immature hematopoietic blast cells and instance, some primitive mutations such as DNMT3A, TET2, and
myeloid cells (Mulloy et al., 2002). This fusion is able to change he- ASXL1 in HSPCs cause clonal hematopoiesis of indeterminate
matopoietic cell fate so that RUNX1–ETO downregulates SCL and potential, which combined with secondary mutations, contribute
GATA1 expression and upregulate PU.1 in myeloid‐erythroid pro- to leukemic transformation, especially myeloid leukemias. In ad-
genitors, which suppress erythropoiesis and promote myelopoiesis dition, a decline in the repression of genes involved in maintenance
(Yeh et al., 2008). The increased COX/β‐catenin signaling pathway and differentiation bias of HSCs, caused by decreased levels of
caused by this fusion enhances hematopoietic self‐renewal and leu- H3K27me3 and 5hmC in mutations of ASXL1 and TET2, may
kemogenesis (Y. Zhang et al., 2013). Increased self‐renewal of pro- trigger the differentiation of HSCs toward myeloid and lymphoid
genitor cells, decreased G‐CFU, and increased M‐CFU as a result of lineages. Furthermore, dysregulation in chromatin modifications
RUNX1–ETO fusion inhibits the differentiation of granulocytic cells, and signaling pathways of HSCs as a result of aberrant TFs ex-
which promote the long‐term growth of myeloid cells as well as an pression related to the maintenance and differentiation of HSCs
expansion of immature granulocytes and myeloblasts in BM (Tonks could contribute to the initiation or acceleration of leukemias. It
et al., 2004). Also, an increase in the motility of preleukemic cells and appears that epigenetic regulators and TFs have important roles in
a decrease in the adhesion of these cells to stromal cells can drive the protecting self‐renewal of both HSCs and LSCs. Therefore, the
migration of preleukemic cells across the BM barrier and cause dis- evaluation of changes in TFs and epigenetic regulators as well as
ease progression in RUNX1–ETO cases (Saia et al., 2016). RUNX1– their functional pathways in genomic instability of HSCs could
ETO in association with ASXL1 mutations results in promoted yield high‐value targets for leukemia therapy.
14 | SHOKOUHIAN ET AL.

A C K N O W L E D GM E N T Nature Communications, 6, 7307. https://doi.org/10.1038/


We wish to thank all our colleagues in Shafa Hospital and Allied ncomms8307
Behrens, K., Triviai, I., Schwieger, M., Tekin, N., Alawi, M., Spohn, M., …
Health Sciences School, Ahvaz Jundishapur University of Medical
Stocking, C. (2016). Runx1 downregulates stem cell and
Sciences. megakaryocytic transcription programs that support niche
interactions. Blood, 127(26), 3369–3381. https://doi.org/10.1182/
CO NFLICT OF I NTERE STS blood‐2015‐09‐668129
Bereshchenko, O., Mancini, E., Moore, S., Bilbao, D., Månsson, R., Luc, S., …
The authors declare that there are no conflict of interests.
Nerlov, C. (2009). Hematopoietic stem cell expansion precedes the
generation of committed myeloid leukemia‐initiating cells in C/EBPα
A UT HO R C ONT RI BU TIO NS mutant AML. Cancer Cell, 16(5), 390–400. https://doi.org/10.1016/j.
N. S. conceived the manuscript and revised it; M. Sh., M. B., B. P., R. ccr.2009.09.036
Bocker, M. T., Hellwig, I., Breiling, A., Eckstein, V., Ho, A. D., & Lyko, F.
Ch., and N. D. wrote the manuscript and prepared the figures.
(2011). Genome‐wide promoter DNA methylation dynamics of human
hematopoietic progenitor cells during differentiation and aging. Blood,
E TH I C A L A P P R O V A L 117(19), e182–e189. https://doi.org/10.1182/blood‐2011‐01‐331926
This article does not contain any studies with human participants or Bocker, M. T., Tuorto, F., Raddatz, G., Musch, T., Yang, F. C., Xu, M., …
animals performed by any of the authors. Breiling, A. (2012). Hydroxylation of 5‐methylcytosine by TET2
maintains the active state of the mammalian HOXA cluster. Nature
Communications, 3, 818. https://doi.org/10.1038/ncomms1826
OR CID Bowman, R. L., & Levine, R. L. (2017). TET2 in normal and malignant
Mohammad Shokouhian http://orcid.org/0000-0001-5061-7058 hematopoiesis. Cold Spring Harbor Perspectives in Medicine, 7(8),
Marziye Bagheri http://orcid.org/0000-0003-3479-0277 a026518. https://doi.org/10.1101/cshperspect.a026518
Broske, A. M., Vockentanz, L., Kharazi, S., Huska, M. R., Mancini, E.,
Najmaldin Saki http://orcid.org/0000-0001-8494-5594
Scheller, M., … Rosenbauer, F. (2009). DNA methylation protects
hematopoietic stem cell multipotency from myeloerythroid
REFERENC ES restriction. Nature Genetics, 41(11), 1207–1215. https://doi.org/10.
Abdel‐Wahab, O., Adli, M., LaFave, L. M., Gao, J., Hricik, T., Shih, A. H., … 1038/ng.463
Levine, R. L. (2012). ASXL1 mutations promote myeloid transformation Brunetti, L., Gundry, M. C., & Goodell, M. A. (2017). DNMT3A in leukemia.
through loss of PRC2‐mediated gene repression. Cancer Cell, 22(2), Cold Spring Harbor Perspectives in Medicine, 7(2), a030320. https://doi.
180–193. https://doi.org/10.1016/j.ccr.2012.06.032 org/10.1101/cshperspect.a030320
Abdel‐Wahab, O., Gao, J., Adli, M., Dey, A., Trimarchi, T., Chung, Y. R., … Cai, X., Gao, L., Teng, L., Ge, J., Oo, Z. M., Kumar, A. R., … Speck, N. A.
Levine, R. L. (2013). Deletion of Asxl1 results in myelodysplasia and (2015). Runx1 deficiency decreases ribosome biogenesis and confers
severe developmental defects in vivo. Journal of Experimental Medicine, stress resistance to hematopoietic stem and progenitor cells. Cell Stem
210(12), 2641–2659. https://doi.org/10.1084/jem.20131141 Cell, 17(2), 165–177. https://doi.org/10.1016/j.stem.2015.06.002
An, J., Gonzalez‐Avalos, E., Chawla, A., Jeong, M., Lopez‐Moyado, I. F., Cai, X., Gaudet, J. J., Mangan, J. K., Chen, M. J., De Obaldia, M. E., Oo, Z., …
Li, W., … Rao, A. (2015). Acute loss of TET function results in Speck, N. A. (2011). Runx1 loss minimally impacts long‐term
aggressive myeloid cancer in mice. Nature Communications, 6, 10071. hematopoietic stem cells. PLoS One, 6(12):e28430. https://doi.org/
https://doi.org/10.1038/ncomms10071 10.1371/journal.pone.0028430
Artinger, E. L., Mishra, B. P., Zaffuto, K. M., Li, B. E., Chung, E. K., Moore, A. Carbuccia, N., Murati, A., Trouplin, V., Brecqueville, M., Adelaide, J., Rey, J.,
W., … Ernst, P. (2013). An MLL‐dependent network sustains … Mozziconacci, M. J. (2009). Mutations of ASXL1 gene in
hematopoiesis. Proceedings of the National Academy of Sciences of the myeloproliferative neoplasms. Leukemia, 23(11), 2183–2186. https://
United States of America, 110(29), 12000–12005. https://doi.org/10. doi.org/10.1038/leu.2009.141
1073/pnas.1301278110 Celik, H., Mallaney, C., Kothari, A., Ostrander, E. L., Eultgen, E.,
Asada, S., Fujino, T., Goyama, S., & Kitamura, T. (2019). The role of ASXL1 Martens, A., … Challen, G. A. (2015). Enforced differentiation of
in hematopoiesis and myeloid malignancies. Cellular and Molecular Life DNMT3a‐null bone marrow leads to failure with c‐Kit mutations
Sciences, 76(13), 2511–2523. https://doi.org/10.1007/s00018‐019‐ driving leukemic transformation. Blood, 125(4), 619–628. https://doi.
03084‐7 org/10.1182/blood‐2014‐08‐594564
Asada, S., Goyama, S., Inoue, D., Shikata, S., Takeda, R., Fukushima, T., … Challen, G. A., Sun, D., Jeong, M., Luo, M., Jelinek, J., Berg, J. S., …
Kitamura, T. (2018). Mutant ASXL1 cooperates with BAP1 to promote Goodell, M. A. (2011). DNMT3a is essential for hematopoietic stem
myeloid leukaemogenesis. Nature Communications, 9(1), 2733. https:// cell differentiation. Nature Genetics, 44(1), 23–31. https://doi.org/10.
doi.org/10.1038/s41467‐018‐05085‐9 1038/ng.1009
Asada, S., & Kitamura, T. (2018). Aberrant histone modifications induced Challen, G. A., Sun, D., Mayle, A., Jeong, M., Luo, M., Rodriguez, B., …
by mutant ASXL1 in myeloid neoplasms. International Journal of Goodell, M. A. (2014). DNMT3a and DNMT3b have overlapping and
Hematology, 110(2), 179–186. https://doi.org/10.1007/s12185‐018‐ distinct functions in hematopoietic stem cells. Cell Stem Cell, 15(3),
2563‐7 350–364. https://doi.org/10.1016/j.stem.2014.06.018
Asai, T., Liu, Y., Bae, N., & Nimer, S. D. (2011). The p53 tumor suppressor Chan, S. M., & Majeti, R. (2013). Role of DNMT3A, TET2, and IDH1/2
protein regulates hematopoietic stem cell fate. Journal of Cellular mutations in pre‐leukemic stem cells in acute myeloid leukemia.
Physiology, 226(9), 2215–2221. https://doi.org/10.1002/jcp.22561 International Journal of Hematology, 98(6), 648–657. https://doi.org/10.
Avellino, R., & Delwel, R. (2017). Expression and regulation of C/EBPα in 1007/s12185‐013‐1407‐8
normal myelopoiesis and in malignant transformation. Blood, 129(15), Chang, Y. I., You, X., Kong, G., Ranheim, E. A., Wang, J., Du, J., … Zhang, J.
2083–2091. https://doi.org/10.1182/blood‐2016‐09‐687822 (2015). Loss of DNMT3a and endogenous Kras(G12D/+) cooperate to
Balasubramani, A., Larjo, A., Bassein, J. A., Chang, X., Hastie, R. B., regulate hematopoietic stem and progenitor cell functions in
Togher, S. M., … Rao, A. (2015). Cancer‐associated ASXL1 mutations leukemogenesis. Leukemia, 29(9), 1847–1856. https://doi.org/10.
may act as gain‐of‐function mutations of the ASXL1‐BAP1 complex. 1038/leu.2015.85
SHOKOUHIAN ET AL. | 15

Chen, Q., Chen, Y., Bian, C., Fujiki, R., & Yu, X. (2013). TET2 promotes Bioscience Reports, 38(1), BSR20171589. https://doi.org/10.1042/
histone O‐GlcNAcylation during gene transcription. Nature, bsr20171589
493(7433), 561–564. https://doi.org/10.1038/nature11742 Dorritie, K. A., Redner, R. L., & Johnson, D. E. (2014). STAT transcription
Chen, S., Gao, R., Yao, C., Kobayashi, M., Liu, S. Z., Yoder, M. C., … Liu, Y. factors in normal and cancer stem cells. Advances in Biological
(2018). Genotoxic stresses promote clonal expansion of Regulation, 56, 30–44. https://doi.org/10.1016/j.jbior.2014.05.004
hematopoietic stem cells expressing mutant p53. Leukemia, 32(3), Dumble, M., Moore, L., Chambers, S. M., Geiger, H., Van Zant, G.,
850–854. https://doi.org/10.1038/leu.2017.325 Goodell, M. A., & Donehower, L. A. (2007). The impact of altered p53
Chen, Y., Costa, R. M., Love, N. R., Soto, X., Roth, M., Paredes, R., & dosage on hematopoietic stem cell dynamics during aging. Blood,
Amaya, E. (2009). C/EBPalpha initiates primitive myelopoiesis in 109(4), 1736–1742. https://doi.org/10.1182/blood‐2006‐03‐010413
pluripotent embryonic cells. Blood, 114(1), 40–48. https://doi.org/10. Ernst, P., Mabon, M., Davidson, A. J., Zon, L. I., & Korsmeyer, S. J. (2004).
1182/blood‐2008‐11‐189159 An Mll‐dependent Hox program drives hematopoietic progenitor
Chin, D. W., Sakurai, M., Nah, G. S., Du, L., Jacob, B., Yokomizo, T., … expansion. Current Biology, 14(22), 2063–2069. https://doi.org/10.
Osato, M. (2016). RUNX1 haploinsufficiency results in granulocyte 1016/j.cub.2004.11.012
colony‐stimulating factor hypersensitivity. Blood Cancer Journal, 6, Ezhkova, E., Pasolli, H. A., Parker, J. S., Stokes, N., Su, I. H., Hannon, G., …
e379. https://doi.org/10.1038/bcj.2015.105 Fuchs, E. (2009). Ezh2 orchestrates gene expression for the stepwise
Choorapoikayil, S., Kers, R., Herbomel, P., Kissa, K., & den Hertog, J. differentiation of tissue‐specific stem cells. Cell, 136(6), 1122–1135.
(2014). Pivotal role of PTEN in the balance between proliferation and https://doi.org/10.1016/j.cell.2008.12.043
differentiation of hematopoietic stem cells in zebrafish. Blood, 123(2), Feng, Y., Li, X., Cassady, K., Zou, Z., & Zhang, X. (2019). TET2 function in
184–190. https://doi.org/10.1182/blood‐2013‐05‐501544 hematopoietic malignancies, immune regulation, and DNA repair.
Cimmino, L., Abdel‐Wahab, O., Levine, R. L., & Aifantis, I. (2011). TET Frontiers in Oncology, 9, 210. https://doi.org/10.3389/fonc.2019.
family proteins and their role in stem cell differentiation and 00210
transformation. Cell Stem Cell, 9(3), 193–204. https://doi.org/10. Fragoso, R., & Barata, J. T. (2014). PTEN and leukemia stem cells. Advances
1016/j.stem.2011.08.007 in Biological Regulation, 56, 22–29. https://doi.org/10.1016/j.jbior.
Cimmino, L., Dawlaty, M. M., Ndiaye‐Lobry, D., Yap, Y. S., Bakogianni, S., 2014.05.005
Yu, Y., … Aifantis, I. (2015). TET1 is a tumor suppressor of Fukuchi, Y., Ito, M., Shibata, F., Kitamura, T., & Nakajima, H. (2008).
hematopoietic malignancy. Nature Immunology, 16(6), 653–662. Activation of CCAAT/enhancer‐binding protein alpha or PU.1 in
https://doi.org/10.1038/ni.3148 hematopoietic stem cells leads to their reduced self‐renewal and
Cimmino, L., Dolgalev, I., Wang, Y., Yoshimi, A., Martin, G. H., Wang, J., … proliferation. Stem Cells, 26(12), 3172–3181. https://doi.org/10.1634/
Mitchell‐Flack, M. (2017). Restoration of TET2 function blocks stemcells.2008‐0320
aberrant self‐renewal and leukemia progression. Cell, 170(6), Fukuchi, Y., Shibata, F., Ito, M., Goto‐Koshino, Y., Sotomaru, Y., Ito, M., …
1079–1095. https://doi.org/10.1016/j.cell.2017.07.032 Nakajima, H. (2006). Comprehensive analysis of myeloid lineage
Cui, K., Zang, C., Roh, T. Y., Schones, D. E., Childs, R. W., Peng, W., & conversion using mice expressing an inducible form of C/EBP alpha.
Zhao, K. (2009). Chromatin signatures in multipotent human The EMBO Journal, 25(14), 3398–3410. https://doi.org/10.1038/sj.
hematopoietic stem cells indicate the fate of bivalent genes during emboj.7601199
differentiation. Cell Stem Cell, 4(1), 80–93. https://doi.org/10.1016/j. Gerritsen, M., Yi, G., Tijchon, E., Kuster, J., Schuringa, J. J., Martens, J. H.
stem.2008.11.011 A., & Vellenga, E. (2019). RUNX1 mutations enhance self‐renewal and
D'Alo, F., Johansen, L. M., Nelson, E. A., Radomska, H. S., Evans, E. K., block granulocytic differentiation in human in vitro models and
Zhang, P., … Tenen, D. G. (2003). The amino terminal and E2F primary AMLs. Blood Advances, 3(3), 320–332. https://doi.org/10.
interaction domains are critical for C/EBP alpha‐mediated induction 1182/bloodadvances.2018024422
of granulopoietic development of hematopoietic cells. Blood, 102(9), Gottgens, B. (2015). Regulatory network control of blood stem cells.
3163–3171. https://doi.org/10.1182/blood‐2003‐02‐0479 Blood, 125(17), 2614–2620. https://doi.org/10.1182/blood‐2014‐08‐
Damnernsawad, A., Kong, G., Wen, Z., Liu, Y., Rajagopalan, A., You, X., … 570226
Zhang, J. (2016). Kras is required for adult hematopoiesis. Stem Cells, Gu, L., Chiang, K. Y., Zhu, N., Findley, H. W., & Zhou, M. (2007).
34(7), 1859–1871. https://doi.org/10.1002/stem.2355 Contribution of STAT3 to the activation of survivin by GM‐CSF in
Darley, R. L., Pearn, L., Omidvar, N., Sweeney, M., Fisher, J., Phillips, S., … CD34+ cell lines. Experimental Hematology, 35(6), 957–966. https://
Burnett, A. K. (2002). Protein kinase C mediates mutant N‐Ras‐ doi.org/10.1016/j.exphem.2007.03.007
induced developmental abnormalities in normal human erythroid Guenther, M. G., Lawton, L. N., Rozovskaia, T., Frampton, G. M., Levine, S.
cells. Blood, 100(12), 4185–4192. https://doi.org/10.1182/blood‐ S., Volkert, T. L., … Young, R. A. (2008). Aberrant chromatin at genes
2002‐05‐1358 encoding stem cell regulators in human mixed‐lineage leukemia. Genes
Davies, C., Yip, B. H., Fernandez‐Mercado, M., Woll, P. S., Agirre, X., & Development, 22(24), 3403–3408. https://doi.org/10.1101/gad.
Prosper, F., … Boultwood, J. (2013). Silencing of ASXL1 impairs the 1741408
granulomonocytic lineage potential of human CD34(+) progenitor Guo, Y., Yang, H., Chen, S., Zhang, P., Li, R., Nimer, S. D., … Yang, F. C.
cells. British Journal of Haematology, 160(6), 842–850. https://doi.org/ (2018). Reduced BAP1 activity prevents ASXL1 truncation‐driven
10.1111/bjh.12217 myeloid malignancy in vivo. Leukemia, 32(8), 1834–1837. https://doi.
Deplus, R., Delatte, B., Schwinn, M. K., Defrance, M., Mendez, J., org/10.1038/s41375‐018‐0126‐9
Murphy, N., … Fuks, F. (2013). TET2 and TET3 regulate GlcNAcylation Guryanova, O. A., Lieu, Y. K., Garrett‐Bakelman, F. E., Spitzer, B., Glass, J.
and H3K4 methylation through OGT and SET1/COMPASS. The EMBO L., Shank, K., … Levine, R. L. (2016). Dnmt3a regulates
Journal, 32(5), 645–655. https://doi.org/10.1038/emboj.2012.357 myeloproliferation and liver‐specific expansion of hematopoietic
Dong, M., & Blobe, G. C. (2006). Role of transforming growth factor‐beta stem and progenitor cells. Leukemia, 30(5), 1133–1142. https://doi.
in hematologic malignancies. Blood, 107(12), 4589–4596. https://doi. org/10.1038/leu.2015.358
org/10.1182/blood‐2005‐10‐4169 Hasemann, M. S., Lauridsen, F. K., Waage, J., Jakobsen, J. S., Frank, A.‐K.,
Dong, Y., Lian, X., Xu, Y., Hu, H., Chang, C., Zhang, H., & Zhang, L. (2018). Schuster, M. B., … Schroeder, T. (2014). C/EBPα is required for long‐
Hematopoietic stem/progenitor cell senescence is associated with term self‐renewal and lineage priming of hematopoietic stem cells and
altered expression profiles of cellular memory‐involved gene. for the maintenance of epigenetic configurations in multipotent
16 | SHOKOUHIAN ET AL.

progenitors. PLoS Genetics, 10(1):e1004079. https://doi.org/10.1371/ hematopoietic stem and progenitor cells. Blood Advance, 2(11),
journal.pgen.1004079 1259–1271. https://doi.org/10.1182/bloodadvances.2018017400
Herviou, L., Cavalli, G., Cartron, G., Klein, B., & Moreaux, J. (2016). EZH2 Jude, C. D., Climer, L., Xu, D., Artinger, E., Fisher, J. K., & Ernst, P. (2007).
in normal hematopoiesis and hematological malignancies. Oncotarget, Unique and independent roles for MLL in adult hematopoietic stem
7(3), 2284–2296. https://doi.org/10.18632/oncotarget.6198 cells and progenitors. Cell Stem Cell, 1(3), 324–337. https://doi.org/10.
Hidalgo, I., & Gonzalez, S. (2013). New epigenetic pathway for stemness 1016/j.stem.2007.05.019
maintenance mediated by the histone methyltransferase Ezh1. Cell Kallin, E. M., Rodriguez‐Ubreva, J., Christensen, J., Cimmino, L., Aifantis, I.,
Cycle, 12(3), 383–384. https://doi.org/10.4161/cc.23550 Helin, K., … Graf, T. (2012). Tet2 facilitates the derepression of
Hidalgo, I., Herrera‐Merchan, A., Ligos, J. M., Carramolino, L., Nunez, J., myeloid target genes during CEBPalpha‐induced transdifferentiation
Martinez, F., … Gonzalez, S. (2012). Ezh1 is required for hematopoietic of pre‐B cells. Molecular Cell, 48(2), 266–276. https://doi.org/10.1016/
stem cell maintenance and prevents senescence‐like cell cycle arrest. j.molcel.2012.08.007
Cell Stem Cell, 11(5), 649–662. https://doi.org/10.1016/j.stem.2012. Kamminga, L. M., Bystrykh, L. V., de Boer, A., Houwer, S., Douma, J.,
08.001 Weersing, E., … de Haan, G. (2006). The Polycomb group gene Ezh2
Hilgendorf, S., Folkerts, H., Schuringa, J. J., & Vellenga, E. (2016). Loss of prevents hematopoietic stem cell exhaustion. Blood, 107(5),
ASXL1 triggers an apoptotic response in human hematopoietic stem 2170–2179. https://doi.org/10.1182/blood‐2005‐09‐3585
and progenitor cells. Experimental Hematology, 44(12), 1188–1196. Kato, Y., Iwama, A., Tadokoro, Y., Shimoda, K., Minoguchi, M., Akira, S., …
https://doi.org/10.1016/j.exphem.2016.08.011 Nakauchi, H. (2005). Selective activation of STAT5 unveils its role in
Holmfeldt, L., & Mullighan, C. G. (2011). The role of TET2 in hematologic stem cell self‐renewal in normal and leukemic hematopoiesis. Journal
neoplasms. Cancer Cell, 20(1), 1–2. https://doi.org/10.1016/j.ccr.2011. of Experimental Medicine, 202(1), 169–179. https://doi.org/10.1084/
06.025 jem.20042541
Hong, D., Fritz, A. J., Gordon, J. A., Tye, C. E., Boyd, J. R., Tracy, K. M., … Katoh, M. (2013). Functional and cancer genomics of ASXL family
Stein, G. S. (2019). RUNX1‐dependent mechanisms in biological members. British Journal of Cancer, 109(2), 299–306. https://doi.org/
control and dysregulation in cancer. Journal of Cellular Physiology, 10.1038/bjc.2013.281
234(6), 8597–8609. https://doi.org/10.1002/jcp.27841 Khan, A. Q., Kuttikrishnan, S., Siveen, K. S., Prabhu, K. S.,
Hong, S. H., Yang, S. J., Kim, T. M., Shim, J. S., Lee, H. S., Lee, G. Y., … Oh, Shanmugakonar, M., Al‐Naemi, H. A., … Uddin, S. (2019). RAS‐
I. H. (2014). Molecular integration of HoxB4 and STAT3 for self‐ mediated oncogenic signaling pathways in human malignancies.
renewal of hematopoietic stem cells: A model of molecular Seminars in Cancer Biology, 54, 1–13. https://doi.org/10.1016/j.
convergence for stemness. Stem Cells, 32(5), 1313–1322. https://doi. semcancer.2018.03.001
org/10.1002/stem.1631 Ko, M., Bandukwala, H. S., An, J., Lamperti, E. D., Thompson, E. C.,
Hu, L., Li, Z., Cheng, J., Rao, Q., Gong, W., Liu, M., … Xu, Y. (2013). Crystal Hastie, R., … Rao, A. (2011). Ten‐eleven‐translocation 2 (TET2)
structure of TET2‐DNA complex: Insight into TET‐mediated 5mC negatively regulates homeostasis and differentiation of
oxidation. Cell, 155(7), 1545–1555. https://doi.org/10.1016/j.cell. hematopoietic stem cells in mice. Proceedings of the National
2013.11.020 Academy of Sciences of the United States of America, 108(35),
Huang, H., Jiang, X., Li, Z., Li, Y., Song, C. X., He, C., … Chen, J. (2013). TET1 14566–14571. https://doi.org/10.1073/pnas.1112317108
plays an essential oncogenic role in MLL‐rearranged leukemia. Ko, M., Huang, Y., Jankowska, A. M., Pape, U. J., Tahiliani, M.,
Proceedings of the National Academy of Sciences of the United States of Bandukwala, H. S., … Rao, A. (2010). Impaired hydroxylation of
America, 110(29), 11994–11999. https://doi.org/10.1073/pnas. 5‐methylcytosine in myeloid cancers with mutant TET2. Nature,
1310656110 468(7325), 839–843. https://doi.org/10.1038/nature09586
Ichikawa, M., Goyama, S., Asai, T., Kawazu, M., Nakagawa, M., Ko, M., & Rao, A. (2011). TET2: Epigenetic safeguard for HSC. Blood,
Takeshita, M., … Kurokawa, M. (2008). AML1/Runx1 negatively 118(17), 4501–4503.
regulates quiescent hematopoietic stem cells in adult hematopoiesis. Kong, G., Chang, Y. I., Damnernsawad, A., You, X., Du, J., Ranheim, E. A., …
Journal of Immunology, 180(7), 4402–4408. https://doi.org/10.4049/ Zhang, J. (2016). Loss of wild‐type Kras promotes activation of all Ras
jimmunol.180.7.4402 isoforms in oncogenic Kras‐induced leukemogenesis. Leukemia, 30(7),
Inoue, D., Fujino, T., Sheridan, P., Zhang, Y. Z., Nagase, R., Horikawa, S., … 1542–1551. https://doi.org/10.1038/leu.2016.40
Kitamura, T. (2018). A novel ASXL1‐OGT axis plays roles in H3K4 Koschmieder, S., Halmos, B., Levantini, E., & Tenen, D. G. (2009).
methylation and tumor suppression in myeloid malignancies. Dysregulation of the C/EBPα differentiation pathway in human
Leukemia, 32(6), 1327–1337. https://doi.org/10.1038/s41375‐018‐ cancer. Journal of Clinical Oncology, 27(4), 619–628. https://doi.org/
0083‐3 10.1200/JCO.2008.17.9812
Inoue, D., Kitaura, J., Togami, K., Nishimura, K., Enomoto, Y., Uchida, T., … Koya, J., Kataoka, K., Sato, T., Bando, M., Kato, Y., Tsuruta‐Kishino, T., …
Kitamura, T. (2013). Myelodysplastic syndromes are induced by Kurokawa, M. (2016). DNMT3A R882 mutants interact with
histone methylation‐altering ASXL1 mutations. Journal of Clinical polycomb proteins to block haematopoietic stem and leukaemic cell
Investigation, 123(11), 4627–4640. https://doi.org/10.1172/jci70739 differentiation. Nature Communications, 7, 10924. https://doi.org/10.
Inoue, S., Lemonnier, F., & Mak, T. W. (2016). Roles of IDH1/2 and TET2 1038/ncomms10924
mutations in myeloid disorders. International Journal of Hematology, Krivtsov, A. V., & Armstrong, S. A. (2007). MLL translocations, histone
103(6), 627–633. modifications and leukaemia stem‐cell development. Nature Reviews
Jeong, J. J., Gu, X., Nie, J., Sundaravel, S., Liu, H., Kuo, W. L., … Cancer, 7(11), 823–833. https://doi.org/10.1038/nrc2253
Wickrema, A. (2019). Cytokine‐regulated phosphorylation and Kubara, K., Yamazaki, K., Ishihara, Y., Naruto, T., Lin, H. T., Nishimura, K., …
activation of TET2 by JAK2 in hematopoiesis. Cancer Discovery, 9(6), Otsu, M. (2018). Status of KRAS in iPSCs impacts upon self‐renewal
778–795. https://doi.org/10.1158/2159‐8290.cd‐18‐1138 and differentiation propensity. Stem Cell Reports, 11(2), 380–394.
Jeong, M., Park, H. J., Celik, H., Ostrander, E. L., Reyes, J. M., Guzman, A., … https://doi.org/10.1016/j.stemcr.2018.06.008
Challen, G. A. (2018). Loss of DNMT3a immortalizes hematopoietic Kumano, K., & Kurokawa, M. (2010). The role of Runx1/AML1 and Evi‐1 in
stem cells in vivo. Cell Reports, 23(1), 1–10. https://doi.org/10.1016/j. the regulation of hematopoietic stem cells. Journal of Cellular
celrep.2018.03.025 Physiology, 222(2), 282–285. https://doi.org/10.1002/jcp.21953
Jin, X., Qin, T., Zhao, M., Bailey, N., Liu, L., Yang, K., … Li, Q. (2018). Oncogenic Kunimoto, H., Fukuchi, Y., Sakurai, M., Sadahira, K., Ikeda, Y., Okamoto, S.,
N‐Ras and Tet2 haploinsufficiency collaborate to dysregulate & Nakajima, H. (2012). TET2 disruption leads to enhanced
SHOKOUHIAN ET AL. | 17

self‐renewal and altered differentiation of fetal liver hematopoietic induces genetic instability and mutagenesis. DNA Repair (Amst), 43,
stem cells. Scientific Reports, 2, 273. https://doi.org/10.1038/ 78–88. https://doi.org/10.1016/j.dnarep.2016.05.031
srep00273 Martin, M. E., Milne, T. A., Bloyer, S., Galoian, K., Shen, W., Gibbs, D., …
Kunimoto, H., Fukuchi, Y., Sakurai, M., Takubo, K., Okamoto, S., & Hess, J. L. (2003). Dimerization of MLL fusion proteins immortalizes
Nakajima, H. (2014). TET2‐mutated myeloid progenitors possess hematopoietic cells. Cancer Cell, 4(3), 197–207.
aberrant in vitro self‐renewal capacity. Blood, 123(18), 2897–2899. Mayani, H. (2016). The regulation of hematopoietic stem cell populations.
https://doi.org/10.1182/blood‐2014‐01‐552471 F1000Research, 5, 1524.
Kunimoto, H., & Nakajima, H. (2017). Epigenetic dysregulation of Mayle, A., Yang, L., Rodriguez, B., Zhou, T., Chang, E., Curry, C. V., …
hematopoietic stem cells and preleukemic state. International Journal Goodell, M. A. (2015). DNMT3a loss predisposes murine
of Hematology, 106(1), 34–44. https://doi.org/10.1007/s12185‐017‐ hematopoietic stem cells to malignant transformation. Blood, 125(4),
2257‐6 629–638. https://doi.org/10.1182/blood‐2014‐08‐594648
Kurokawa, M. (2006). AML1/Runx1 as a versatile regulator of McMahon, K. A., Hiew, S. Y., Hadjur, S., Veiga‐Fernandes, H., Menzel, U.,
hematopoiesis: Regulation of its function and a role in adult Price, A. J., … Brady, H. J. (2007). Mll has a critical role in fetal and
hematopoiesis. International Journal of Hematology, 84(2), 136–142. adult hematopoietic stem cell self‐renewal. Cell Stem Cell, 1(3),
https://doi.org/10.1532/ijh97.06070 338–345. https://doi.org/10.1016/j.stem.2007.07.002
Lam, K., Muselman, A., Du, R., Yan, M., Matsuura, S., & Zhang, D. E. (2016). Mercher, T., Quivoron, C., Couronne, L., Bastard, C., Vainchenker, W., &
Loss of RUNX1 function results in enhanced granulocyte‐colony‐ Bernard, O. A. (2012). TET2, a tumor suppressor in hematological
stimulating factor‐mediated mobilization. Blood Cancer Journal, 6, disorders. Biochimica et Biophysica Acta/General Subjects, 1825(2),
e407. https://doi.org/10.1038/bcj.2016.20 173–177. https://doi.org/10.1016/j.bbcan.2011.12.002
Langlois, T., da Costa Reis Monte‐Mor, B., Lenglet, G., Droin, N., Marty, C., Micol, J. B., Pastore, A., Inoue, D., Duployez, N., Kim, E., Lee, S. C., & Abdel‐
Le Couedic, J. P., … Plo, I. (2014). TET2 deficiency inhibits mesoderm Wahab, O. (2017). ASXL2 is essential for haematopoiesis and acts as a
and hematopoietic differentiation in human embryonic stem cells. haploinsufficient tumour suppressor in leukemia. Nature
Stem Cells, 32(8), 2084–2097. https://doi.org/10.1002/stem.1718 Communications, 8, 15429. https://doi.org/10.1038/ncomms15429
Li, C., Lan, Y., Schwartz‐Orbach, L., Korol, E., Tahiliani, M., Evans, T., & Mochizuki‐Kashio, M., Aoyama, K., Sashida, G., Oshima, M., Tomioka, T.,
Goll, M. G. (2015). Overlapping requirements for Tet2 and Tet3 in Muto, T., & Iwama, A. (2015). Ezh2 loss in hematopoietic stem cells
normal development and hematopoietic stem cell emergence. Cell predisposes mice to develop heterogeneous malignancies in an Ezh1‐
Reports, 12(7), 1133–1143. https://doi.org/10.1016/j.celrep.2015. dependent manner. Blood, 126(10), 1172–1183. https://doi.org/10.
07.025 1182/blood‐2015‐03‐634428
Li, J., Zhang, J., Tang, M., Xin, J., Xu, Y., Volk, A., … Zhang, J. (2016). Mochizuki‐Kashio, M., Mishima, Y., Miyagi, S., Negishi, M., Saraya, A.,
Hematopoietic stem cell activity is regulated by PTEN Konuma, T., & Iwama, A. (2011). Dependency on the polycomb gene
phosphorylation through a niche‐dependent mechanism. Stem Cells, Ezh2 distinguishes fetal from adult hematopoietic stem cells. Blood,
34(8), 2130–2144. https://doi.org/10.1002/stem.2382 118(25), 6553–6561. https://doi.org/10.1182/blood‐2011‐03‐340554
Li, Z., Zhang, P., Yan, A., Guo, Z., Ban, Y., Li, J., … Yang, F. C. (2017). ASXL1 Morotti, A., Panuzzo, C., Crivellaro, S., Carrà, G., Torti, D., Guerrasio, A., &
interacts with the cohesin complex to maintain chromatid separation Saglio, G. (2015). The role of PTEN in myeloid malignancies.
and gene expression for normal hematopoiesis. Science Advances, 3(1): Hematology Reports, 7(4), 5844.
e1601602. https://doi.org/10.1126/sciadv.1601602 Morrow, M., Horton, S., Kioussis, D., Brady, H. J., & Williams, O. (2004).
Liu, X., Jia, X., Yuan, H., Ma, K., Chen, Y., Jin, Y., … Zhu, J. (2015). DNA TEL‐AML1 promotes development of specific hematopoietic lineages
methyltransferase 1 functions through C/ebpa to maintain consistent with preleukemic activity. Blood, 103(10), 3890–3896.
hematopoietic stem and progenitor cells in zebrafish. Journal of https://doi.org/10.1182/blood‐2003‐10‐3695
Hematology & Oncology, 8, 15. https://doi.org/10.1186/s13045‐015‐ Motoda, L., Osato, M., Yamashita, N., Jacob, B., Chen, L. Q., Yanagida, M., &
0115‐7 Ito, Y. (2007). Runx1 protects hematopoietic stem/progenitor cells
Liu, Y., Elf, S. E., Miyata, Y., Sashida, G., Liu, Y., Huang, G., … Nimer, S. D. from oncogenic insult. Stem Cells, 25(12), 2976–2986. https://doi.org/
(2009). p53 regulates hematopoietic stem cell quiescence. Cell Stem 10.1634/stemcells.2007‐0061
Cell, 4(1), 37–48. https://doi.org/10.1016/j.stem.2008.11.006 Mulloy, J. C., Cammenga, J., MacKenzie, K. L., Berguido, F. J., Moore, M. A.,
Loberg, M. A., Bell, R. K., Goodwin, L. O., Eudy, E., Miles, L. A., SanMiguel, J. & Nimer, S. D. (2002). The AML1‐ETO fusion protein promotes the
M., … Trowbridge, J. J. (2019). Sequentially inducible mouse models expansion of human hematopoietic stem cells. Blood, 99(1), 15–23.
reveal that Npm1 mutation causes malignant transformation of https://doi.org/10.1182/blood.v99.1.15
DNMT3a‐mutant clonal hematopoiesis. Leukemia, 33, 1635–1649. Nabinger, S. C., Chen, S., Gao, R., Yao, C., Kobayashi, M., Vemula, S., &
https://doi.org/10.1038/s41375‐018‐0368‐6. Liu, Y. (2019). Mutant p53 enhances leukemia‐initiating cell self‐
Madan, V., Han, L., Hattori, N., Teoh, W. W., Mayakonda, A., Sun, Q. Y., … renewal to promote leukemia development. Leukemia, 33(6),
Koeffler, H. P. (2018). ASXL2 regulates hematopoiesis in mice and its 1535–1539. https://doi.org/10.1038/s41375‐019‐0377‐0
deficiency promotes myeloid expansion. Haematologica, 103(12), Nagase, R., Inoue, D., Pastore, A., Fujino, T., Hou, H. ‐A., Yamasaki, N., &
1980–1990. https://doi.org/10.3324/haematol.2018.189928 Sera, Y. (2018). Expression of mutant Asxl1 perturbs hematopoiesis
Madzo, J., Liu, H., Rodriguez, A., Vasanthakumar, A., Sundaravel, S., and promotes susceptibility to leukemic transformation. Journal of
Caces, D. B. D., … Godley, L. A. (2014). Hydroxymethylation at gene Experimental Medicine, 215(6), 1729–1747.
regulatory regions directs stem/early progenitor cell commitment Nakajima, H., & Kunimoto, H. (2014). TET2 as an epigenetic master
during erythropoiesis. Cell Reports, 6(1), 231–244. https://doi.org/10. regulator for normal and malignant hematopoiesis. Cancer Prevention
1016/j.celrep.2013.11.044 Research, 105(9), 1093–1099. https://doi.org/10.1111/cas.12484
Maethner, E., Garcia‐Cuellar, M. P., Breitinger, C., Takacova, S., Divoky, V., Neri, F., Incarnato, D., Krepelova, A., Dettori, D., Rapelli, S., Maldotti, M., &
Hess, J. L., & Slany, R. K. (2013). MLL‐ENL inhibits polycomb Oliviero, S. (2015). TET1 is controlled by pluripotency‐associated
repressive complex 1 to achieve efficient transformation of factors in ESCs and downmodulated by PRC2 in differentiated cells
hematopoietic cells. Cell Reports, 3(5), 1553–1566. https://doi.org/ and tissues. Nucleic Acids Research, 43(14), 6814–6826. https://doi.
10.1016/j.celrep.2013.03.038 org/10.1093/nar/gkv392
Mahfoudhi, E., Talhaoui, I., Cabagnols, X., Della Valle, V., Secardin, L., North, T. E., Stacy, T., Matheny, C. J., Speck, N. A., & de Bruijn, M. F.
Rameau, P., … Plo, I. (2016). TET2‐mediated 5‐hydroxymethylcytosine (2004). Runx1 is expressed in adult mouse hematopoietic stem cells
18 | SHOKOUHIAN ET AL.

and differentiating myeloid and lymphoid cells, but not in maturing Biomedicine and Pharmacotherapy, 98, 626–635. https://doi.org/10.
erythroid cells. Stem Cells, 22(2), 158–168. https://doi.org/10.1634/ 1016/j.biopha.2017.12.059
stemcells.22‐2‐158 Saia, M., Termanini, A., Rizzi, N., Mazza, M., Barbieri, E., Valli, D., &
Oak, J. S., & Ohgami, R. S. (2017). Focusing on frequent ASXL1 mutations Alcalay, M. (2016). AML1/ETO accelerates cell migration and impairs
in myeloid neoplasms, and considering rarer ASXL2 and ASXL3 cell‐to‐cell adhesion and homing of hematopoietic stem/progenitor
mutations. Current Medical Research and Opinion, 33(4), 781–782. cells. Scientific Reports, 6, 34957. https://doi.org/10.1038/srep34957
https://doi.org/10.1080/03007995.2017.1284049 Saika, M., Inoue, D., Nagase, R., Sato, N., Tsuchiya, A., Yabushita, T., &
Ohlsson, E., Schuster, M. B., Hasemann, M., & Porse, B. T. (2016). The Goyama, S. (2018). ASXL1 and SETBP1 mutations promote
multifaceted functions of C/EBPalpha in normal and malignant leukaemogenesis by repressing TGFbeta pathway genes through
haematopoiesis. Leukemia, 30(4), 767–775. https://doi.org/10.1038/ histone deacetylation. Scientific Reports, 8(1):15873. https://doi.org/
leu.2015.324 10.1038/s41598‐018‐33881‐2
Pan, F., Weeks, O., Yang, F. C., & Xu, M. (2015). The TET2 interactors and Sato, H., Wheat, J. C., Steidl, U., & Ito, K. (2016). DNMT3A and TET2 in the
their links to hematological malignancies. IUBMB Life, 67(6), 438–445. pre‐leukemic phase of hematopoietic disorders. Frontiers in Oncology,
https://doi.org/10.1002/iub.1389 6, 187. https://doi.org/10.3389/fonc.2016.00187
Pan, F., Wingo, T. S., Zhao, Z., Gao, R., Makishima, H., Qu, G., & Xu, M. Schepers, H., van Gosliga, D., Wierenga, A. T., Eggen, B. J., Schuringa, J. J.,
(2017). TET2 loss leads to hypermutagenicity in haematopoietic stem/ & Vellenga, E. (2007). STAT5 is required for long‐term maintenance of
progenitor cells. Nature Communications, 8, 15102. https://doi.org/10. normal and leukemic human stem/progenitor cells. Blood, 110(8),
1038/ncomms15102 2880–2888. https://doi.org/10.1182/blood‐2006‐08‐039073
Perry, J. M., He, X. C., Sugimura, R., Grindley, J. C., Haug, J. S., Ding, S., & Schindler, J. W., Van Buren, D., Foudi, A., Krejci, O., Qin, J., Orkin, S. H., &
Li, L. (2011). Cooperation between both Wnt/{beta}‐catenin and Hock, H. (2009). TEL‐AML1 corrupts hematopoietic stem cells to
PTEN/PI3K/Akt signaling promotes primitive hematopoietic stem cell persist in the bone marrow and initiate leukemia. Cell Stem Cell, 5(1),
self‐renewal and expansion. Genes and Development, 25(18), 43–53. https://doi.org/10.1016/j.stem.2009.04.019
1928–1942. https://doi.org/10.1101/gad.17421911 Serrano, M., Hannon, G. J., & Beach, D. (1993). A new regulatory motif in
Porse, B. T., Bryder, D., Theilgaard‐Monch, K., Hasemann, M. S., cell‐cycle control causing specific inhibition of cyclin D/CDK4. Nature,
Anderson, K., Damgaard, I., & Nerlov, C. (2005). Loss of C/EBP alpha 366(6456), 704–707. https://doi.org/10.1038/366704a0
cell cycle control increases myeloid progenitor proliferation and Sharma, S., & Gurudutta, G. (2016). Epigenetic regulation of
transforms the neutrophil granulocyte lineage. Journal of Experimental hematopoietic stem cells. International Journal of Stem Cells, 9(1),
Medicine, 202(1), 85–96. https://doi.org/10.1084/jem.20050067 36–43. https://doi.org/10.15283/ijsc.2016.9.1.36
Porter, S. N., Cluster, A. S., Signer, R. A. J., Voigtmann, J., Monlish, D. A., Shen, S., Passioura, T., Symonds, G., & Dolnikov, A. (2007). N‐RAS
Schuettpelz, L. G., & Magee, J. A. (2016). PTEN cell autonomously oncogene‐induced gene expression in human hematopoietic
modulates the hematopoietic stem cell response to inflammatory progenitor cells: Upregulation of p16INK4a and p21CIP1/WAF1
cytokines. Stem Cell Reports, 6(6), 806–814. https://doi.org/10.1016/j. correlates with myeloid differentiation. Experimental Hematology,
stemcr.2016.04.008 35(6), 908–919. https://doi.org/10.1016/j.exphem.2007.02.011
Pronier, E., Almire, C., Mokrani, H., Vasanthakumar, A., Simon, A., da Costa Shen, S. W., Dolnikov, A., Passioura, T., Millington, M., Wotherspoon, S.,
Reis Monte Mor, B., & Delhommeau, F. (2011). Inhibition of Rice, A., & Symonds, G. (2004). Mutant N‐RAS preferentially drives
TET2‐mediated conversion of 5‐methylcytosine to human CD34+ hematopoietic progenitor cells into myeloid
5‐hydroxymethylcytosine disturbs erythroid and granulomonocytic differentiation and proliferation both in vitro and in the NOD/SCID
differentiation of human hematopoietic progenitors. Blood, 118(9), mouse. Experimental Hematology, 32(9), 852–860. https://doi.org/10.
2551–2555. https://doi.org/10.1182/blood‐2010‐12‐324707 1016/j.exphem.2004.06.001
Pronier, E., & Delhommeau, F. (2012). Role of TET2 mutations in Shi, H., Yamamoto, S., Sheng, M., Bai, J., Zhang, P., Chen, R., & Yang, F. C.
myeloproliferative neoplasms. Current Hematologic Malignancy (2016). ASXL1 plays an important role in erythropoiesis. Scientific
Reports, 7(1), 57–64. https://doi.org/10.1007/s11899‐011‐0108‐8 Reports, 6, 28789. https://doi.org/10.1038/srep28789
Qu, X., Zhang, S., Wang, S., Wang, Y., Li, W., Huang, Y., & An, X. (2018). Sood, R., Kamikubo, Y., & Liu, P. (2017). Role of RUNX1 in hematological
TET2 deficiency leads to stem cell factor‐dependent clonal expansion malignancies. Blood, 129(15), 2070–2082.
of dysfunctional erythroid progenitors. Blood, 132(22), 2406–2417. Suh, H. C., Gooya, J., Renn, K., Friedman, A. D., Johnson, P. F., & Keller, J. R.
https://doi.org/10.1182/blood‐2018‐05‐853291 (2006). C/EBPalpha determines hematopoietic cell fate in
Quintana‐Bustamante, O., Lan‐Lan Smith, S., Griessinger, E., Reyal, Y., multipotential progenitor cells by inhibiting erythroid differentiation
Vargaftig, J., Lister, T. A., & Bonnet, D. (2012). Overexpression of wild‐ and inducing myeloid differentiation. Blood, 107(11), 4308–4316.
type or mutants forms of CEBPA alter normal human hematopoiesis. https://doi.org/10.1182/blood‐2005‐06‐2216
Leukemia, 26(7), 1537–1546. https://doi.org/10.1038/leu.2012.38 Sui, X., Price, C., Li, Z., & Chen, J. (2012). Crosstalk between DNA and
Quivoron, C., Couronne, L., Della Valle, V., Lopez, C. K., Plo, I., Wagner‐ histones: TET's new role in embryonic stem cells. Current Genomics,
Ballon, O., & Bernard, O. A. (2011). TET2 inactivation results in 13(8), 603–608. https://doi.org/10.2174/138920212803759730
pleiotropic hematopoietic abnormalities in mouse and is a recurrent Suzuki, T., Shimizu, Y., Furuhata, E., Maeda, S., Kishima, M., Nishimura, H.,
event during human lymphomagenesis. Cancer Cell, 20(1), 25–38. & Suzuki, H. (2017). RUNX1 regulates site specificity of DNA
https://doi.org/10.1016/j.ccr.2011.06.003 demethylation by recruitment of DNA demethylation machineries in
Reddy, V. A., Iwama, A., Iotzova, G., Schulz, M., Elsasser, A., Vangala, R. K., & hematopoietic cells. Blood Advances, 1(20), 1699–1711. https://doi.
Behre, G. (2002). Granulocyte inducer C/EBPalpha inactivates the myeloid org/10.1182/bloodadvances.2017005710
master regulator PU.1: Possible role in lineage commitment decisions. Tan, Y., Liu, H., & Chen, S. (2015). Mutant DNA methylation regulators
Blood, 100(2), 483–490. https://doi.org/10.1182/blood.v100.2.483 endow hematopoietic stem cells with the preleukemic stem cell
Sabnis, A. J., Cheung, L. S., Dail, M., Kang, H. C., Santaguida, M., property, a requisite of leukemia initiation and relapse. Frontiers in
Hermiston, M. L., & Braun, B. S. (2009). Oncogenic KRAS initiates Medicine, 9(4), 412–420. https://doi.org/10.1007/s11684‐015‐0423‐x
leukemia in hematopoietic stem cells. PLoS Biology, 7(3), e59. https:// Tesio, M., Oser, G. M., Baccelli, I., Blanco‐Bose, W., Wu, H., Gothert, J. R., &
doi.org/10.1371/journal.pbio.1000059 Trumpp, A. (2013). PTEN loss in the bone marrow leads to G‐CSF‐
Safaei, S., Baradaran, B., Hagh, M. F., Alivand, M. R., Talebi, M., Gharibi, T., mediated HSC mobilization. Journal of Experimetnal Medicine, 210(11),
& Solali, S. (2018). Double sword role of EZH2 in leukemia. 2337–2349. https://doi.org/10.1084/jem.20122768
SHOKOUHIAN ET AL. | 19

Tonks, A., Tonks, A. J., Pearn, L., Pearce, L., Hoy, T., Couzens, S., & levels. Molecular and Cellular Biology, 28(21), 6668–6680. https://doi.
Darley, R. L. (2004). Expression of AML1‐ETO in human org/10.1128/mcb.01025‐08
myelomonocytic cells selectively inhibits granulocytic differentiation Wu, X., Bekker‐Jensen, I. H., Christensen, J., Rasmussen, K. D., Sidoli, S.,
and promotes their self‐renewal. Leukemia, 18(7), 1238–1245. https:// Qi, Y., & Helin, K. (2015). Tumor suppressor ASXL1 is essential for the
doi.org/10.1038/sj.leu.2403396 activation of INK4B expression in response to oncogene activity and
Trowbridge, J. J., & Orkin, S. H. (2011). DNMT3a silences hematopoietic anti‐proliferative signals. Cell Research, 25(11), 1205–1218. https://
stem cell self‐renewal. Nature Genetics, 44(1), 13–14. https://doi.org/ doi.org/10.1038/cr.2015.121
10.1038/ng.1043 Yan, H., Wang, Y., Qu, X., Li, J., Hale, J., Huang, Y., & An, X. (2017). Distinct
Trowbridge, J. J., Snow, J. W., Kim, J., & Orkin, S. H. (2009). DNA roles for TET family proteins in regulating human erythropoiesis.
methyltransferase 1 is essential for and uniquely regulates Blood, 129(14), 2002–2012. https://doi.org/10.1182/blood‐2016‐08‐
hematopoietic stem and progenitor cells. Cell Stem Cell, 5(4), 736587
442–449. https://doi.org/10.1016/j.stem.2009.08.016 Yang, H., Kurtenbach, S., Guo, Y., Lohse, I., Durante, M. A., Li, J., & Yang, F.
Tsuzuki, S., Hong, D., Gupta, R., Matsuo, K., Seto, M., & Enver, T. (2007). C. (2018). Gain of function of ASXL1 truncating protein in the
Isoform‐specific potentiation of stem and progenitor cell engraftment pathogenesis of myeloid malignancies. Blood, 131(3), 328–341.
by AML1/RUNX1. PLoS Medicine, 4(5), e172. https://doi.org/10.1371/ https://doi.org/10.1182/blood‐2017‐06‐789669
journal.pmed.0040172 Yang, L., Rodriguez, B., Mayle, A., Park, H. J., Lin, X., Luo, M., & Goodell, M.
Tsuzuki, S., & Seto, M. (2012). Expansion of functionally defined mouse A. (2016). DNMT3A loss drives enhancer hypomethylation in FLT3‐
hematopoietic stem and progenitor cells by a short isoform of ITD‐associated leukemias. Cancer Cell, 29(6), 922–934. https://doi.
RUNX1/AML1. Blood, 119(3), 727–735. https://doi.org/10.1182/ org/10.1016/j.ccell.2016.05.003
blood‐2011‐06‐362277 Ye, M., Zhang, H., Amabile, G., Yang, H., Staber, P. B., Zhang, P., &
Uni, M., Masamoto, Y., Sato, T., Kamikubo, Y., Arai, S., Hara, E., & Tenen, D. G. (2013). C/EBPa controls acquisition and maintenance of
Kurokawa, M. (2019). Modeling ASXL1 mutation revealed impaired adult haematopoietic stem cell quiescence. Nature Cell Biology, 15(4),
hematopoiesis caused by derepression of p16Ink4a through aberrant 385–394. https://doi.org/10.1038/ncb2698
PRC1‐mediated histone modification. Leukemia, 33(1), 191–204. Yeh, J. R., Munson, K. M., Chao, Y. L., Peterson, Q. P., Macrae, C. A., &
https://doi.org/10.1038/s41375‐018‐0198‐6 Peterson, R. T. (2008). AML1‐ETO reprograms hematopoietic cell fate
Van Meter, M. E., Diaz‐Flores, E., Archard, J. A., Passegue, E., Irish, J. M., by downregulating SCL expression. Development, 135(2), 401–410.
Kotecha, N., & Braun, B. S. (2007). K‐RasG12D expression induces https://doi.org/10.1242/dev.008904
hyperproliferation and aberrant signaling in primary hematopoietic Yilmaz, O. H., Valdez, R., Theisen, B. K., Guo, W., Ferguson, D. O., Wu, H., &
stem/progenitor cells. Blood, 109(9), 3945–3952. https://doi.org/10. Morrison, S. J. (2006). PTEN dependence distinguishes
1182/blood‐2006‐09‐047530 haematopoietic stem cells from leukaemia‐initiating cells. Nature,
Wang, J., Kong, G., Liu, Y., Du, J., Chang, Y.‐I., Tey, S. R., & Meloche, S. 441(7092), 475–482. https://doi.org/10.1038/nature04703
(2013). NrasG12D/+ promotes leukemogenesis by aberrantly Yokoyama, A. (2015). Molecular mechanisms of MLL‐associated leukemia.
regulating hematopoietic stem cell functions. Blood, 121(26), International Journal of Hematology, 101(4), 352–361.
5203–5207. Yokoyama, A. (2017). Transcriptional activation by MLL fusion proteins in
Wang, J., Li, Z., He, Y., Pan, F., Chen, S., Rhodes, S., & Yang, F. C. (2014). Loss leukemogenesis. Experimental Hematology, 46, 21–30. https://doi.org/
of Asxl1 leads to myelodysplastic syndrome‐like disease in mice. Blood, 10.1016/j.exphem.2016.10.014
123(4), 541–553. https://doi.org/10.1182/blood‐2013‐05‐500272 Yu, M., Mazor, T., Huang, H., Huang, H. T., Kathrein, K. L., Woo, A. J., &
Wang, J., Liu, Y., Li, Z., Wang, Z., Tan, L. X., Ryu, M. J., & Zhang, J. (2011). Cantor, A. B. (2012). Direct recruitment of polycomb repressive
Endogenous oncogenic NRAS mutation initiates hematopoietic complex 1 to chromatin by core binding transcription factors.
malignancies in a dose‐ and cell type‐dependent manner. Blood, Molecular Cell, 45(3), 330–343. https://doi.org/10.1016/j.molcel.
118(2), 368–379. https://doi.org/10.1182/blood‐2010‐12‐326058 2011.11.032
Wang, T., Li, C., Xia, C., Dong, Y., Yang, D., Geng, Y., & Wang, J. (2015). Yuzugullu, H., Baitsch, L., Von, T., Steiner, A., Tong, H., Ni, J., & Zhao, J. J.
Oncogenic NRAS hyper‐activates multiple pathways in human cord (2015). A PI3K p110beta‐RAC signalling loop mediates PTEN‐loss‐
blood stem/progenitor cells and promotes myelomonocytic induced perturbation of haematopoiesis and leukaemogenesis. Nature
proliferation in vivo. American Journal of Translational Research, 7(10), Communications, 6, 8501. https://doi.org/10.1038/ncomms9501
1963–1973. Zhang, J., Grindley, J. C., Yin, T., Jayasinghe, S., He, X. C., Ross, J. T., & Li, L.
Wang, Z., & Bunting, K. D. (2013). STAT5 in hematopoietic stem cell (2006). PTEN maintains haematopoietic stem cells and acts in lineage
biology and transplantation. JAK‐STAT, 2(4):e27159. https://doi.org/ choice and leukaemia prevention. Nature, 441(7092), 518–522.
10.4161/jkst.27159 https://doi.org/10.1038/nature04747
Wang, Z., Li, G., Tse, W., & Bunting, K. D. (2009). Conditional deletion of Zhang, J., Wang, J., Liu, Y., Sidik, H., Young, K. H., Lodish, H. F., &
STAT5 in adult mouse hematopoietic stem cells causes loss of Fleming, M. D. (2009). Oncogenic Kras‐induced leukemogeneis:
quiescence and permits efficient nonablative stem cell replacement. Hematopoietic stem cells as the initial target and lineage‐specific
Blood, 113(20), 4856–4865. https://doi.org/10.1182/blood‐2008‐09‐ progenitors as the potential targets for final leukemic transformation.
181107 Blood, 113(6), 1304–1314. https://doi.org/10.1182/blood‐2008‐01‐
Wang, Z., Medrzycki, M., Bunting, S. T., & Bunting, K. D. (2015). STAT5‐ 134262
deficient hematopoiesis is permissive for Myc‐induced B‐cell Zhang, P., Chen, Z., Li, R., Guo, Y., Shi, H., Bai, J., & Yang, F. C. (2018). Loss
leukemogenesis. Oncotarget, 6(30), 28961–28972. https://doi.org/10. of ASXL1 in the bone marrow niche dysregulates hematopoietic stem
18632/oncotarget.5009 and progenitor cell fates. Cell Discovery, 4, 4. https://doi.org/10.1038/
Wierenga, A. T., Schepers, H., Moore, M. A., Vellenga, E., & Schuringa, J. J. s41421‐017‐0004‐z
(2006). STAT5‐induced self‐renewal and impaired myelopoiesis of Zhang, P., He, F., Bai, J., Yamamoto, S., Chen, S., Zhang, L., & Yang, F. C.
human hematopoietic stem/progenitor cells involves down‐ (2018). Chromatin regulator ASXL1 loss and NF1 haploinsufficiency
modulation of C/EBPalpha. Blood, 107(11), 4326–4333. https://doi. cooperate to accelerate myeloid malignancy. Journal of Clinical
org/10.1182/blood‐2005‐11‐4608 Investigation, 128(12), 5383–5398. https://doi.org/10.1172/jci121366
Wierenga, A. T., Vellenga, E., & Schuringa, J. J. (2008). Maximal STAT5‐ Zhang, P., Iwasaki‐Arai, J., Iwasaki, H., Fenyus, M. L., Dayaram, T.,
induced proliferation and self‐renewal at intermediate STAT5 activity Owens, B. M., & Tenen, D. G. (2004). Enhancement of hematopoietic
20 | SHOKOUHIAN ET AL.

stem cell repopulating capacity and self‐renewal in the absence of the Zhao, Z., Chen, S., Zhu, X., Pan, F., Li, R., Zhou, Y., & Xu, M. (2016). The
transcription factor C/EBP alpha. Immunity, 21(6), 853–863. https:// catalytic activity of TET2 is essential for its myeloid malignancy‐
doi.org/10.1016/j.immuni.2004.11.006 suppressive function in hematopoietic stem/progenitor cells.
Zhang, X., Su, J., Jeong, M., Ko, M., Huang, Y., Park, H. J., & Goodell, M. A. Leukemia, 30(8), 1784–1788. https://doi.org/10.1038/leu.2016.56
(2016). DNMT3A and TET2 compete and cooperate to repress lineage Zjablovskaja, P., Kardosova, M., Danek, P., Angelisova, P., Benoukraf, T.,
‐specific transcription factors in hematopoietic stem cells. Nature Wurm, A. A., & Alberich‐Jorda, M. (2017). EVI2B is a C/EBPalpha
Genetics, 48(9), 1014–1023. https://doi.org/10.1038/ng.3610 target gene required for granulocytic differentiation and functionality
Zhang, Y., Wang, J., Wheat, J., Chen, X., Jin, S., Sadrzadeh, H., & Yeh, J. R. of hematopoietic progenitors. Cell Death and Differentiation, 24(4),
(2013). AML1‐ETO mediates hematopoietic self‐renewal and 705–716. https://doi.org/10.1038/cdd.2017.6
leukemogenesis through a COX/beta‐catenin signaling pathway. Blood,
121(24), 4906–4916. https://doi.org/10.1182/blood‐2012‐08‐447763
Zhang, Y., Yan, X., Sashida, G., Zhao, X., Rao, Y., Goyama, S., & Huang, G.
(2012). Stress hematopoiesis reveals abnormal control of self‐renewal, How to cite this article: Shokuhian M, Bagheri M, Poopak B,
lineage bias, and myeloid differentiation in Mll partial tandem duplication Chegeni R, Davari N, Saki N. Altering chromatin methylation
(Mll‐PTD) hematopoietic stem/progenitor cells. Blood, 120(5),
patterns and the transcriptional network involved in
1118–1129. https://doi.org/10.1182/blood‐2012‐02‐412379
Zhao, X., Chen, A., Yan, X., Zhang, Y., He, F., Hayashi, Y., & Huang, G. regulation of hematopoietic stem cell fate. J Cell Physiol.
(2014). Downregulation of RUNX1/CBFbeta by MLL fusion proteins 2020;1–20. https://doi.org/10.1002/jcp.29642
enhances hematopoietic stem cell self‐renewal. Blood, 123(11),
1729–1738. https://doi.org/10.1182/blood‐2013‐03‐489575

Das könnte Ihnen auch gefallen