You are on page 1of 15

REVIEWS

The genetics of myelodysplastic


syndrome: from clonal haematopoiesis
to secondary leukaemia
Adam S. Sperling1,2, Christopher J. Gibson1,2 and Benjamin L. Ebert1,2
Abstract | Myelodysplastic syndrome (MDS) is a clonal disease that arises from the expansion of
mutated haematopoietic stem cells. In a spectrum of myeloid disorders ranging from clonal
haematopoiesis of indeterminate potential (CHIP) to secondary acute myeloid leukaemia (sAML),
MDS is distinguished by the presence of peripheral blood cytopenias, dysplastic haematopoietic
differentiation and the absence of features that define acute leukaemia. More than 50 recurrently
mutated genes are involved in the pathogenesis of MDS, including genes that encode proteins
involved in pre-mRNA splicing, epigenetic regulation and transcription. In this Review we discuss
the molecular processes that lead to CHIP and further clonal evolution to MDS and sAML. We also
highlight the ways in which these insights are shaping the clinical management of MDS, including
classification schemata, prognostic scoring systems and therapeutic approaches.

Cytopenias
Myelodysplastic syndrome (MDS) is driven by a com- is probably underestimated, as a bone marrow biopsy is
Decreased blood counts of any plex combination of genetic mutations that result in required for diagnosis and many older patients with mild
kind (white cells, red cells or heterogeneity in both clinical phenotype and disease out- cytopenias do not undergo marrow evaluation. The lim-
platelets). come, as is the case for most cancers. The World Health ited data available suggest that the incidence is increasing
Blasts
Organization (WHO) classifies MDS as a clonal disease over time, which may be due to an ageing population,
Immature, hypofunctional characterized by morphological dysplasia, ineffec­tive improving survival after treatments for other neoplasms
leukaemic cells found in the haematopoiesis leading to cytopenias and risk of trans- that place patients at risk of subsequent development of
peripheral blood or bone formation to acute myeloid leukaemia (AML)1. It is now therapy-related MDS (t‑MDS) and increased awareness
marrow.
appreciated that most of the clinical and pathological fea- of MDS among general practitioners5.
tures of MDS are the direct result of recurrent acquired Although patients with MDS can be asymptomatic
somatic genetic lesions (FIG. 1). Although MDS and related at diagnosis, identified only by the incidental discov-
myeloproliferative neoplasms (MPNs) are defined by dis- ery of cytopenias, many patients present with clinical
tinct clinical and morphological criteria, they share many symptoms such as fatigue, often related to anaemia,
of the same genetic mutations, and the composite geno- bleeding owing to thrombocytopenia and fevers or
type of individual cases — including the specific genes recurrent infections as a result of neutropenia. The clin-
mutated, the order in which they were mutated and the ical course is variable: some patients live for many years
interactions between clones and subclones — is likely to with minimal therapy, whereas others rapidly progress
1
Department of Medical underlie these phenotypic differences (BOX 1). In addition, to AML. Morbidity and mortality in MDS are related
Oncology, Dana-Farber there are several rare familial syndromes associated with primarily to complications arising from cytopenias and
Cancer Institute and
Brigham and Women’s
a predisposition to early onset MDS2 (BOX 2). transformation to AML. The risk of either event can be
Hospital, Boston, MDS is among the most common of the haematolog- assessed using one of several prognostic systems, includ-
Massachusetts 02115, USA. ical malignancies, with current estimates placing its inci- ing the International Prognostic Scoring System (IPSS),
2
Division of Hematology, dence in the United States as between 5.3 and 13.1 cases the revised IPSS (IPSS‑R), the WHO-based Prognostic
Department of Medicine,
per 100,000 persons3. In adults, advanced age is the pre- Scoring System (WPSS) and the MD Anderson Com­
Brigham and Women’s
Hospital, Boston, dominant risk factor for developing MDS, with a median prehensive Scoring System (MDA-CSS)6–8. Although the
Massachusetts 02115, USA. age at diagnosis of 71–76 years4,5. Precise enumeration IPSS and, more recently, the IPSS‑R have become the
Correspondence to B.L.E. of the incidence of MDS has been challenging because most widely used, these scoring systems are largely inter-
bebert@partners.org it was not independently recorded in the National changeable and all take into account some combination
doi:10.1038/nrc.2016.112 Cancer Institute’s Survey, Epidemiology and End Results of the degree of cytopenia, the proportion of bone mar-
Published online 11 Nov 2016 (SEER) cancer databases until 2001; the true prevalence row blasts and the karyotype. These prognostic scoring

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 1


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Frequency Frequency Frequency Frequency Frequency


Category Gene in CHIP (#) in MDS (%) in sAML (%) in AML (%) in AA (%) Morphological and clinical features
Epigenetic TET2 • Monocytosis
ASXL1 • Improved response to HMAs
DNMT3A • Decreased response to HMAs
EZH2 • In AML, associated with antecedent MDS
BCOR In AML, associated with antecedent MDS
IDH2
• In AML, associated with antecedent MDS
IDH1
• Poor prognosis
BCORL1
ATRX Poor prognosis
Cohesins STAG2 In AML, associated with antecedent MDS
CTCF
SMC3
SMC1A
RAD21
Splicing SF3B1 • Ring sideroblasts
SRSF2 • Good prognosis
U2AF1 • Associated with sAML
ZRSR2 • Often seen in CMML
LUC7L2 • Associated with sAML
U2AF2 Poor prognosis
PRPF8
SF1 In AML, associated with antecedent MDS

Transcription RUNX1 • Thrombocytopenia


factors CUX1 • Familial platelet disorder
ETV6 • Poor prognosis
NPM1 Poor prognosis
GATA2
MonoMAC syndrome
WT1
CEBPA Poor prognosis in MDS

p53 TP53 • Associated with complex karyotype


PPM1D ND ND ND ND • Poor prognosis, espeically with alloHSCT

Signalling JAK2 • Thrombocytosis


NF1 • Associated with MPN
NRAS Increased risk of progression to AML
CBL Associated with JMML
MPL Associated with MPN
KRAS Increased risk of progression to AML
FLT3 • Poor prognosis
GNAS • Increased risk of progression to AML
KIT Associated with mast cell disorders
PTPN11
GNB1
Frequency
Others BRCC3
ATM High Low
FANCL
ND No data No mutations found
PIGA

Figure 1 | Recurrent mutations in CHIP and MDS. Mutations are sorted by their frequency in myelodysplastic
Nature Reviews | Cancer
syndrome (MDS) into functional categories. Mutation percentages (%) shown for all categories except clonal
haematopoiesis of indeterminate potential (CHIP), for which the absolute mutation count is shown (#). Data shown are
derived from the following sources: CHIP12,21, MDS9,25,26,35, secondary acute myeloid leukaemia (sAML) 28, acute myeloid
leukaemia (AML)192–194 and aplastic anaemia (AA)45. alloHSCT, allogeneic haematopoietic stem cell transplantation;
ASXL1, additional sex combs-like 1; ATM, ataxia telangiectasia mutated; BCOR, BCL6 co-repressor; BCORL1, BCL6
co-repressor-like 1; BRCC3, BRCA1/BRCA2‑containing complex subunit 3; CEBPA, CCAAT/enhancer binding protein α;
CMML, chronic myelomonocytic leukaemia; CTCF, CCCTC-binding factor; CUX1, cut-like homeobox 1; DNMT3A, DNA
methyltransferase 3A; ETV6, ETS variant 6; EZH2, enhancer of zeste 2; FANCL, Fanconi anaemia complementation group L;
FLT3, fms related tyrosine kinase 3; GATA2, GATA binding protein 2; GNB1, G protein subunit β1; HMAs, hypomethylating
agents; IDH1, isocitrate dehydrogenase 1; JAK2, Janus kinase 2; JMML, juvenile myelomonocytic leukaemia; LUC7L2, LUC7
like 2, pre-mRNA splicing factor; MonoMAC, monocytosis with increased susceptibility to mycobacterial infections;
MPL, gene that encodes the thrombopoietin receptor; MPN, myeloproliferative neoplasm; NF1, nuclear factor 1;
NPM1, nucleophosmin; PIGA, phosphatidylinositol glycan anchor biosynthesis class A; PPM1D, protein phosphatase, Mg2+/
Mn2+-dependent 1D; PRPF8, pre-mRNA processing factor 8; PRPN11, protein tyrosine phosphatase, non-receptor type 11;
RUNX1, runt related transcription factor 1; SF1, splicing factor 1; SF3B1, splicing factor 3b subunit 1; SMC1A, structural
maintenance of chromosomes 1A; SRSF2, serine and arginine rich splicing factor 2; STAG2, stromal antigen 2; TET2, tet
methylcytosine dioxygenase2; U2AF1, U2 small nuclear RNA auxiliary factor 1; WT1, Wilms tumour 1; ZRSR2, zinc finger
CCCH-type, RNA binding motif and serine/arginine rich 2.

2 | ADVANCE ONLINE PUBLICATION www.nature.com/nrc


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

systems do not include information about somatic muta- overlap or a common early precursor stem cell12. Many
Secondary AML
(sAML). Acute myeloid tions in individual genes, although this genetic infor­ of the mutations commonly found in myeloid malig-
leukaemia (AML) that arises mation can predict prognosis independently of each of nancies, such as tet methylcytosine dioxygenase 2
out of a pre-existing myeloid the prognostic scoring systems9,10. (TET2) and splicing factor 3b subunit 1 (SF3B1), have
neoplasm such as In this Review we discuss the molecular and genetic also been identified in lymphoid cells or have been
myelodysplastic syndrome
(MDS) or myeloproliferative
basis of MDS, beginning with initiating mutations in described in various lymphomas13–17.
neoplasms. Distinguished from haematopoietic stem cells (HSCs) that lead to the devel- Although most patients with MDS follow a course
MDS by the presence of 20% opment of clonal haematopoiesis. This initiating event dominated by cytopenias and their consequences,
or more blasts in the bone is followed by the accumulation of additional cooperat- approximately one-third progress to high-risk MDS
marrow or peripheral blood.
ing mutations and eventual progression to overt clinical and sAML18. The genetic mutations in MDS seem to
Lineage disease, including MDS and secondary AML (sAML). be initiated in an HSC. Sequencing of sequential sam-
A collection of cell surface Finally, we discuss how our evolving understanding of ples from individual patients demonstrated that in the
markers that defines mature the genetics of MDS provides insights into the clinical 5q− ­syndrome clinical stability was associated with muta-
blood cells, including B cells, course, prognosis and treatment of patients with MDS. tional stability 11. Conversely, those patients who pro-
T cells, monocytes,
granulocytes and red
gressed to AML developed multiple new mutations in
blood cells. The molecular genetics of MDS the leukaemic stem cell compartment, coupled with new
Cell of origin. A central challenge in understanding myeloid progenitor (that is, non-HSC) populations that
5q− syndrome the development of MDS, as in other malignancies, had gained self-renewal potential11. The expansion of
Myelodysplastic syndrome
has been identifying the cell of origin. Recent work has self-renewal activity outside the stem cell compartment,
associated with isolated
deletion of chromosome 5q demonstrated that distinct MDS stem cells bearing the and expansion of the population of cells with proliferative
and characterized by a immunophenotype of normal HSCs (Lineagelow, CD34+, potential, are two other key steps in the transition from
macrocytic anaemia, normal or CD38–, CD90+ (also known as THY1+), CD45RA– (also MDS to AML. Eradicating the mutated MDS stem cells,
elevated platelet count, a low known as PTPRC–)) are able to sustain the generation of which are capable of maintaining the disease indefinitely,
marrow blast count and a
relatively indolent course.
myeloid progenitors in vitro and in vivo, whereas other is likely to be essential to curing the disease.
early myeloid progenitors are unable to do so11.
Paroxysmal nocturnal Although the founding genetic event in MDS patho- Clonal haematopoiesis: initiators of disease. The pres-
haemoglobinuria genesis has long been assumed to occur in a myeloid-­ ence of initiating mutations leading to clonal expansion,
(PNH). Caused by mutations in
biased HSC, patients with clonal haematopoiesis have and thus a pre-malignant state, has long been suspected
the phosphatidylinositol glycan
anchor biosynthesis class A an increased risk of developing both lymphoid and to precede the development of most malignancies
(PIGA) gene leading to the loss myeloid malignancies, perhaps suggesting mutational (FIG. 2). Initial studies of healthy women demonstrated
of glycosylphosphatidylinositol
(GPI), a chemical linker that
functions to anchor several Box 1 | The spectrum of myeloid neoplasms
proteins to blood cell
membranes including those Myelodysplastic syndrome (MDS) is a haematological disorder within the larger spectrum of myeloid neoplasms. Other
that block complement-­ myeloid neoplasms include the myeloproliferative neoplasms (MPNs), acute myeloid leukaemia (AML) and MDS/MPN
mediated haemolysis. overlap syndromes. Although these were previously felt to be biologically distinct entities, it is now appreciated that
there is a considerable degree of genetic overlap between them.
Aplastic anaemia
MPNs include chronic myeloid leukaemia (CML), polycythemia vera (PV), essential thrombocythemia (ET), myelofibrosis
(AA). Pancytopenia in the
setting of aplastic bone marrow
(MF), chronic neutrophilic leukaemia (CNL) and systemic mastocytosis (SM)1. Each has specific pathological
caused by immune-mediated characteristics, but all are distinguished from MDS by the absence of morphological dysplasia and normal, or more often
destruction of haematopoietic increased, production of blood cells. All MPNs have mutations that constitutively activate signalling cascades and
progenitors. It is often difficult cytokine-independent proliferation. These include mutations in Janus kinase 2 (JAK2) (PV, ET and MF), calreticulin (CALR)
to distinguish morphologically (ET and MF), thrombopoietin receptor (MPL) (ET and MF), colony stimulating factor 3 receptor (CSF3R) (CNL), KIT (SM) and
from hypoplastic BCR–ABL rearrangements (CML)115–119,163.
myelodysplastic syndrome. Some myeloid neoplasms display features of both MPNs (elevated peripheral blood cell counts or bone marrow fibrosis)
and MDS (morphological dysplasia) and are termed MDS/MPN overlap syndromes. These include chronic myelomonocytic
Hypoplastic MDS
leukaemia (CMML), which typically presents with peripheral blood monocytosis and dysplasia in the bone marrow; atypical
Myelodysplastic syndrome
(MDS) with low bone marrow
CML, which clinically resembles CML but lacks a BCR–ABL rearrangement; and unclassifiable MDS/MPN overlap
cellularity. Hypoplastic MDS syndromes1,164. Overlap syndromes frequently harbour concomitant MDS-associated mutations (for example, serine and
can be difficult to distinguish arginine rich splicing factor 2 (SRSF2) and additional sex combs-like 1 (ASXL1)) and MPN-associated mutations
from aplastic anaemia owing to (for example, JAK2). They may also harbour mutations not specific for any disorder (for example, tet methylcytosine
the small number of cells dioxygenase 2 (TET2) and DNA methyltransferase 3A (DNMT3A)), as well as frequent mutations activating the RAS pathway
available to evaluate for (for example, NRAS, KRAS, CBL and protein tyrosine phosphatase, non-receptor type 11 (PTPN11))165.
morphological dysplasia. Although not classified as myeloid neoplasms, paroxysmal nocturnal haemoglobinuria (PNH) and aplastic anaemia (AA)
can overlap with and transform into MDS and AML. PNH is defined by clonal, somatic mutations in phosphatidylinositol
glycan anchor biosynthesis class A (PIGA). In AA, a subset of patients have clonal mutations in genes specifically
associated with myeloid malignancy, including DNMT3A, ASXL1, JAK2 and TP53 (which encodes p53), at a relative
frequency very similar to that seen in clonal haematopoiesis of indeterminate potential (CHIP)45,166,167. In addition,
mutations in PIGA and BCL6 co-repressor (BCOR) are more common in AA than in CHIP and are associated with a better
response to immunosuppressive therapy and improved overall survival45. Mutations in CHIP and myeloid malignancy
genes, such as DNMT3A, ASXL1 or TP53, were associated with worse overall survival45. It is unclear how specific clonal
somatic aberrations affect disease pathogenesis and evolution in AA and PNH and whether they can be used clinically to
help differentiate AA from the morphologically related hypoplastic MDS.

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 3


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Box 2 | Inherited bone marrow failure and MDS


Although most cases of myelodysplastic syndrome (MDS) are caused by somatic mutations, a few are due to inherited
mutations that lead to increased risk of bone marrow failure1,2. These typically affect patients in adolescence or early
adulthood and include the following disorders.
• Diamond–Blackfan anaemia: characterized by absent erythroid precursors due to mutations in ribosomal protein
genes, causing abnormal ribosome biogenesis, and rarely MDS or AML168.
• Telomerase complex disorders: due to mutations in genes encoding members of the telomerase complex (dyskerin
pseudouridine synthase 1 (DKC1), telomerase reverse transcriptase (TERT), telomerase RNA component (TERC) and
GAR1 ribonucleoprotein (GAR1)169–172. The best known is dyskeratosis congenita (DKC), characterized by bone marrow
failure, pulmonary fibrosis, oral leukoplakia, dystrophic nails, hyperpigmentation and cirrhosis.
• Familial platelet disorder with propensity to myeloid malignancy: characterized by childhood thrombocytopenia and a
propensity to progress to MDS. It is caused by mutations in runt related transcription factor 1 (RUNX1), which encodes
a component of the core binding factor transcriptional activator complex101.
• Familial thrombocytopenia and malignancy: characterized by dominantly inherited thrombocytopenia, an elevated
mean corpuscular volume (MCV), and a propensity to haematological malignancies, including MDS and acute myeloid
leukaemia (AML). It is caused by mutations in ETS variant 6 (ETV6)99,173.
• Fanconi anaemia: an early-onset aplastic anaemia often associated with dysmorphic features, short stature and limb
defects, caused by mutations in any of more than 15 genes involved in DNA repair168.
• Li–Fraumeni syndrome: caused by loss‑of‑function mutations in TP53, leading to an increased predisposition to many
cancers, including MDS and AML174.
• Inherited GATA binding protein 2 (GATA2) mutations: a heterogeneous group of disorders, including MonoMAC
syndrome (monocytosis with increased susceptibility to mycobacterial infections)175 and Emberger syndrome (familial
lymphoedema and deafness)176, both of which also carry a risk of early-onset MDS.
• Inherited DEAD-box helicase 41 (DDX41) mutations: associated with mild monocytosis and an increased risk of MDS and
AML, with a longer latency to disease initiation (average age 62 years) than other genetic syndromes discussed here138,177.
• Shwachman–Diamond syndrome (SDS): characterized by skeletal defects, pancreatic insufficiency and bone marrow
failure, with risk of transformation to MDS. It is caused by mutations in the SBDS gene, which is required for normal
ribosome maturation178.

skewing of X‑chromosome inactivation in almost 40% In fact, the most significant driver of decreased survival
Mean corpuscular volume
of women more than 60 years of age, and a subset of these in association with CHIP was an increased propensity
(MCV). A measure of the women were later found to harbour mutations in TET2, for thrombosis, coronary artery disease and stroke, the
average volume of red blood suggestive of clonal haematopoiesis driven by a somatic reason for which is unclear and remains under active
cells. Increased size is mutation19. More recently, exome sequencing of periph- investigation12. Despite these effects on survival, there
associated with abnormal
eral blood samples from more than 30,000 patients with- are currently no data to suggest that screening of asymp-
or delayed red blood cell
differentiation. out known haematological malignancies demonstrated tomatic patients for the presence of CHIP is clinically
recurrent somatic myeloid malignancy-associated muta- indicated, especially in the absence of an intervention
Clonal haematopoiesis of tions in up to 10% of patients over the age of 65 and more that could restore polyclonal haematopoiesis.
indeterminate potential than 20% of patients over the age of 90 (REFS 12,20,21).
(CHIP). The presence of a
somatic mutation associated
This phenomenon has subsequently been termed clonal From CHIP to MDS: a blurred distinction. In a model
with haematological haematopoiesis of indeterminate p ­ otential (CHIP)22. The of clonal evolution beginning with CHIP and ending
malignancy at a variant allele most common recurrently mutated genes were DNA in frank haematological malignancy, the transition to
fraction of at least 2% and the methyltransferase 3A (DNMT3A), TET2, additional sex MDS probably involves a complex interplay between
absence of morphological
combs-like 1 (ASXL1), TP53 (which encodes p53), Janus epigenetic alterations in the HSC, a dysfunctional bone
evidence of malignancy or
diagnostic criteria for kinase 2 (JAK2) and SF3B1, all of which are also mutated marrow microenvironment (BOX 3) and the stepwise
paroxysmal nocturnal in MDS12,21 (FIG. 1). acquisition of additional driver mutations (FIG. 2). The
haemoglobinuria, monoclonal In these studies, the presence of CHIP was a strong clinical diagnosis of MDS, as currently defined, does
gammopathy of undetermined predictor of the development of subsequent haemato- not incorporate somatic mutations, but is instead based
significance or monoclonal
B‑lymphocytosis.
logical malignancy (hazard ratio of 11.1), with an annual on the morphology of haematopoietic cells in the bone
risk of approximately 0.5–1%, and decreased overall sur- marrow, the finding of cytogenetic abnormalities and
Hazard ratio vival (hazard ratio for all-cause mortality of 1.4) com- the development of cytopenias in the peripheral blood1.
A statistical measure that pared with age-matched controls12. In two patients with As currently defined, a diagnosis of CHIP requires the
corresponds to the probability
CHIP who later developed AML, sequencing of bone presence of a somatic mutation with a mutant allele frac-
of a particular outcome
attributable to a given variable marrow-derived DNA demonstrated that the leukae- tion of at least 2% in the peripheral blood and no other
compared with normal mias were clonally derived from the previously identified evidence of a haematological malignancy 22. In contrast,
controls. CHIP21. Importantly, however, the absolute risk of trans- the presence of a mutation and otherwise unexplained
formation to overt malignancy in patients with CHIP cytopenias or borderline dysplasia is suggestive of, but
Annual risk
The probability of acquiring a
was low during the time periods under study, probably does not confirm, progression to MDS, as the formal
condition over the course of reflecting the need to acquire additional mutations in diagnosis still requires the fulfilment of specific mor-
1 year. a relatively small pool of potential cooperating genes. phological criteria23,24. Approximately 35% of patients

4 | ADVANCE ONLINE PUBLICATION www.nature.com/nrc


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Idiopathic cytopenias of with idiopathic cytopenias of undetermined significance zinc finger CCCH-type, RNA binding motif and serine/­
undetermined significance (ICUS) have somatic mutations characteristic of MDS, arginine rich 2 (ZRSR2)) and epi­genetic modi­fiers, espe-
(ICUS). Cytopenias that remain although whether all such patients go on to develop cially DNMT3A and TET2, tend to be mutated early in the
unexplained after thorough morphological dysplasia, or have a clinical course simi­ evolution of MDS, whereas mutations in transcription fac-
evaluation and do not meet
World Health Organization
lar to that of patients with MDS even in the absence of tors (runt related transcription factor 1 (RUNX1), GATA
criteria for a haematological dysplasia, requires further study 24. Conversely, as the full binding protein 2 (GATA2) and cut like h ­ omeobox 1
neoplasm. complement of mutations involved in MDS has yet to (CUX1)) can be either early or late events26.
be defined, the absence of a known somatic mutation The genes most frequently mutated in CHIP par-
de novo AML
also does not exclude the diagnosis of MDS22. On the tially correspond to the initiating mutations in MDS.
Acute myeloid leukaemia (AML)
that arises without a pre-existing
whole, however, MDS is a more genetically complex dis- The two most commonly mutated genes in CHIP,
myeloid neoplasm or a history ease than CHIP, with most patients harbouring at least DNMT3A and TET2, both encode epigenetic regula-
of cytotoxic therapy. More two, and sometimes many more, somatic mutations in tors that, when mutated in MDS, tend to occur early in
common in younger patients recurrent driver genes, often with a high mutant allele disease pathogenesis26. On the other hand, mutations
and associated with an overall
better prognosis.
fraction (>10%), at the time of diagnosis25,26. in splicing factors are less common in CHIP than would
The genetic lesions that initiate MDS promote self-­ be expected based on their frequency in MDS. This
Induction therapy renewal, leading to a proliferative advantage over normal observation suggests that they may be more morpholog-
High-dose intensive HSCs and asymptomatic clonal expansion and eventu- ically deterministic than mutations in epigenetic regu­
chemotherapy directed at
ally to overt disease (BOX 3). Mutations that occur early lators and that the patients who acquire splicing factor
inducing remission in acute
leukaemias.
in disease evolution can be detected by calculating allele mutations develop overt dysplasia more rapidly and
frequency in bulk sequencing studies, or by single cell are thus relatively under-represented in the cohorts of
Polycomb repressive sequencing, and these two approaches correlate well with ‘healthy’ adults in whom CHIP was defined. Other genes
complex each other 27. Using these methods, several studies have mutated frequently in CHIP cohorts are either mutated
(PRC). Multiprotein complex
involved in epigenetic
demonstrated that, as general groups, the splicing factors rarely (CBL) or not yet assessed (protein phosphatase,
repression of gene (SF3B1, serine and arginine rich splicing factor 2 (SRSF2), Mg 2+/Mn2+-dependent 1D (PPM1D)) in MDS12,21.
transcription. U2 small nuclear RNA auxiliary factor 1 (U2AF1) and
Progression to leukaemia. sAML is a disease distinct
from de  novo AML, and is characterized by a poorer
response to induction therapy, a significantly higher
relapse rate and an overall inferior prognosis. Mounting
evidence suggests that sAML is also biologically distinct
from de novo disease, reflecting its evolution from MDS.
Careful study of rigorously defined cases of de novo AML
and sAML has shown that mutations in SRSF2, SF3B1,
U2AF1, ZRSR2, ASXL1, enhancer of zeste 2 (EZH2),
BCL6 co-repressor (BCOR, part of a polycomb repressive
complex (PRC)) and stromal antigen 2 (STAG2, a com-
ponent of the cohesin complex) are strongly associated
with an antecedent MDS and are thus highly specific
for sAML28. These mutations define a group of patients
with AML that behaves clinically like sAML even in cases
when no pre-existing dysplasia or cytopenias have been
documented. In contrast, mutations in nucleophosmin
(NPM1) and rearrangements involving mixed-lineage
leukaemia 1 (MLL1, also known as KMT2A, located
at chromosome 11q23) or genes that encode compo-
nents of the core binding factor are primarily restricted to
de novo AML. Most other mutations, including those in
Clonal haematopoiesis DNMT3A, TET2 and RUNX1, are not unique to either
Cytopenias and dysplasia disease entity. TP53‑mutated AML comprises its own
Excess blasts and sAML
unique category of disease that tends to have fewer coop-
erating point mutations, which may be functionally repli-
cated by a high frequency of cytogenetic rearrangements
Normal blood cell Mutation 1 Mutation 2 Mutation 3 Mutation 4
that disrupt global chromosomal architecture9,28.
Some of the mutations that occur during progression
Figure 2 | Clonal expansion in MDS. Early mutations tend to leadNature Reviews | Cancer
to increased from MDS to AML are found in core haematopoietic
haematopoietic stem cell (HSC) self-renewal, clonal expansion and the development
transcription factor genes, including RUNX1, GATA2
of clonal haematopoiesis of indeterminate potential (CHIP). As the mutant clone
continues to enlarge, it gives rise to an expanding population of cells in which and CCAAT/enhancer binding protein α (CEBPA),
acquisition of additional genetic or epigenetic lesions can promote progression to overt which abrogate normal differentiation28. Activating
malignancy. These secondary subclonal events tend to lead to the development of mutations in signalling pathway components such as
overt dysplasia, myelodysplastic syndrome (MDS) and eventually secondary acute fms related tyrosine kinase 3 (FLT3) and RAS family
myeloid leukaemia (sAML). members, which control cellular proliferation, also

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 5


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Cohesin complex commonly occur during the progression to sAML, and lethality, such that co‑mutation would have either a neu-
A multisubunit protein their subclonal presence in otherwise lower-risk MDS tral or a negative consequence. For instance, the individ-
complex that forms a ring is associated with impending transformation26,29–31. ual splicing factors are almost never co‑mutated with each
structure capable of encircling other, and neither are cohesin complex genes33,34. Other
two chromosomal strands of
DNA and is required for sister
Positive and negative cooperativity between mutations. combinations of mutations have a relative paucity of co‑­
chromatid cohesion during At the time of diagnosis, most cases of MDS and sAML are occurrence, such as ASXL1 and DNMT3A25,26,35. Definitive
mitosis. clonally and genetically complex, with many clones con- functional evidence of mutual exclusivity does not exist
taining more than three cooperating disease-­associated for most mutation pairs. In the absence of such evidence,
Core binding factor
mutations32. In MDS, as in other cancers, certain muta- it is therefore important not to overstate the biological
A core haematopoietic
transcription factor complex,
tions co‑occur at frequencies greater than or less than basis for these associations.
mutations of which are would be expected by chance. These variances are most As discussed above, some specific mutations are
associated with acute myeloid likely driven by patterns of functional complementarity, associated with the initial development of CHIP,
leukaemia in younger patients redundancy and synthetic lethality. Complementarity is whereas subsequent evolution is probably guided by
and a relatively good prognosis.
most easily appreciated in cases with mutations in genes the temporal acquisition of mutations that cooperate to
Red cell distribution width from different classes of biological function, as for exam- generate overt malignancy. The timing and context of
A measure of the distribution ple, in a patient with TET2, SRSF2 and RUNX1 mutations. each serial mutation may influence disease phenotype
of red blood cell sizes; On the other hand, certain mutations co‑occur much less and progression. This concept has been best described
indicates the degree of
often than would be expected by chance, which presum- in the MPNs, in which mutations in JAK2 and TET2
variation within a sample.
ably implies either functional redundancy or synthetic co‑occur in approximately 10% of patients. Sequencing

Box 3 | Aberrant haematopoiesis, the bone marrow microenvironment and MDS


The somatic mutations that drive myelodysplastic syndrome (MDS) that MDS can be cured in some cases by transplantation of HSCs
increase self-renewal, aberrant differentiation and morphological alone suggests that the stroma is at least not always the primary
abnormalities in haematopoietic stem cells (HSCs). The effects of MDS disease-initiating element.
mutations on HSCs are apparent in some mouse models179. For Like aplastic anaemia, MDS is often associated with immune dysfunction.
example, mutations in or loss of Dnmt3a or Tet2 expand the mutant Inappropriate immune targeting of haematopoietic progenitors may
HSC clone and increase HSC function at the expense of normal contribute to the cytopenias observed in some cases of MDS and is
polyclonal haematopoiesis75,77,180. another example of an abnormal haematopoietic environment
Asymptomatic patients with clonal haematopoiesis of indeterminate contributing to disease phenotype188. Up to 30% of patients with MDS
potential (CHIP), by definition, have normal blood counts, but many show improvements in their cytopenias following treatment with
have an elevated mean red cell distribution width, suggesting some immunosuppressants, and patients with trisomy 8 are particularly likely to
dyserythropoiesis even with a single mutation12. Certain mutations respond to this form of therapy189–191.
are associated with specific abnormalities in
differentiation, such as the formation of ring Normal
sideroblasts in patients with splicing factor 3b subunit 1 Bone marrow Peripheral blood
(SF3B1) mutations and impaired erythropoiesis and Stem cells Progenitor cells
hypolobated ­micromegakaryocytes in patients with
5q syndrome
− 54,181
.
The ineffective haematopoiesis and cytopenias that
define MDS may seem paradoxical, as the genetic
lesions in MDS give HSCs a clonal advantage, and
generally expand progenitor cells, leading to a
hypercellular bone marrow. Terminal maturation,
however, is impaired, leading to increased apoptosis of
differentiating cells and thus peripheral cytopenias
(see the figure).
HSCs exist in a supportive stromal microenvironment
comprised of sinusoidal vascular endothelial cells,
osteoblasts, adipocytes, and other components, which
can also alter HSC function and differentiation182–184. MDS
In mice, selective disruption of certain genes only in
the stroma, such as deletion of β‑catenin (Ctnnb1), the Cytopenias
Dicer1 ribonuclease or Sbds, the gene mutated in Blasts
the Schwachman–Diamond syndrome (BOX 2), can cause
abnormal differentiation of HSCs and the development
of dysplasia185,186. This interaction is bidirectional, as
myeloid neoplasms can remodel the bone marrow niche,
often leading to increased fibrosis187. Although in these
model systems MDS can be induced purely by stromal
defects, it is unclear whether human MDS can be caused Normal cell
by defects isolated to the bone marrow stroma or
whether genetic and epigenetic alterations in the HSCs Dysplasia Dysplastic or
mutant cell
are an essential component of human disease. The fact

Nature Reviews | Cancer

6 | ADVANCE ONLINE PUBLICATION www.nature.com/nrc


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Nascent pre-mRNA AML in which chemotherapy eliminates leukaemic blasts


and leaves bone marrow in which there is no morpholog-
5′ EXON ESE A YYY AG ESE EXON 3′
ical evidence of disease, pre-malignant mutations may
still be detectable and able to reinitiate disease41–43. It is
important to recognize that this post-treatment clonality
Splicing complex formation is an entity distinct from CHIP, carrying a much higher
SF1 risk of progression to overt leukaemia28,41–43.
A
SF3B1 U2AF2
Recurrently mutated pathways
YY
Y
U2
Several studies have used targeted sequencing panels
ZRS
U2A R2 built around genes mutated in other myeloid neoplasms
U6 AG F1
ESE SRS
F2 to describe the spectrum of somatic mutations in MDS
U4
EXON U1 ESE and their clinical and pathological consequences9,25,26,44.
5′ U5
EXO
N
Between 75% and 90% of patients have a mutation in
3′ at least one known, recurrently mutated gene, and this
number will probably increase as additional genes are
identified25,26 (FIG. 1).
Mature mRNA Lariat
U2 Mutational processes. MDS arises as a consequence
U5 of the sequential acquisition of somatic mutations in
5′ EXON ESE ESE EXON 3′
U6 HSCs. It is estimated that HSCs acquire approximately
0.13 exonic single nucleotide variants (SNVs) per year
Figure 3 | Splicing factor mutations in myeloid neoplasms. Shown is a simplified of life32. This calculation assumes a constant mutational
schematic of the steps involved in mRNA splicing. Splicing factor mutations cluster rate, but it is possible that initiating mutations alter the
Nature Reviews | Cancer
in components of the 3′ spliceosome. Proteins encoded by genes mutated in mutational milieu of the cell, thus shaping the subse-
myelodysplastic syndrome (MDS) are denoted in red. A, branch site; AG, splice acceptor quent evolutionary path26. The most frequent base-pair
site; ESE, exonic splicing enhancer; SF1, splicing factor 1; SF3B1, splicing factor 3b change seen in clonal haematopoiesis and MDS is the
subunit 1; SRSF2, serine and arginine rich splicing factor 2; U2AF1, U2 small nuclear RNA C>T transition, a hallmark of methylcytidine deami-
auxiliary factor 1; YYY, polypyrimidine tract; ZRSR2, zinc finger CCCH-type, RNA binding
nation associated with ageing 12,45–47. Other age-related
motif and serine/arginine rich 2. Adapted with permission from REF. 195, Elsevier.
mutational processes also contribute to the pathogenesis
of MDS, including large chromosome rearrangements,
of single cell clones isolated from patients with MPN the occurrence of small insertions and deletions and
showed that those with a JAK2 mutation occurring first progressive telomeric shortening 48.
are more likely to develop polycythemia vera and have
increased thrombotic risk, whereas those with a TET2 Splicing factors. Alternative splicing of pre-mRNA is a
mutation occurring first are more likely to develop common feature of eukaryotic genes and is one of the
essential thrombocythemia36. JAK2 mutations frequently most commonly dysregulated processes in cancer 49.
co‑occur with SF3B1 mutations in the MDS/MPN over- Mutations in components of the spliceosome are the
lap syndrome (BOX 1) MDS/MPN with ring sideroblasts most common recurrent lesions in MDS and are found in
Polycythemia vera and thrombocytosis (MDS/MPN‑RS‑T)1. As in MPNs, up to 60% of cases50. The majority are in components of
(PV). Myeloproliferative
neoplasm characterized by an
either mutation can occur first, and the original pheno- the 3′ spliceosome, including SF3B1, SRSF2, U2AF1 and,
elevated red blood cell count, type can subsequently be modified by acquisition of the to a lesser extent ZRSR2, pre-mRNA processing factor 8
and almost exclusively second mutation37,38. (PRPF8), U2AF2, LUC7 like 2, pre-mRNA splicing factor
associated with activating (LUC7L2) and splicing factor 1 (SF1)10,25,26,51–53 (FIG. 3).
mutations in Janus kinase 2
Evolution in response to therapy. Although MDS Mutations in SF3B1, SRSF2 and U2AF1 occur exclu-
(JAK2).
undergoes clonal evolution even in untreated patients, sively as heterozygous missense mutations at defined
Essential thrombocythemia the administration of disease-altering therapy, such hot spots, leading to highly recurrent amino acid sub-
(ET). Myeloproliferative as hypomethylating agents or lenalidomide, functions as stitutions that alter the function of the splicing machin-
neoplasm characterized by an external selective pressure that can influence the rela­ ery. SF3B1 mutations, the most common spliceo­some
elevated platelet count,
associated with mutations in
tive proportions of coexisting clones. To date, there are alteration in MDS, are highly associated with the
Janus kinase 2 (JAK2), limited genetic data for how MDS evolves with treatment, presence of ring sideroblasts and a relatively benign
calreticulin (CALR) or MPL and much of our understanding is extrapolated from prognosis38,52. They lead to altered selection of 3ʹ splice
(gene that encodes the related diseases. Analysis of serial AML samples collected sites and aberrant splicing of important genes involved
thrombopoietin receptor).
at diagnosis and relapse shows that subclones apparent at in iron homeostasis that probably mediate the ring
Ring sideroblasts diagnosis can become undetectable, whereas mutations sideroblast phenotype54,55. SRSF2 is the second most
Early red cell precursors conferring resistance to therapy may evolve over time commonly mutated splicing factor in MDS and is also
containing aberrant through either the acquisition of resistance mutations frequently mutated in MDS/MPN overlap syndromes.
mitochondrial iron staining; in the dominant subclone or outgrowth of a pre-existing Missense mutations alter SRSF2’s binding to exonic
associated with mutations in
splicing factors, most
subclone39,40. In some cases, these resistant subclones can splice enhancers, which in turn leads to mis-splicing
commonly splicing factor 3b be detected even in samples from patients in remission of several important genes, including EZH2 (REF. 56).
subunit 1 (SF3B1). and provide a source for leukaemic relapse40–42. In cases of Mutations in U2AF1 occur in 10–15% of MDS cases,

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 7


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Active Repressed IDH1 Derepressed Active Repressed


and IDH2 H3K27 H3K27me3
DNMT3A TET2 EZH2
C 5mC 5hmC C
Promoter
CTCF BCOR H2AK119Ub

SM
TF (RUNX1,

C3
ETV6 or WT1)

C1
SM
ASXL1

A
CTCF TF
RAD21
Enhancer
STAG2 DNA

Figure 4 | Multiple steps in gene expression are recurrently disrupted in MDS. Shown is a prototypical gene promoter
Nature Reviews | Cancer
with chromosomal looping facilitated by CCCTC-binding factor (CTCF) and the cohesin complex, enabling transcription
factors (TFs) bound at distant enhancers to interact with the promoter. Alterations in epigenetic marks such as DNA
methylation and histone post-translational modifications function to regulate the transcription of genes. Green signifies
loss-of-function (or dominant-negative function) mutations. Red signifies gain-of-function mutations. ASXL1, additional
sex combs-like 1; BCOR, BCL6 co-repressor; C, cytosine; 5mC, 5‑methylcytosine; 5hmC, 5‑hydroxymethylcytosine;
DNMT3A, DNA methyltransferase 3A; ETV6, ETS variant 6; EZH2, enhancer of zeste 2; H2AK119Ub, histone H2A lysine 119
ubiquitylation; H3K27, histone H3 lysine 27; H3K27me3, H3K27 trimethylation; IDH, isocitrate dehydrogenase; RUNX1,
runt related transcription factor 1; SMC1A, structural maintenance of chromosomes 1A; STAG2, stromal antigen 2; TET2,
tet methylcytosine dioxygenase2; WT1, Wilms tumour 1.

are not associated with any particular morphological Alterations in DNA methylation also occur in
phenotype and promote increased exon skipping 51,57,58. patients with mutations in the genes encoding the iso­
Neither SRSF2 nor U2AF1 mutations confer the favour- citrate dehydrogenase (IDH) enzymes, IDH1 and IDH2,
able prognosis associated with mutations in SF3B1 found in approximately 5% of patients with MDS.
(REF. 38). Although these studies have shown that muta- These mutations produce a neomorphic enzyme that
tions in different splicing factor genes lead to distinct converts isocitrate into R-2‑hydroxyglutarate (2HG)
patterns of aberrant splicing and alter the abundance or instead of α‑ketoglutarate (α‑KG)25,70,71. 2HG acts as
function of independent sets of target genes, no specific an onco­metabolite and diffuses to the nucleus, where
alternatively spliced isoform has been demonstrated to it promotes neoplasia by, among other things, inhib-
directly cause disease. Furthermore, although the data iting α‑KG‑dependent ­dioxygenases, including TET2
are still limited, there has been little overlap of altered (REFS 72–74) (FIG. 4).
splicing events reported between mice and humans or The mechanisms by which changes in methylation
between individual spliceosomal mutants, raising the contribute to the pathogenesis of MDS are complex, and
possibility that mutations in these genes could promote multiple studies have been unable to identify a clear cor-
MDS through some alternative mechanism. relation between methylation and gene expression60,75.
Although DNMT3A and TET2 are seemingly biochem-
Epigenetic regulators: DNA methylation and histone ical opposites, their genes are frequently co‑mutated in
modification. Post-translational modifications of DNA MDS25. Mice deficient in either Dnmt3a or Tet2 have
and histones are important mechanisms of cellular epi- pheno­typic similarities and the double mutants show
genetic regulation. Mutations in genes involved in these accelerated development of malignancy75–77. Interrogation
processes are the second most common set of recurrent of the global landscape of 5mC and 5hmC in haemato-
lesions in MDS. poietic cells has shown that many genes known to be dys-
Methylation of cytosines in repetitive CpG elements regulated in myeloid malignancies lie within ‘canyons’ of
in DNA, mediated by the DNA methyltransferases sparse methylation, where their expression is regulated
(DNMTs), is one of the most common epigenetic modifi- by epigenetic histone modifications. DNMT3A is impor-
cations, and functions by altering the accessibility of DNA tant for maintaining the borders of these canyons, but
regulatory regions59 (FIG. 4). Inactivating mutations in the those same borders are also enriched in 5hmC, suggest-
gene encoding one such enzyme, DNMT3A, occur in ing synergistic, rather than divergent, roles for TET2 and
10–15% of MDS cases25,26,60,61. The opposing process, DNMT3A in this capacity 76,78.
DNA demethylation, is mediated by the ten-eleven Genes that encode histone modifying enzymes also
translocation (TET) family of proteins, dioxygenases that contribute to MDS. The covalent modification of his-
catalyse the conversion of 5‑methylcytosine (5mC) into tone tails leads to changes in chromatin structure and
5‑hydroxymethylcytosine (5hmC) as part of a multi­step altered binding of regulatory proteins (FIG. 4). The PRCs
reaction that eventually leads to DNA demethyl­ation62–65 are two distinct protein complexes (PRC1 and PRC2)
(FIG. 4). TET2 is one of the most commonly altered genes that are both required for maintaining the transcrip-
in MDS, with inactivating mutations found in approx- tional silencing of key developmental regulators dur-
imately 30% of cases25,59,66. These mutations are associ- ing differentiation. PRC2 trimethylates histone H3 on
ated with hypermethylation of cytosines at enhancer lysine 27 (H3K27me3), whereas PRC1 ubiquitylates
sequences and subsequent repression of several genes histone H2A at lysine 119; both alterations lead to chro-
important for myeloid differentiation67–69. matin compaction79–81. Components of both complexes

8 | ADVANCE ONLINE PUBLICATION www.nature.com/nrc


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

can be mutated in MDS. EZH2 encodes a PRC2 catalytic in MDS. For example, Wilms tumour 1 (WT1) encodes
subunit and is mutated in approximately 5% of patients a sequence-specific DNA-binding transcription factor
with MDS; loss of Ezh2 promotes the development of mutated in fewer than 5% of cases of MDS that functions
MDS in mouse models25,82. BCOR and BCOR-like 1 to recruit TET2 to specific genomic loci103–105.
(BCORL1) are components of a PRC1 complex known
as PRC1.1, are mutated in approximately 5% of cases The role of p53 in MDS. TP53 is the most frequently
of MDS and are associated with poor prognosis25,83–86. mutated tumour suppressor gene across all human can-
ASXL1 is recurrently mutated in approximately 20% cers and is recurrently mutated in MDS49 (see FIG. 1).
of patients with MDS25. Although not itself a constitu- Humans born with a single mutant allele of TP53, the
ent of either PRC, ASXL1 forms a polycomb repressive Li–Fraumeni syndrome, have a dramatically increased
deubquity­lase complex with BRCA1‑associated pro- risk of many types of cancer, including MDS and AML
tein 1 (BAP1) that physically interacts with PRC2 and (BOX 2). Somatic disruption of TP53 in MDS is strongly
deubiquitylates histone H2A87,88. Pathogenic ASXL1 associated with low platelet levels, a high blast count,
mutations are restricted to exons 11 and 12, and lead to a complex karyotype and previous exposure to chemo-
truncated protein product that increases the deubiquity­ therapy 9,28. Many patients with deletion of one TP53
lation activity of BAP1, which is associated with allele, including cases with del(17p), carry a second inac-
decreased global H3K27me3 (REFS 88,89). Mutations in tivating mutation in the other allele of TP53 (REF. 106).
EZH2, ASXL1 and BCOR all lead to dysregulation of sev- Loss of TP53 in MDS and AML carries a particularly
eral important haematopoietic lineage genes, including dismal prognosis9,107.
the homeobox A (HOXA) cluster, possibly explaining p53 mediates the response to cellular stress by increas-
their role in promoting dysplasia and cytopenias82,88,90. ing expression of genes involved in apoptosis and cell
cycle arrest 108,109. This pathway is negatively regulated
Cohesin complex. The cohesins (STAG2, structural by the phosphatase PPM1D. Truncating mutations in
maintenance of chromosomes 3 (SMC3), SMC1A and PPM1D have been identified in CHIP and are found at
RAD21) form a ring-shaped multiprotein structure increased frequency in the blood of patients with ovarian
that encircles DNA and helps to maintain sister chro- cancer who have previously been treated with chemo-
matid cohesion, which in turn prevents collapse of the therapy 21,110. Inappropriate entry into the cell cycle before
replication fork and facilitates homologous recombina- DNA repair is complete and inadequate activation of
tion-mediated DNA repair. Loss‑of‑function mutations the DNA damage repair machinery are believed to con-
in cohesin genes occur in approximately 15% of MDS tribute to the chemotherapy resistance, accumulation of
cases25,26,34,91. Despite the role of these proteins in sister genomic alterations and chromosomal instability seen in
chromatid cohesion, cohesin mutations in MDS are not patients with inactivation of the p53 pathway.
associated with aneuploidy or chromosomal aberra- Although loss of p53 in vitro can promote aber-
tions91. Cohesins also function to stabilize DNA loops rant self-renewal in some assays, HSCs with hetero­
that promote interaction between promoters and distant zygous inactivation of Trp53 do not have an advantage
enhancers92 (FIG. 4). It is now thought that cohesin muta- over normal HSCs in competitive mouse transplant
tions drive MDS pathogenesis primarily through dysreg- models111,112. However, if those mice are then exposed
ulation of long-range chromatin interactions, leading to to alkylating agents or ionizing radiation, the Trp53
altered gene expression, rather than through their roles mutant clone rapidly expands at the expense of normal
in replication and homologous recombination, although HSCs111,113. Indeed, mutations in TP53 are present in
further work is needed to confirm this hypothesis93–96. approximately 5% of MDS cases, but more than 30% of
therapy-­related myeloid neoplasms25,28,44,111,114. Of note,
Transcription factors. A small number of core haemato- TP53 and PPM1D are among the genes most frequently
poietic transcription factors are recurrently mutated in found to be mutated in individuals with CHIP, raising
MDS. Germline loss‑of‑function mutations in RUNX1, the possibility that detection of somatic mutations in
GATA2 and ETS variant 6 (ETV6) are associated with patients scheduled to receive chemotherapy for other
inherited bone marrow failure disorders that carry a cancers could identify those most at risk of developing
risk of MDS and AML97–99 (BOX 2). RUNX1 is the DNA- therapy-related myeloid neoplasms12,21.
binding subunit of the core binding factor, which reg-
ulates several genes involved in haemato­p oiesis. In Abnormal cell signalling. Mutations in signalling path-
addition to germline mutations, somatic mutations are way components are associated with pro-proliferative
found in approximately 10% of cases of MDS, often asso- states and occur in a range of myeloid malignancies,
ciated with severe thrombocytopenia9,25,100,101. RUNX1 including AML (FLT3), polycythemia vera (JAK2),
mutations co‑occur with cohesin mutations, as well essential thrombocythemia (JAK2 and the gene encod-
as in other genetic contexts26,91. GATA2 encodes a zinc ing the thrombopoietin receptor (MPL)), chronic
finger transcription factor that is highly expressed in myelomonocytic leukaemia (CMML) (CBL) and mast
HSCs, and is essential for normal haematopoietic dif- cell disorders (KIT)115–119. Mutations in these genes all
ferentiation. Like RUNX1, both germline and somatic occur at a relatively low frequency in MDS compared
mutations occur in GATA2, but somatic mutations are with AML, CMML or MPN. Many of these mutations
present in only 1–2% of patients with MDS25,102. Several affect the cell through activation of the MAPK path-
other transcription factors are mutated less commonly way, as well as other signalling pathways. Mutations in

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 9


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

the MAPK pathway (NRAS, KRAS, neurofibromin 1 predictive of a poor prognosis126,127. Elucidation of the
(NF1) and protein tyrosine phosphatase, non-receptor pathogenic genes within large chromosomal deletions
type 11 (PTPN11)) are those most frequently found has been challenging. Among those that are most com-
in MDS but still occur in only approximately 10% of mon and best understood are isolated deletion of 5q, loss
cases overall25,26. When such mutations do occur they of chromosome 7 and deletion of 17p (discussed above
are typically found in subclonal populations and occur with TP53).
late in disease evolution, often heralding the transi-
tion to sAML30,31. Most signalling pathway mutations Chromosome 5q deletions. The single most common
are missense mutations or involve small insertions or isolated cytogenetic abnormality in MDS is deletion
deletions that lead to constitutive activation. One excep- of chromosome 5q. Patients with MDS with isolated
tion to this is the CBL gene, which encodes a tyrosine deletions of chromosome 5q often have a consistent
kinase-associated ubiquitin ligase120. CBL mutations clinical pheno­type, termed the 5q− syndrome, which is
lead to upregulation of signalling proteins such as FLT3 more common in women and has a relatively indolent
and MPL121,122. Although CBL mutations are relatively course128,129. Deletion of 5q leads to haploinsufficiency
common in CHIP, they occur less frequently in MDS, of a small number of genes, including ribosomal protein
perhaps owing to tropism for myelomonocytic lineages, S14 (RPS14), casein kinase 1 α1 (CSNK1A1), adenoma-
resulting instead in enrichment of MDS/MPN overlap tous polyposis coli (APC), heat shock protein family A
syndromes123. When CBL mutations do occur in MDS, (HSP70) member 9 (HSPA9), early growth response 1
they are often late events12,21,26,118. (EGR1), DEAD-box helicase 41 (DDX41), NPM1,
TRAF-interacting protein with forkhead-­associated
Recurrent cytogenetic rearrangements domain  B (TIFAB), Diaphanous-related formin 1
Cytogenetic analysis of bone marrow samples from (DIAPH1), microRNA (miR)-145 and miR‑146a, most
patients with MDS is part of routine clinical practice, and of which lack evidence of bi-allelic inactivation130–136.
large chromosomal rearrangements are seen in approx- A small subset of cases also have point mutations in
imately half of cases (TABLE 1, reviewed in detail else- CSNK1A1 or DDX41, but no other genes have been
where 124). Like individual gene mutations, these identified with bi‑allelic deletion or mutation137–139.
large-scale copy-number alterations can serve as found- Targeted short hairpin RNA (shRNA) screening of the
ing events and drive disease evolution in MDS and genes in the commonly deleted region in 5q− syndrome
AML (reviewed in detail elsewhere125). Acquisition of demonstrated that loss of RPS14 leads to a block in
cytogenetic abnormalities during disease evolution is pre-­ribosomal RNA (rRNA) processing and abnormal

Table 1 | Recurrent cytogenetic abnormalities in MDS


Chromosomal Key genes deleted* IPSS‑R risk Clinical features
abnormality category6
Normal NA Good NA
del(5q) CSNK1A1, RPS14, EGR1, Good Sensitive to lenalidomide196
APC, DDX41, HSPA9, NPM1,
TIFAB, DIAPH1, miR‑145 and
miR‑146a130–140
Monosomy 7 or del(7q) EZH2, MLL3 and CUX1 Poor Monosomy 7 may have a worse
(REFS 148–150) prognosis than del(7q)197
Trisomy 8 Unknown Intermediate High response rate to
immunosuppression190
Trisomy 19 Unknown Intermediate Unknown
del(20q) MYBL2, TP53RK and TP53TG5 Good Often associated with mutations in
(REF. 198) splicing factors198
del(17p) TP53 (REF. 109) N/A Poor response to alloHSCT35.
Complex and monosomal
‡ §
TP53 (REF. 109) Poor to very Associated with TP53 mutation35
poor
del(11q) MLL and ATM199 Very good Unknown
Y chromosome loss (–Y) Unknown Very good May not be pathogenic, but
instead may be lost during
normal ageing200
alloHSCT, allogeneic haematopoietic stem cell transplantation; APC, adenomatous polyposis coli; ATM, ataxia telangiectasia
mutated; del, deletion; CSNK1A1, casein kinase 1 α1; CUX1, cut like homeobox 1; DDX41, DEAD-box helicase 41; DIAPH1,
Diaphanous-related formin 1; EGR1, early growth response 1; EZH2, enhancer of zeste 2; HSPA9, heat shock protein family A
(HSP70) member 9; IPSS‑R, revised International Prognostic Scoring System; MDS, myelodysplastic syndrome; miR, microRNA;
MLL, mixed lineage leukaemia; MYBL2, v‑myb avian myeloblastosis viral oncogene homologue-like 2; NA, not applicable; N/A, not
included in the IPSS‑R; RPS14, ribosomal protein S14; TP53RK, TP53 regulating kinase; TP53TG5, TP53‑target gene 5; TIFAB,
TRAF-interacting protein with forkhead-associated domain B.*Other genes have been implicated in some studies; ‡complex: ≥3
abnormalities; §monosomal: >2 monosomies.

10 | ADVANCE ONLINE PUBLICATION www.nature.com/nrc


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Chromosome 7 deletions. Deletion of 7q and/or mon-


Ub Ub
Ub osomy 7 are also common in MDS and are associated
Ub
CK1α with a poor prognosis6. Like deletion of 5q, these chro-
CK1α CK1α
mosomal aberrations lead to haploinsufficiency of sev-
LEN LEN
eral genes implicated in haematological malignancies,
CRBN CRBN CRBN including EZH2, CUX1 and MLL3 (also known as
KMT2C), which encodes a histone H3 lysine 4 (H3K4)
methyltransferase148–150. In mouse models, Mll3 haplo­
b CK1α insufficiency cooperates with mutations in the RAS
level
­pathway and Trp53 to promote leukaemia148.

Genetics and MDS management


The standard backbone of therapy for MDS involves
supportive measures such as transfusion, infection
control, growth factor support and iron chelation. The
use of disease-altering therapy, including lenalidomide,
5‑­a zacitidine, decitabine or allogeneic HSC trans­
+LEN Time
CSNK1A1 +/– +/+ +/– +/+ +/– +/+ plantation (alloHSCT), is based on careful evaluation
of the individual patient, including age, prognostic risk
Figure 5 | Mechanism of lenalidomide efficacy in 5q− syndrome. a | Lenalidomide
Nature Reviews
(LEN) functions through modulation of the substrate binding specificity | Cancer
of cereblon stratification and genetics18.
(CRBN), an E3 ubiquitin ligase, for casein kinase 1 α (CK1α, encoded by CSNK1A1). In the
absence of LEN, CRBN has low affinity for CK1α. Binding of LEN, however, induces a Genetic predictors of response to therapy. The hypo-
conformational change in CRBN that significantly increases this affinity, thereby methylating agent 5‑azacitidine and its derivative
catalysing efficient ubquitylation (Ub) and degradation of CK1α. b | Haematopoietic 5‑aza‑2ʹ‑­deoxycytidine (decitabine) inhibit DNMTs
stem cells (HSCs) harbouring 5q deletion (5q− syndrome), which lack one copy of and thus lead to global hypomethylation. Both drugs
CSNK1A1 and have a lower CK1α level, have a clonal advantage over wild-type cells at are active in MDS, and 5‑azacitidine improves overall
baseline. Treatment with LEN selectively depletes CK1α in all HSCs; in 5q− cells this survival when compared with standard supportive care
pushes the CK1α level below a critical threshold and triggers cell death, whereas
in patients at higher risk as defined by the IPSS151,152.
wild-type CSNK1A1 cells retain enough CK1α for survival and can eventually repopulate
Molecular analysis may help to predict which patients
the bone marrow.
will benefit the most from these therapies. Loss of TET2
across the entire MDS clone predicts improved response
erythroid differentiation that phenocopies the macro­ to hypomethylating agents, which may be related to the
cytosis and anaemia seen in the 5q − syndrome 140. finding that these patients often have low-risk disease.
Mouse models with conditional heterozygous inactiva­ The presence of additional mutations, such as in ASXL1,
tion of Rps14 display a similar erythroid defect that is are associated with a poorer prognosis, and subclonal
p53 dependent and accompanied by upregulation of TET2 mutations do not similarly predict for response
components of innate immune signalling 141,142. This is to treatment 103,153.
a similar mechanism to that proposed for Diamond– Isolated deletion of chromosome 5q predicts
Blackfan anaemia, caused by germline heterozygous response to treatment with lenalidomide, which leads to
inactivation of ribosomal protein genes143 (BOX 2). An complete cytogenetic remission in more than half of
abundance of evi­dence now indicates that the full clin- patients and a reduced need for transfusions in 75%
ical phenotype of 5q− MDS is caused by combinatorial of those treated146. As discussed above, lenalidomide
­haploinsufficiency of multiple factors133,144. functions via selective degradation of CK1α, leading to
Another important gene involved in the pathogenesis acti­vation of p53. TP53‑mutant subclones can often be
of 5q− syndrome is CSNK1A1, which encodes a serine/ detected in pretreatment samples and expand following
threonine kinase that when heterozygously deleted leads treatment with lenalidomide, leading to the develop-
to upregulation of WNT signalling and stem cell expan- ment of resistance154. This may also explain the lack of
sion139,145. Homozygous deletion of CSNK1A1, on the efficacy of lenalidomide in patients with 5q deletion in
other hand, leads to the accumulation of p53 and loss the setting of complex karyotypes or in AML, as many
of the competitive advantage seen in the heterozygous of these patients are likely to carry concomitant muta-
setting. It is this dose-dependent effect of CSNK1A1 that tions in TP53. The molecular correlates of response
leads to one of the hallmark features of 5q− syndrome: its to lenalidomide in patients with non‑5q− syndrome
responsiveness to treatment with the thalidomide deriv- (approximately 25% become transfusion independent
ative lenalidomide146 (FIG. 5). Lenalidomide binds to the in this context) are not known155.
Conditioning regimens substrate recognition component (cereblon, CRBN) of
High-dose preparative the Cullin-RING E3 ubiquitin ligase CRL4, altering its Allogeneic bone marrow transplantation. The only
chemotherapy regimens given substrate affinity and leading to the selective degrada- curative treatment for MDS is alloHSCT. Although some
before stem cell tion of the CSNK1A1 gene product, CK1α (REFS 145,147) patients receive myeloablative conditioning regimens,
transplantation. Can be either
(FIG. 5). Loss of CK1α leads to activation of p53 and apop- recent studies showing equivalent success rates with
myeloablative (doses sufficient
to completely ablate the bone tosis, and knockdown of TP53 abrogates the effect of reduced-­intensity regimens suggest that the primary
marrow) or non-myeloablative. lenalidomide in vitro139,145. mechanism of action is immunological elimination of

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 11


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

the MDS stem cell clone156. The efficacy is far from com- neomorphic IDH1 and IDH2 enzymes have shown ini-
plete, however, and specific mutations have been associ- tial promising results and induce differentiation of pri-
ated with poor survival following alloHSCT, including mary AML cells in vitro158,159. IDH mutations also lead
TP53, owing to increased rates of post-­transplant to increased sensitivity to pro-apoptotic BH3 mimetics,
relapse35,157. TP53 mutation is a powerful marker of poor such as the small-molecule BCL‑2 inhibitor veneto-
survival after transplant: in a retro­spective analysis of clax, probably owing to altered mitochondrial function
87 patients, not one of the 18 with a TP53 mutation was related to elevated 2HG160. Initial attempts to target epi-
still alive 5 years after alloHSCT with reduced-intensity genetic regulation with histone deacetylase inhibitors
conditioning regimens35. Interestingly, the rare patients have thus far not shown clinical efficacy 161. MDS cells
with a complex karyotype who lack a concomitant TP53 carrying mutations in splicing factors may be uniquely
mutation have a similar prognosis to patients with a sensitive to further inhibition of splicing via mecha-
­normal karyotype35. nisms probably analogous to haploinsufficiency, and
this observation has led to the development of small-­
Clinical implications of molecular genetics. A more molecule splicing inhibitors162. Clinical trials of these
complete understanding of MDS genetics can inform agents are expected to begin soon.
multiple aspects of our clinical practice, including diag-
nosis, prognosis and prediction of response to therapy. Conclusions
First, although morphological analysis is still required The rapid accumulation of genetic data over the past
to diagnose MDS, it is evident that morphology often decade has provided a molecular taxonomy of MDS, a
directly relates to the underlying genetic lesions. It is guide to the genetic progression of CHIP to MDS and
likely that genetics will have an increasing role in the MDS to sAML, and the identification of core molec-
diagnosis of MDS in coming years as accumulating ular processes that are functionally disrupted through
evidence strengthens the association between specific somatic mutations. CHIP, a pre-malignant condition,
mutations and clinicopathological features of the dis- is common in older adults and is associated with a spe-
ease. It is also clear that certain mutations influence the cific subset of mutations, most commonly in DNMT3A,
prognosis of patients with MDS. Initial attempts to inte- TET2, ASXL1, TP53 and SF3B1, suggesting that muta-
grate mutational data into the IPSS have demonstrated tions in these key genes are important for initiating
improved accuracy in predicting overall survival, and disease. In contrast, mutations in haematopoietic tran-
large-scale international collaborations are now under scription factors and activated signalling pathways tend
way to fully incorporate molecular genetics into the to occur during disease progression to high-risk MDS
next generation of prognostic scoring systems33. and sAML. Patterns of cooperativity and mutual exclu-
As discussed above, specific mutations may predict sivity act to define the evolutionary path from disease
response to standard therapies such as hypomethylat- initiation to leukaemia, and it is now clear that clin-
ing agents, lenalidomide and alloHSCT, and prospec- ical phenotype, prognosis and response to therapy in
tive studies are needed to validate these findings. Finally, MDS are influenced by combinations of genetic lesions
understanding the molecular underpinnings of MDS and the order and context in which they occur. Further
can aid in the development of new targeted therapies. understanding of these relationships will enable refined
Targeting epigenetic modifiers and splicing factors is an prognostic staging models, the identification of patients
attractive option, as mutations in these are often found- most likely to respond to therapy and the development
ing events. For example, small-molecule inhibitors of of new targeted therapeutics.

1. Arber, D. A. et al. The 2016 revision to the World 9. Bejar, R. et al. Clinical effect of point mutations in 15. Nilsson, L. et al. Isolation and characterization of
Health Organization (WHO) classification of myeloid myelodysplastic syndromes. N. Engl. J. Med. 364, hematopoietic progenitor/stem cells in 5q‑deleted
neoplasms and acute leukemia. Blood 127, 2496–2506 (2011). myelodysplastic syndromes: evidence for involvement
2391–2405 (2016). The first paper showing that individual at the hematopoietic stem cell level. Blood 96,
2. Liew, E. & Owen, C. Familial myelodysplastic MDS-associated mutations are associated with 2012–2021 (2000).
syndromes: a review of the literature. Haematologica patient outcomes. 16. Odejide, O. et al. A targeted mutational landscape of
96, 1536–1542 (2011). 10. Bejar, R. et al. Validation of a prognostic model and angioimmunoblastic T‑cell lymphoma. Blood 123,
3. Cogle, C. R. Incidence and burden of the the impact of mutations in patients with lower-risk 1293–1296 (2014).
myelodysplastic syndromes. Curr. Hematol. myelodysplastic syndromes. J. Clin. Oncol. 30, 17. Quivoron, C. et al. TET2 inactivation results in
Malignancy Rep. 10, 272–281 (2015). 3376–3382 (2012). pleiotropic hematopoietic abnormalities in mouse and
4. Ma, X., Does, M., Raza, A. & Mayne, S. T. 11. Woll, P. S. et al. Myelodysplastic syndromes are is a recurrent event during human lymphomagenesis.
Myelodysplastic syndromes: incidence and survival in propagated by rare and distinct human cancer stem Cancer Cell 20, 25–38 (2011).
the United States. Cancer 109, 1536–1542 (2007). cells in vivo. Cancer Cell 25, 794–808 (2014). 18. Steensma, D. P. Myelodysplastic syndromes: diagnosis
5. Sekeres, M. A. Epidemiology, natural history, and This paper describes the most detailed and treatment. Mayo Clin. Proc. 90, 969–983 (2015).
practice patterns of patients with myelodysplastic characterization of MDS stem cells. 19. Busque, L. et al. Recurrent somatic TET2 mutations in
syndromes in 2010. J. Natl Compr. Canc. Netw. 9, 12. Jaiswal, S. et al. Age-related clonal hematopoiesis normal elderly individuals with clonal hematopoiesis.
57–63 (2011). associated with adverse outcomes. N. Engl. J. Med. Nat. Genet. 44, 1179–1181 (2012).
6. Greenberg, P. L. et al. Revised international prognostic 371, 2488–2498 (2014). 20. Xie, M. et al. Age-related mutations associated with
scoring system for myelodysplastic syndromes. Blood Description of CHIP and its association with overall clonal hematopoietic expansion and malignancies.
120, 2454–2465 (2012). and cardiovascular mortality. Nat. Med. 20, 1472–1478 (2014).
7. Kantarjian, H. et al. Proposal for a new risk model in 13. Miyamoto, T., Weissman, I. L. & Akashi, K. 21. Genovese, G. et al. Clonal hematopoiesis and blood-
myelodysplastic syndrome that accounts for events AML1/ETO-expressing nonleukemic stem cells in cancer risk inferred from blood DNA sequence.
not considered in the original International Prognostic acute myelogenous leukemia with 8;21 chromosomal N. Engl. J. Med. 371, 2477–2487 (2014).
Scoring System. Cancer 113, 1351–1361 (2008). translocation. Proc. Natl Acad. Sci. USA 97, 22. Steensma, D. P. et al. Clonal hematopoiesis of
8. Malcovati, L. et al. Impact of the degree of anemia on 7521–7526 (2000). indeterminate potential and its distinction from
the outcome of patients with myelodysplastic 14. Nilsson, L. et al. Involvement and functional myelodysplastic syndromes. Blood 126, 9–16 (2015).
syndrome and its integration into the WHO impairment of the CD34+CD38–Thy‑1+ hematopoietic 23. Cargo, C. A. et al. Targeted sequencing identifies
classification-based Prognostic Scoring System stem cell pool in myelodysplastic syndromes with patients with preclinical MDS at high risk of disease
(WPSS). Haematologica 96, 1433–1440 (2011). trisomy 8. Blood 100, 259–267 (2002). progression. Blood 126, 2362–2365 (2015).

12 | ADVANCE ONLINE PUBLICATION www.nature.com/nrc


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

24. Kwok, B. et al. MDS-associated somatic mutations and 48. Colla, S. et al. Telomere dysfunction drives aberrant 74. Lu, C. et al. IDH mutation impairs histone
clonal hematopoiesis are common in idiopathic hematopoietic differentiation and myelodysplastic demethylation and results in a block to cell
cytopenias of undetermined significance. Blood 126, syndrome. Cancer Cell 27, 644–657 (2015). differentiation. Nature 483, 474–478 (2012).
2355–2361 (2015). 49. Lawrence, M. S. et al. Discovery and saturation References 70–74 demonstrate that IDH
25. Haferlach, T. et al. Landscape of genetic lesions in 944 analysis of cancer genes across 21 tumour types. mutations cause disease through the generation of
patients with myelodysplastic syndromes. Leukemia Nature 505, 495–501 (2014). a novel oncometabolite.
28, 241–247 (2014). 50. Inoue, D., Bradley, R. K. & Abdel-Wahab, O. 75. Challen, G. A. et al. Dnmt3a is essential for
The largest panel of MDS patients sequenced to Spliceosomal gene mutations in myelodysplasia: hematopoietic stem cell differentiation. Nat. Genet.
date. molecular links to clonal abnormalities of 44, 23–31 (2012).
26. Papaemmanuil, E. et al. Clinical and biological hematopoiesis. Genes Dev. 30, 989–1001 (2016). 76. Zhang, X. et al. DNMT3A and TET2 compete and
implications of driver mutations in myelodysplastic 51. Graubert, T. A. et al. Recurrent mutations in the cooperate to repress lineage-specific transcription
syndromes. Blood 122, 3616–3627 (2013). U2AF1 splicing factor in myelodysplastic syndromes. factors in hematopoietic stem cells. Nat. Genet. 48,
Identification of early and late mutations in MDS Nat. Genet. 44, 53–57 (2012). 1014–1023 (2016).
using variant allele fractions. 52. Papaemmanuil, E. et al. Somatic SF3B1 mutation in 77. Moran-Crusio, K. et al. Tet2 loss leads to increased
27. Hughes, A. E. et al. Clonal architecture of secondary myelodysplasia with ring sideroblasts. N. Engl. J. Med. hematopoietic stem cell self-renewal and myeloid
acute myeloid leukemia defined by single-cell 365, 1384–1395 (2011). transformation. Cancer Cell 20, 11–24 (2011).
sequencing. PLoS Genet. 10, e1004462 (2014). 53. Yoshida, K. et al. Frequent pathway mutations of 78. Jeong, M. et al. Large conserved domains of low DNA
28. Lindsley, R. C. et al. Acute myeloid leukemia ontogeny splicing machinery in myelodysplasia. Nature 478, methylation maintained by Dnmt3a. Nat. Genet. 46,
is defined by distinct somatic mutations. Blood 125, 64–69 (2011). 17–23 (2014).
1367–1376 (2015). Initial description of the finding of common splicing 79. Cao, R. et al. Role of histone H3 lysine 27
Characterization of different mutational spectra mutations in MDS. methylation in Polycomb-group silencing. Science
found in de novo AML and sAML. 54. Conte, S. et al. Aberrant splicing of genes involved in 298, 1039–1043 (2002).
29. Corces-Zimmerman, M. R., Hong, W. J., Weissman, I. L., haemoglobin synthesis and impaired terminal 80. Francis, N. J., Kingston, R. E. & Woodcock, C. L.
Medeiros, B. C. & Majeti, R. Preleukemic mutations in erythroid maturation in SF3B1 mutated refractory Chromatin compaction by a polycomb group protein
human acute myeloid leukemia affect epigenetic anaemia with ring sideroblasts. Br. J. Haematol. 171, complex. Science 306, 1574–1577 (2004).
regulators and persist in remission. Proc. Natl Acad. 478–490 (2015). 81. Wang, H. et al. Role of histone H2A ubiquitination
Sci. USA 111, 2548–2553 (2014). 55. Darman, R. B. et al. Cancer-associated SF3B1 hotspot in Polycomb silencing. Nature 431, 873–878 (2004).
30. Murphy, D. M. et al. NRAS mutations with low allele mutations induce cryptic 3’ splice site selection 82. Sashida, G. et al. Ezh2 loss promotes development of
burden have independent prognostic significance for through use of a different branch point. Cell Rep. 13, myelodysplastic syndrome but attenuates its
patients with lower risk myelodysplastic syndromes. 1033–1045 (2015). predisposition to leukaemic transformation. Nat.
Leukemia 27, 2077–2081 (2013). 56. Kim, E. et al. SRSF2 mutations contribute to Commun. 5, 4177 (2014).
31. Takahashi, K. et al. Dynamic acquisition of FLT3 or myelodysplasia by mutant-specific effects on exon 83. Gao, Z. et al. PCGF homologs, CBX proteins, and RYBP
RAS alterations drive a subset of patients with lower recognition. Cancer Cell 27, 617–630 (2015). define functionally distinct PRC1 family complexes.
risk MDS to secondary AML. Leukemia 27, 57. Ilagan, J. O. et al. U2AF1 mutations alter splice site Mol. Cell 45, 344–356 (2012).
2081–2083 (2013). recognition in hematological malignancies. Genome 84. Gearhart, M. D., Corcoran, C. M., Wamstad, J. A. &
32. Welch, J. S. et al. The origin and evolution of Res. 25, 14–26 (2015). Bardwell, V. J. Polycomb group and SCF ubiquitin
mutations in acute myeloid leukemia. Cell 150, 58. Przychodzen, B. et al. Patterns of missplicing due to ligases are found in a novel BCOR complex that is
264–278 (2012). somatic U2AF1 mutations in myeloid neoplasms. recruited to BCL6 targets. Mol. Cell. Biol. 26,
33. Gerstung, M. et al. Combining gene mutation with Blood 122, 999–1006 (2013). 6880–6889 (2006).
gene expression data improves outcome prediction in 59. Yang, L., Rau, R. & Goodell, M. A. DNMT3A in 85. Huynh, K. D., Fischle, W., Verdin, E. & Bardwell, V. J.
myelodysplastic syndromes. Nat. Commun. 6, 5901 haematological malignancies. Nat. Rev. Cancer 15, BCoR, a novel corepressor involved in BCL‑6
(2015). 152–165 (2015). repression. Genes Dev. 14, 1810–1823 (2000).
34. Kon, A. et al. Recurrent mutations in multiple 60. Ley, T. J. et al. DNMT3A mutations in acute myeloid 86. Damm, F. et al. BCOR and BCORL1 mutations in
components of the cohesin complex in myeloid leukemia. N. Engl. J. Med. 363, 2424–2433 (2010). myelodysplastic syndromes and related disorders.
neoplasms. Nat. Genet. 45, 1232–1237 (2013). 61. Walter, M. J. et al. Recurrent DNMT3A mutations in Blood 122, 3169–3177 (2013).
35. Bejar, R. et al. Somatic mutations predict poor patients with myelodysplastic syndromes. Leukemia 87. Scheuermann, J. C. et al. Histone H2A deubiquitinase
outcome in patients with myelodysplastic syndrome 25, 1153–1158 (2011). activity of the Polycomb repressive complex PR‑DUB.
after hematopoietic stem-cell transplantation. J. Clin. 62. Ko, M. et al. TET proteins and 5‑methylcytosine Nature 465, 243–247 (2010).
Oncol. 32, 2691–2698 (2014). oxidation in hematological cancers. Immunol. Rev. 88. Abdel-Wahab, O. et al. ASXL1 mutations promote
36. Ortmann, C. A. et al. Effect of mutation order on 263, 6–21 (2015). myeloid transformation through loss of PRC2‑mediated
myeloproliferative neoplasms. N. Engl. J. Med. 372, 63. Tahiliani, M. et al. Conversion of 5‑methylcytosine to gene repression. Cancer Cell 22, 180–193 (2012).
601–612 (2015). 5‑hydroxymethylcytosine in mammalian DNA by MLL 89. Balasubramani, A. et al. Cancer-associated ASXL1
The first demonstration that mutational order partner TET1. Science 324, 930–935 (2009). mutations may act as gain‑of‑function mutations of the
alters clinical phenotype and outcomes in MPN. 64. Lorsbach, R. B. et al. TET1, a member of a novel ASXL1−BAP1 complex. Nat. Commun. 6, 7307 (2015).
37. Jeromin, S. et al. Refractory anemia with ring protein family, is fused to MLL in acute myeloid 90. Cao, Q. et al. BCOR regulates myeloid cell
sideroblasts and marked thrombocytosis cases harbor leukemia containing the t(10;11)(q22;q23). Leukemia proliferation and differentiation. Leukemia 30,
mutations in SF3B1 or other spliceosome genes 17, 637–641 (2003). 1155–1165 (2016).
accompanied by JAK2V617F and ASXL1 mutations. 65. Ono, R. et al. LCX, leukemia-associated protein with a 91. Thota, S. et al. Genetic alterations of the cohesin
Haematologica 100, e125–e127 (2015). CXXC domain, is fused to MLL in acute myeloid complex genes in myeloid malignancies. Blood 124,
38. Malcovati, L. et al. SF3B1 mutation identifies a leukemia with trilineage dysplasia having t(10;11) 1790–1798 (2014).
distinct subset of myelodysplastic syndrome with ring (q22;q23). Cancer Res. 62, 4075–4080 (2002). 92. Losada, A. Cohesin in cancer: chromosome
sideroblasts. Blood 126, 233–241 (2015). 66. Okano, M., Xie, S. & Li, E. Cloning and segregation and beyond. Nat. Rev. Cancer 14,
39. Ding, L. et al. Clonal evolution in relapsed acute characterization of a family of novel mammalian DNA 389–393 (2014).
myeloid leukaemia revealed by whole-genome (cytosine‑5) methyltransferases. Nat. Genet. 19, 93. Horsfield, J. A. et al. Cohesin-dependent regulation of
sequencing. Nature 481, 506–510 (2012). 219–220 (1998). Runx genes. Development 134, 2639–2649 (2007).
40. Garg, M. et al. Profiling of somatic mutations in acute 67. Kallin, E. M. et al. Tet2 facilitates the derepression 94. Leeke, B., Marsman, J., O’Sullivan, J. M. &
myeloid leukemia with FLT3‑ITD at diagnosis and of myeloid target genes during CEBPα-induced Horsfield, J. A. Cohesin mutations in myeloid
relapse. Blood 126, 2491–2501 (2015). transdifferentiation of pre‑B cells. Mol. Cell 48, malignancies: underlying mechanisms. Exp. Hematol.
41. Jan, M. et al. Clonal evolution of preleukemic 266–276 (2012). Oncol. 3, 13 (2014).
hematopoietic stem cells precedes human acute myeloid 68. Rasmussen, K. D. et al. Loss of TET2 in hematopoietic 95. Marsman, J. et al. Cohesin and CTCF differentially
leukemia. Sci. Transl Med. 4, 149ra118 (2012). cells leads to DNA hypermethylation of active regulate spatiotemporal runx1 expression during
42. Shlush, L. I. et al. Identification of pre-leukaemic enhancers and induction of leukemogenesis. Genes zebrafish development. Biochim. Biophys. Acta 1839,
haematopoietic stem cells in acute leukaemia. Nature Dev. 29, 910–922 (2015). 50–61 (2014).
506, 328–333 (2014). 69. Yamazaki, J. et al. TET2 mutations affect non-CpG island 96. Mazumdar, C. et al. Leukemia-associated cohesin
43. Klco, J. M. et al. Association between mutation DNA methylation at enhancers and transcription factor- mutants dominantly enforce stem cell programs and
clearance after induction therapy and outcomes in binding sites in chronic myelomonocytic leukemia. impair human hematopoietic progenitor
acute myeloid leukemia. JAMA 314, 811–822 Cancer Res. 75, 2833–2843 CAN‑14‑0739 (2015). differentiation. Cell Stem Cell 17, 675–688 (2015).
(2015). 70. Dang, L. et al. Cancer-associated IDH1 mutations 97. Hahn, C. N. et al. Heritable GATA2 mutations
44. Walter, M. J. et al. Clonal diversity of recurrently produce 2‑hydroxyglutarate. Nature 465, 966 associated with familial myelodysplastic syndrome and
mutated genes in myelodysplastic syndromes. (2010). acute myeloid leukemia. Nat. Genet. 43, 1012–1017
Leukemia 27, 1275–1282 (2013). 71. Ward, P. S. et al. The common feature of leukemia- (2011).
45. Yoshizato, T. et al. Somatic mutations and clonal associated IDH1 and IDH2 mutations is a neomorphic 98. Ito, Y., Bae, S. C. & Chuang, L. S. The RUNX family:
hematopoiesis in aplastic anemia. N. Engl. J. Med. enzyme activity converting α-ketoglutarate to developmental regulators in cancer. Nat. Rev. Cancer
373, 35–47 (2015). 2‑hydroxyglutarate. Cancer Cell 17, 225–234 (2010). 15, 81–95 (2015).
The largest sequenced cohort of aplastic anaemia 72. Xu, W. et al. Oncometabolite 2‑hydroxyglutarate is a 99. Zhang, M. Y. et al. Germline ETV6 mutations in familial
patients. competitive inhibitor of alpha-ketoglutarate- thrombocytopenia and hematologic malignancy. Nat.
46. Alexandrov, L. B. et al. Signatures of mutational dependent dioxygenases. Cancer Cell 19, 17–30 Genet. 47, 180–185 (2015).
processes in human cancer. Nature 500, 415–421 (2011). 100. Growney, J. D. et al. Loss of Runx1 perturbs
(2013). 73. Losman, J. A. et al. (R)-2‑hydroxyglutarate is sufficient adult hematopoiesis and is associated with a
47. Greenman, C. et al. Patterns of somatic mutation in to promote leukemogenesis and its effects are myeloproliferative phenotype. Blood 106, 494–504
human cancer genomes. Nature 446, 153–158 (2007). reversible. Science 339, 1621–1625 (2013). (2005).

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 13


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

101. Owen, C. J. et al. Five new pedigrees with inherited 124. Haase, D. et al. New insights into the prognostic 149. McNerney, M. E. et al. CUX1 is a haploinsufficient
RUNX1 mutations causing familial platelet disorder impact of the karyotype in MDS and correlation with tumor suppressor gene on chromosome 7 frequently
with propensity to myeloid malignancy. Blood 112, subtypes: evidence from a core dataset of 2124 inactivated in acute myeloid leukemia. Blood 121,
4639–4645 (2008). patients. Blood 110, 4385–4395 (2007). 975–983 (2013).
102. Tsai, F. Y. et al. An early haematopoietic defect in mice 125. Bochtler, T., Frohling, S. & Kramer, A. Role of 150. Nikoloski, G. et al. Somatic mutations of the histone
lacking the transcription factor GATA‑2. Nature 371, chromosomal aberrations in clonal diversity and methyltransferase gene EZH2 in myelodysplastic
221–226 (1994). progression of acute myeloid leukemia. Leukemia 29, syndromes. Nat. Genet. 42, 665–667 (2010).
103. Bejar, R. et al. TET2 mutations predict response to 1243–1252 (2015). 151. Fenaux, P. et al. Efficacy of azacitidine compared with
hypomethylating agents in myelodysplastic syndrome 126. Bernasconi, P. et al. Does cytogenetic evolution have that of conventional care regimens in the treatment
patients. Blood 124, 2705–2712 (2014). any prognostic relevance in myelodysplastic of higher-risk myelodysplastic syndromes: a
104. Rampal, R. et al. DNA hydroxymethylation profiling syndromes? A study on 153 patients from a single randomised, open-label, phase III study. Lancet.
reveals that WT1 mutations result in loss of TET2 institution. Ann. Hematol. 89, 545–551 (2010). Oncol. 10, 223–232 (2009).
function in acute myeloid leukemia. Cell Rep. 9, 127. Wang, H., Wang, X. Q., Xu, X. P. & Lin, G. W. This clinical trial showed that 5‑azacitidine
1841–1855 (2014). Cytogenetic evolution correlates with poor prognosis improves survival in higher risk patients with MDS
105. Wang, Y. et al. WT1 recruits TET2 to regulate its target in myelodysplastic syndrome. Cancer Genet. and led to its clinical approval.
gene expression and suppress leukemia cell Cytogenet. 196, 159–166 (2010). 152. Kantarjian, H. et al. Decitabine improves patient
proliferation. Mol. Cell 57, 662–673 (2015). 128. Komrokji, R. S., Padron, E., Ebert, B. L. & List, A. F. outcomes in myelodysplastic syndromes: results of a
106. Soenen, V. et al. 17p Deletion in acute myeloid Deletion 5q MDS: molecular and therapeutic phase III randomized study. Cancer 106, 1794–1803
leukemia and myelodysplastic syndrome. Analysis of implications. Best practice and research. Clin. (2006).
breakpoints and deleted segments by fluorescence Haematol. 26, 365–375 (2013). 153. Traina, F. et al. Impact of molecular mutations on
in situ. Blood 91, 1008–1015 (1998). 129. Van den Berghe, H. et al. Distinct haematological treatment response to DNMT inhibitors in
107. Rucker, F. G. et al. TP53 alterations in acute myeloid disorder with deletion of long arm of no. 5 myelodysplasia and related neoplasms. Leukemia 28,
leukemia with complex karyotype correlate with chromosome. Nature 251, 437–438 (1974). 78–87 (2014).
specific copy number alterations, monosomal 130. Chen, T. H. et al. Knockdown of Hspa9, a del(5q31.2) 154. Jadersten, M. et al. Clonal heterogeneity in the 5q−
karyotype, and dismal outcome. Blood 119, gene, results in a decrease in hematopoietic syndrome: 53 expressing progenitors prevail during
2114–2121 (2012). progenitors in mice. Blood 117, 1530–1539 lenalidomide treatment and expand at disease
108. Lowe, S. W., Cepero, E. & Evan, G. Intrinsic tumour (2011). progression. Haematologica 94, 1762–1766
suppression. Nature 432, 307–315 (2004). 131. Kumar, M. S. et al. Coordinate loss of a microRNA and (2009).
109. Kaneko, H., Misawa, S., Horiike, S., Nakai, H. & protein-coding gene cooperate in the pathogenesis of 155. Raza, A. et al. Phase 2 study of lenalidomide in
Kashima, K. TP53 mutations emerge at early phase of 5q- syndrome. Blood 118, 4666–4673 (2011). transfusion-dependent, low-risk, and intermediate‑1
myelodysplastic syndrome and are associated with 132. Starczynowski, D. T. et al. Identification of miR‑145 risk myelodysplastic syndromes with karyotypes other
complex chromosomal abnormalities. Blood 85, and miR‑146a as mediators of the 5q- syndrome than deletion 5q. Blood 111, 86–93 (2008).
2189–2193 (1995). phenotype. Nat. Med. 16, 49–58 (2010). 156. Koreth, J. et al. Role of reduced-intensity
110. Swisher, E. M. et al. Somatic mosaic mutations in 133. Stoddart, A. et al. Haploinsufficiency of del(5q) genes, conditioning allogeneic hematopoietic stem-cell
PPM1D and TP53 in the blood of women with ovarian Egr1 and Apc, cooperate with Tp53 loss to induce transplantation in older patients with de novo
carcinoma. JAMA Oncol. 2, 370–372 (2016). acute myeloid leukemia in mice. Blood 123, myelodysplastic syndromes: an international
111. Wong, T. N. et al. Role of TP53 mutations in the origin 1069–1078 (2014). collaborative decision analysis. J. Clin. Oncol. 31,
and evolution of therapy-related acute myeloid 134. Grisendi, S. et al. Role of nucleophosmin in embryonic 2662–2670 (2013).
leukaemia. Nature 518, 552–555 (2015). development and tumorigenesis. Nature 437, 157. Della Porta, M. G. et al. Clinical effects of driver
TP53 mutant HSCs expand preferentially during 147–153 (2005). somatic mutations on the outcomes of patients with
treatment with chemotherapy, explaining their 135. Varney, M. E. et al. Loss of Tifab, a del(5q) MDS gene, myelodysplastic syndromes treated with allogeneic
increased frequency in therapy-related disease. alters hematopoiesis through derepression of Toll-like hematopoietic stem-cell transplantation. J. Clin.
112. Zhao, Z. et al. p53 loss promotes acute myeloid receptor‑TRAF6 signaling. J. Exp. Med. 212, Oncol. 30, 3627–3637 (2016).
leukemia by enabling aberrant self-renewal. Genes 1967–1985 (2015). 158. Chaturvedi, A. et al. Mutant IDH1 promotes
Dev. 24, 1389–1402 (2010). 136. DeWard, A. D., Leali, K., West, R. A., leukemogenesis in vivo and can be specifically
113. Marusyk, A., Porter, C. C., Zaberezhnyy, V. & Prendergast, G. C. & Alberts, A. S. Loss of RhoB targeted in human AML. Blood 122, 2877–2887
DeGregori, J. Irradiation selects for p53‑deficient expression enhances the myelodysplastic phenotype (2013).
hematopoietic progenitors. PLoS Biol. 8, e1000324 of mammalian diaphanous-related Formin mDia1 159. Wang, F. et al. Targeted inhibition of mutant IDH2 in
(2010). knockout mice. PLoS ONE 4, e7102 (2009). leukemia cells induces cellular differentiation. Science
114. Walter, M. J. et al. Clonal architecture of secondary 137. Graubert, T. A. et al. Integrated genomic analysis 340, 622–626 (2013).
acute myeloid leukemia. N. Engl. J. Med. 366, implicates haploinsufficiency of multiple chromosome 160. Chan, S. M. et al. Isocitrate dehydrogenase 1 and 2
1090–1098 (2012). 5q31.2 genes in de novo myelodysplastic syndromes mutations induce BCL‑2 dependence in acute myeloid
115. Furitsu, T. et al. Identification of mutations in the pathogenesis. PloS ONE 4, e4583 (2009). leukemia. Nat. Med. 21, 178–184 (2015).
coding sequence of the proto-oncogene c‑kit in a 138. Polprasert, C. et al. Inherited and somatic defects 161. Fandy, T. E. et al. Early epigenetic changes and DNA
human mast cell leukemia cell line causing ligand- in DDX41 in myeloid neoplasms. Cancer Cell 27, damage do not predict clinical response in an
independent activation of c‑kit product. J. Clin. Invest. 658–670 (2015). overlapping schedule of 5‑azacytidine and entinostat
92, 1736–1744 (1993). 139. Schneider, R. K. et al. Role of casein kinase 1A1 in the in patients with myeloid malignancies. Blood 114,
116. Kottaridis, P. D. et al. The presence of a FLT3 internal biology and targeted therapy of del(5q) MDS. Cancer 2764–2773 (2009).
tandem duplication in patients with acute myeloid Cell 26, 509–520 (2014). 162. Lee, S. C. et al. Modulation of splicing catalysis for
leukemia (AML) adds important prognostic information 140. Ebert, B. L. et al. Identification of RPS14 as a 5q- therapeutic targeting of leukemia with mutations in
to cytogenetic risk group and response to the first cycle syndrome gene by RNA interference screen. Nature genes encoding spliceosomal proteins. Nat. Med. 22,
of chemotherapy: analysis of 854 patients from the 451, 335–339 (2008). 672–678 (2016).
United Kingdom Medical Research Council AML 10 141. Barlow, J. L. et al. A p53‑dependent mechanism 163. Maxson, J. E. et al. Oncogenic CSF3R mutations in
and 12 trials. Blood 98, 1752–1759 (2001). underlies macrocytic anemia in a mouse model of chronic neutrophilic leukemia and atypical CML.
117. Kralovics, R. et al. A gain‑of‑function mutation of JAK2 human 5q− syndrome. Nat. Med. 16, 59–66 (2010). N. Engl. J. Med. 368, 1781–1790 (2013).
in myeloproliferative disorders. N. Engl. J. Med. 352, 142. Schneider, R. K. et al. Rps14 haploinsufficiency causes a 164. Meggendorfer, M. et al. SETBP1 mutations occur in
1779–1790 (2005). block in erythroid differentiation mediated by S100A8 9% of MDS/MPN and in 4% of MPN cases and are
118. Makishima, H. et al. Mutations of e3 ubiquitin ligase and S100A9. Nat. Med. 22, 288–297 (2016). strongly associated with atypical CML, monosomy 7,
cbl family members constitute a novel common 143. Ludwig, L. S. et al. Altered translation of GATA1 in isochromosome i(17)(q10), ASXL1 and CBL mutations.
pathogenic lesion in myeloid malignancies. J. Clin. Diamond-Blackfan anemia. Nat. Med. 20, 748–753 Leukemia 27, 1852–1860 (2013).
Oncol. 27, 6109–6116 (2009). (2014). 165. Mason, C. C. et al. Age-related mutations and chronic
119. Pardanani, A. D. et al. MPL515 mutations in 144. Vlachos, A. et al. Diminutive somatic deletions in the myelomonocytic leukemia. Leukemia 30, 906–913
myeloproliferative and other myeloid disorders: a 5q region lead to a phenotype atypical of classical 5q− (2016).
study of 1182 patients. Blood 108, 3472–3476 syndrome. Blood 122, 2487–2490 (2013). 166. Lane, A. A. et al. Low frequency clonal mutations
(2006). 145. Kronke, J. et al. Lenalidomide induces ubiquitination recoverable by deep sequencing in patients with
120. Sanada, M. et al. Gain‑of‑function of mutated C‑CBL and degradation of CK1α in del(5q) MDS. Nature aplastic anemia. Leukemia 27, 968–971 (2013).
tumour suppressor in myeloid neoplasms. Nature 523, 183–188 (2015). 167. Shen, W. et al. Deep sequencing reveals stepwise
460, 904–908 (2009). This study describes the mechanism of action of mutation acquisition in paroxysmal nocturnal
121. Sargin, B. et al. Flt3‑dependent transformation by lenalidomide in MDS. hemoglobinuria. J. Clin. Invest. 124, 4529–4538
inactivating c‑Cbl mutations in AML. Blood 110, 146. List, A. et al. Lenalidomide in the myelodysplastic (2014).
1004–1012 (2007). syndrome with chromosome 5q deletion. N. Engl. 168. Wilson, D. B., Link, D. C., Mason, P. J. & Bessler, M.
122. Saur, S. J., Sangkhae, V., Geddis, A. E., Kaushansky, K. J. Med. 355, 1456–1465 (2006). Inherited bone marrow failure syndromes in
& Hitchcock, I. S. Ubiquitination and degradation of The first clinical trial demonstrating that adolescents and young adults. Ann. Med. 46,
the thrombopoietin receptor c‑Mpl. Blood 115, lenalidomide improves outcomes in the 353–363 (2014).
1254–1263 (2010). 5q− syndrome. 169. Yamaguchi, H. et al. Mutations of the human
123. Javadi, M., Richmond, T. D., Huang, K. & 147. Ito, T. et al. Identification of a primary target of telomerase RNA gene (TERC) in aplastic anemia and
Barber, D. L. CBL linker region and RING finger thalidomide teratogenicity. Science 327, 1345–1350 myelodysplastic syndrome. Blood 102, 916–918
mutations lead to enhanced granulocyte-macrophage (2010). (2003).
colony-stimulating factor (GM‑CSF) signaling via 148. Chen, C. et al. MLL3 is a haploinsufficient 7q tumor 170. Yamaguchi, H. et al. Mutations in TERT, the gene for
elevated levels of JAK2 and LYN. J. Biol. Chem. 288, suppressor in acute myeloid leukemia. Cancer Cell 25, telomerase reverse transcriptase, in aplastic anemia.
19459–19470 (2013). 652–665 (2014). N. Engl. J. Med. 352, 1413–1424 (2005).

14 | ADVANCE ONLINE PUBLICATION www.nature.com/nrc


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

171. Mitchell, J. R., Wood, E. & Collins, K. A telomerase 183. Ding, L., Saunders, T. L., Enikolopov, G. & 194. Papaemmanuil, E. et al. Genomic Classification and
component is defective in the human disease Morrison, S. J. Endothelial and perivascular cells Prognosis in Acute Myeloid Leukemia. N. Engl. J. Med.
dyskeratosis congenita. Nature 402, 551–555 maintain haematopoietic stem cells. Nature 481, 374, 2209–2221 (2016).
(1999). 457–462 (2012). 195. Boultwood, J., Dolatshad, H., Varanasi, S. S.,
172. Vulliamy, T. et al. Disease anticipation is associated 184. Zhou, B. O., Yue, R., Murphy, M. M., Peyer, J. G. & Yip, B. H. & Pellagatti, A. The role of splicing factor
with progressive telomere shortening in families with Morrison, S. J. Leptin-receptor-expressing mutations in the pathogenesis of the
dyskeratosis congenita due to mutations in TERC. mesenchymal stromal cells represent the main source myelodysplastic syndromes. Adv. Biol. Regul. 54,
Nat. Genet. 36, 447–449 (2004). of bone formed by adult bone marrow. Cell Stem Cell 153–161 (2014).
173. Noetzli, L. et al. Germline mutations in ETV6 are 15, 154–168 (2014). 196. Fenaux, P. et al. A randomized phase 3 study of
associated with thrombocytopenia, red cell 185. Raaijmakers, M. H. et al. Bone progenitor dysfunction lenalidomide versus placebo in RBC transfusion-
macrocytosis and predisposition to lymphoblastic induces myelodysplasia and secondary leukaemia. dependent patients with Low-/Intermediate‑1‑risk
leukemia. Nat. Genet. 47, 535–538 (2015). Nature 464, 852–857 (2010). myelodysplastic syndromes with del5q. Blood 118,
174. McBride, K. A. et al. Li‑Fraumeni syndrome: cancer 186. Kode, A. et al. Leukaemogenesis induced by an 3765–3776 (2011).
risk assessment and clinical management. Nat. Rev. activating beta-catenin mutation in osteoblasts. 197. Cordoba, I. et al. Better prognosis for patients with
Clin. Oncol. 11, 260–271 (2014). Nature 506, 240–244 (2014). del(7q) than for patients with monosomy 7 in
175. Hsu, A. P. et al. Mutations in GATA2 are associated with 187. Schepers, K. et al. Myeloproliferative neoplasia myelodysplastic syndrome. Cancer 118, 127–133
the autosomal dominant and sporadic monocytopenia remodels the endosteal bone marrow niche into a (2012).
and mycobacterial infection (MonoMAC) syndrome. self-reinforcing leukemic niche. Cell Stem Cell 13, 198. Bacher, U. et al. Investigation of 305 patients with
Blood 118, 2653–2655 (2011). 285–299 (2013). myelodysplastic syndromes and 20q deletion for
176. Ostergaard, P. et al. Mutations in GATA2 cause 188. Epperson, D. E., Nakamura, R., Saunthararajah, Y., associated cytogenetic and molecular genetic lesions
primary lymphedema associated with a predisposition Melenhorst, J. & Barrett, A. J. Oligoclonal T cell and their prognostic impact. Br. J. Haematol. 164,
to acute myeloid leukemia (Emberger syndrome). Nat. expansion in myelodysplastic syndrome: evidence 822–833 (2014).
Genet. 43, 929–931 (2011). for an autoimmune process. Leukemia Res. 25, 199. Wang, S. A. et al. Myelodysplastic syndromes with
177. Lewinsohn, M. et al. Novel germ line DDX41 1075–1083 (2001). deletions of chromosome 11q lack cryptic MLL
mutations define families with a lower age of MDS/ 189. Molldrem, J. J. et al. Antithymocyte globulin for rearrangement and exhibit characteristic
AML onset and lymphoid malignancies. Blood 127, treatment of the bone marrow failure associated with clinicopathologic features. Leuk. Res. 35, 351–357
1017–1023 (2016). myelodysplastic syndromes. Ann. Intern. Med. 137, (2011).
178. Boocock, G. R. et al. Mutations in SBDS are 156–163 (2002). 200. Wong, A. K. et al. Loss of the Y chromosome: an
associated with Shwachman–Diamond syndrome. 190. Saunthararajah, Y., Nakamura, R., Wesley, R., age-related or clonal phenomenon in acute
Nat. Genet. 33, 97–101 (2003). Wang, Q. J. & Barrett, A. J. A simple method to myelogenous leukemia/myelodysplastic syndrome?
179. Zhou, T., Kinney, M. C., Scott, L. M., Zinkel, S. S. & predict response to immunosuppressive therapy in Arch. Pathol. Lab. Med. 132, 1329–1332
Rebel, V. I. Revisiting the case for genetically patients with myelodysplastic syndrome. Blood 102, (2008).
engineered mouse models in human myelodysplastic 3025–3027 (2003).
syndrome research. Blood 126, 1057–1068 (2015). 191. Sloand, E. M., Wu, C. O., Greenberg, P., Young, N. &
Acknowledgements
180. Li, Q. et al. Oncogenic Nras has bimodal effects on Barrett, J. Factors affecting response and survival in
This work was supported by grants from the US National
stem cells that sustainably increase competitiveness. patients with myelodysplasia treated with
Institutes of Health (NIH) (R01 HL082945 and R24
Nature 504, 143–147 (2013). immunosuppressive therapy. J. Clin. Oncol. 26,
DK099808), the Department of Defense, the Edward P.
181. Dolatshad, H. et al. Disruption of SF3B1 results in 2505–2511 (2008).
Evans Foundation, the Leukemia and Lymphoma Society and
deregulated expression and splicing of key genes and 192. Cancer Genome Atlas Research Network. Genomic and
the STARR Cancer Consortium to B.L.E.
pathways in myelodysplastic syndrome hematopoietic epigenomic landscapes of adult de novo acute myeloid
stem and progenitor cells. Leukemia 29, 1798 (2015). leukemia. N. Engl. J. Med. 368, 2059–2074 (2013).
182. Chan, C. K. et al. Endochondral ossification is required 193. Patel, J. P. et al. Prognostic relevance of integrated Competing interests statement
for haematopoietic stem-cell niche formation. Nature genetic profiling in acute myeloid leukemia. N. Engl. The authors declare competing interests: see Web version
457, 490–494 (2009). J. Med. 366, 1079–1089 (2012). for details.

NATURE REVIEWS | CANCER ADVANCE ONLINE PUBLICATION | 15


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like