Fanconi anemia (FA) is the most frequent inherited cause of BM failure (BMF). Fifteen FANC genes have been identified to date, the most prevalent being FANCA, FANCC, FANCG, and FANCD2. In addition to classical presentations with progressive BMF during childhood and a positive chromosome breakage test in the blood, atypical clinical and/or biological situations can be seen in which a FA diagnosis has to be confirmed or eliminated. For this, a range of biological tools have been developed, including analysis of skin fibroblasts. FA patients experience a strong selective pressure in the BM that predisposes to clonal evolution and to the emergence in their teens or young adulthood of myelodysplasia syndrome (MDS) and/or acute myeloid leukemia (AML) with a specific pattern of somatic chromosomal lesions. The cellular mechanisms underlying (1) the hematopoietic defect which leads to progressive BMF and (2) somatic clonal evolutions in this background, are still largely elusive. Elucidation of these mechanisms at the molecular and cellular levels should be useful to understand the physiopathology of the disease and to adapt the follow-up and treatment of FA patients. This may also ultimately benefit older, non-FA patients with aplastic anemia, MDS/AML for whom FA represents a model genetic condition.

Fanconi anemia (FA) is the most frequent inherited cause of BM failure (BMF).1  The FA genes (genes that have been to be found mutated in FA patients) are called FANC, the most frequent being FANCA, FANCC, FANCG, and FANCD2.2  Except for the very rare FANCB, which is located on the X chromosome,3  all other FANC genes are autosomic and the disease is recessive. There are typically several clinical stages in FA that are related to age.1,4,5  At birth and early childhood, only physical signs are present and range from discreet to extensive. Many patients then experience BMF between 5 and 15 years of age, and the diagnosis of FA is often made at this stage. Later on, during their teens or young adulthood, the risk of acute myeloid leukemia (AML) and myelodysplastic syndrome (MDS)/AML becomes very high. Still later, in adult patients, a range of solid cancers can be seen, especially mouth cancer. Throughout life, the hematopoietic situation can change spontaneously by genetic reversion or other clonal evolution, including progression to MDS or AML. Hematopoietic stem cell transplantation (HSCT) is currently the best treatment to cure severe aplastic anemia or MDS/AML.6  However, HSCT has to be carefully considered based upon clinical and biological criteria including age, severity of the cytopenia, significant BM dysplasia, excess of blast cells, cytogenetic abnormalities, and immunological compatibility with the donor.

To date, 15 FA genes have been identified (FANCA, FANCB, FANCC, FANCD1/BRCA2, FANCD2, FANCE, FANCF, FANCG, FANCI, FANCJ/BRIP1, FANCL, FANCM, FANCN/PALB2, FANCO/RAD51C, and FANCP/SLX4).2,7  It is noteworthy that FANCD1 is identical to the BRCA2 gene, one of the genes predisposing to hereditary breast and ovarian cancer.8  Products of the FA genes function in a common DNA repair signaling pathway, the FA/BRCA pathway (Figure 1), which closely cooperates with other DNA repair proteins for resolving DNA interstrand cross-links (ICLs) during replication (reviewed in Moldovan et al7  and in Kee et al9 ). A central event in the pathway is the mono-ubiquitination of FANCD2 and FANCI upon DNA damage, which is mediated by a group of upstream FA proteins (FANCA, FANCB, FANCC, FANCE, FANCF, FANCG, FANCL, and FANCM) that are assembled into a large nuclear E3 ubiquitin ligase complex called the FA core complex. The mono-ubiquitinated FANCD2 and FANCI heterodimer functionally interacts with downstream FA proteins such as FANCD1/BRCA2, FANCN/PALB2, FANCJ/BRIP1, FANCP/SLX4, RAD51C, and their associated protein, BRCA1. FAN1 (Fanconi-associated nuclease 1), a recently identified FA-associated protein, provides a nuclease activity during the ICL repair.10–13  FA proteins could also have other functions or participate in pathways other than DNA repair in response to cellular stress.14–17 

Figure 1.

The FA/BRCA biological pathway. FANC proteins are shown in green.

Figure 1.

The FA/BRCA biological pathway. FANC proteins are shown in green.

Close modal

FA patients often, but not always, present with a combination of various congenital abnormalities such as short stature; Fanconi facies and microphthalmia; thumb and radius deformities; skin hyperpigmentation such as “café-au-lait spots”; and cardiac, renal, genitourinary, and/or other malformations (reviewed in Shimamura and Alter1 ). Physical abnormalities can be subtle or absent. Many FA patients develop BMF some time during the course of the disease, usually during their first and second decades of life,4,5  and for the majority of patients, the suspicion of FA will only be made after the onset of pancytopenia. In some patients, the underlying diagnosis of FA is not known until MDS/AML occurs. Increases of fetal hemoglobin, serum alpha-fetoprotein, and macrocytosis are commonly noted in FA, but their absence does not rule out the disease (although not specific for FA, they may help to distinguish an inherited from an acquired BMF).1  A significant, and probably underestimated, number of patients do not develop overt BMF but still have an increased risk of malignancies.

It is crucial for family counseling and treatment centers to identify patients with FA. Patients with BMF who happen to have underlying undiagnosed FA will not respond to the immunosuppression therapy used to treat patients with idiopathic aplastic anemia. Moreover, due to a hypersensitivity to chemotherapy agents, patients with FA will often die of toxicity if given conventional conditioning for HSCT, and therefore less myeloablative regimens are used in this population.6  In addition, to being at a higher risk for developing malignancies, patients with FA need appropriate cancer surveillance throughout life.18,19  Therefore, the consensus is that FA has to be screened for in all children or young adults with hypoplastic or aplastic anemia or cytopenias. It is also safe to screen children and young adults with MDS or AML with physical signs and/or FA-associated chromosomal abnormalities in the BM, such as 1q+ or 3q+ (see below).

Due to the many genes and mutations associated with FA, a single genetic test cannot be used as a first approach for FA diagnosis in unselected BMF patients. The biological diagnosis of FA is primarily based on the exquisite sensitivity of FA cells to DNA ICL chemicals such as diepoxybutane or mitomycin C. The chromosomal breakage test with these agents is the technique of reference for diagnosing FA.1,20  In the majority of cases, a precise diagnosis can be made with careful history, physical examination, and a positive chromosomal breakage blood test (Figure 2). Other blood tests include cell-cycle analysis21  and evaluation of FANCD2 mono-ubiquitination, which can positively diagnose FA core patients.22  However, all of these tests can be ambiguous or even falsely negative in patients who develop hematopoietic reversion and somatic mosaicism. Hematopoietic reversion occurs when, after a spontaneous genetic event in a hematopoietic stem cell (ie, a reverse point mutation or intragenic recombination), one FA allele is corrected, with a consequent recovery of a normal or subnormal protein activity and cellular phenotype.23,24  Because there has been no evidence that this same phenomenon could happen in primary skin fibroblasts, these cells have been used to overcome misleading results in blood due to somatic mosaicism.25–28  Primary fibroblast cells can be tested for their ICL sensitivity by chromosome breaks (although less conveniently than in blood cells), growth curves, FANCD2 test, or flow-derived tests.26–29  Moreover, fibroblast DNA is useful for subsequent screening for the FANC mutation and also as a constitutional reference when analyzing somatic molecular and chromosomal abnormalities in the BM.30  Once the FA diagnosis is established at the cellular level, FANC gene mutations can be screened. The most frequent gene (FANCA) is analyzed first, then FANCG, FANCC, and so on. Alternatively, or in difficult cases, retroviral complementation can be used to ascertain the FA group before sequencing. The FANCD2 mono-ubiquitination test in fibroblasts is useful to orientate the search toward a FA core, FANCD2 or downstream gene. FA mutations are being referred to in the Fanconi Anemia Mutation Database (www.rockefeller.edu/fanconi). For FANCD2 and genes from the downstream groups, including FANCD1/BRCA2, at least one of the mutations is hypomorphic because the complete absence of protein is lethal.

Figure 2.

FA biological diagnosis is based on the hypersensitivity of FA cells to ICL agents.

Figure 2.

FA biological diagnosis is based on the hypersensitivity of FA cells to ICL agents.

Close modal

There are few clear genetic-phenotype correlations in the classical FA core patients, although “hypomorphic” mutations might be associated with milder phenotypes and could allow time to develop to clonal evolution and late-onset solid cancer.31–33  In contrast, FANCD2 patients usually experience a more severe phenotype,34  and FANCD1/BRCA2 patients develop an early and rapidly lethal cancer-prone syndrome.35–37 

FA patients have a cumulative hematopoietic dysfunction likely related to an excess of genetic instability, cellular stress, p53 activation and cell death, which is proved to be intrinsic by the excellent efficiency of allogeneic HSCT in this population. In such a background of deficient hematopoiesis, FA BM cells experience a strong selective pressure toward clonal evolution, probably favored by the constitutive chromosomal instability (Figure 3).38–40  Myelodysplasia and AML cells have experienced clonal genetic evolution. Although a case of an AML cell line (from the D1/BRCA2 group) with genetic reversion has been reported,41  reversion does not seem frequent in the MDS and AML of FA patients,30  suggesting that genetic and/or epigenetic evolutions other than FA reversion happen. Acquired attenuation of the FA-prototypical G2 checkpoint and resistance to TNFα have been described in humans and mice, respectively.42,43  It is likely that these and other cellular phenotypes rescue FA cells and confer a selective advantage, but also predispose patients to develop malignancies. Therefore, a thorough characterization of the molecular and cellular features associated with the BM progression in FA should lead to an understanding the step-wise mechanisms of transformation, and to better prevention or treatment in these patients.

Figure 3.

Frequent spontaneous clonal evolution in FA BM.

Figure 3.

Frequent spontaneous clonal evolution in FA BM.

Close modal

Clinically, FA patients experience MDS and AML with the highest frequency during their teens or young adulthood.1,44  These diseases are often, but not always, preceded by an hypoplastic or aplastic phase. MDS in FA often presents as refractory cytopenia with multilineage dysplasia (according to the WHO 2008 classification), with or without excess of blasts.30,45  A certain level of dyserythropoiesis is almost constant in FA, and a mild dyserythropoiesis is not considered as an MDS criteria in this population. Acute leukemia can be diagnosed primarily (in approximately 30% of the AML in FA patients) or after a MDS phase with an increasing fraction of blast cells in the BM. Karyotypic abnormalities are frequently found, and translocations of chromosome 1q, monosomy 7, and gains of 3q have been reported.30,44–48  Using array-based high-density chromosomal profiling and oncogene sequencing, we recently analyzed BM samples from a large series of FA patients at various stages of the disease (aplastic, MDS, and AML).30  We found a highly recurrent pattern of somatic abnormalities that were related to unbalanced chromosomal translocations and led to partial chromosomal arm duplications or losses (Figure 4). In contrast to what is seen in non-FA MDS/AML patients,49  somatic copy-neutral loss of heterozygosity (uniparental disomy) were rarely found in FA, which is consistent with a constitutive defect of homologous recombination repair in FA. The most frequent lesion was partial duplication of chromosome 1q (1q+, 44.8% of 29 MDS/AML patients), after by 3q+ (41.3%), RUNX1/21q− (20.7%), monosomy 7/7q− (17.2%), and 11q− (13.8%). Mutations of MDS/AML oncogenes and tumor-suppressor genes were rarely found (isolated FLT3-ITD, NRAS, MLL-PTD, but no TP53, CBL, TET2, CEBPα, NPM1, and FLT3-TKD mutations in this series). It appears that the oncogenesis in FA shares common pathways with non-FA patients (monosomy 7/7q−, 5q−, RUNX1/21q− lesions, PRDM16 translocations, MLL-PTD), but also has additional FA-specific chromosomal lesions, especially 1q+ and 3q+. The molecular targets in these 2 chromosomal lesions are not known, although the EVI1 oncogene at 3q26 is a candidate. Interestingly, whereas the lesions 7q−, 3q+, and RUNX1 abnormalities were found at the MDS and AML stages only, translocation/duplications 1q+ can be seen at all stages in the BM, including “normal” or hypoplastic BM without apparent transformation signs (Figure 4), suggesting that 1q+ could rescue the FA cells without necessarily transforming them into MDS/AML.30 

Figure 4.

A recurrent profile of acquired chromosomal abnormalities in BM samples of FA patients. Samples are grouped by BM stages (normal, aplasia, MDS, and AML). The 1q+ and 3q+ lesions are specific from FA compared with MDS/AML of non-FA patients. Other abnormalities, including 7q− and RUNX1 mutations, can be found in non-FA patients. 1q+ is found at all stages of the BM progression, whereas most other lesions are found at the most advanced MDS/AML stages. (Used with permission from Quentin et al.30 )

Figure 4.

A recurrent profile of acquired chromosomal abnormalities in BM samples of FA patients. Samples are grouped by BM stages (normal, aplasia, MDS, and AML). The 1q+ and 3q+ lesions are specific from FA compared with MDS/AML of non-FA patients. Other abnormalities, including 7q− and RUNX1 mutations, can be found in non-FA patients. 1q+ is found at all stages of the BM progression, whereas most other lesions are found at the most advanced MDS/AML stages. (Used with permission from Quentin et al.30 )

Close modal

Because MDS/AML is a frequent and severe occurrence in FA, it is necessary to follow up on patients with regular BM aspirate tests with expert morphological and karyotype analyses to detect transformation before the onset of overt MDS/AML.1,20  In our study, conventional karyotype analysis appeared to be more sensitive than comparative genomic hybridization/single nucleotide polymorphism arrays for the early detection of discrete 1q+ subclones due to the possibility of observing individual cells and probably a clonal advantage in culture. Systematic interphasic FISH screening using probes for chromosome 7q, 3q, and break-apart RUNX1 might increase the sensitivity of detection of transformed cells. MDS/AML cases can have a normal karyotype but cryptic chromosomal or genomic abnormalities detected by FISH and/or array analysis.30  Therefore, when MDS or AML is diagnosed with a BM smear in FA patients, we now implement the karyotype analysis using high-density DNA arrays and RUNX1 FISH. This approach is useful in detecting cryptic lesions and in characterizing the abnormal karyotypes precisely, frequently highlighting meaningful chromosomal/molecular lesions.30  The predictive value of the various chromosomal/genomic abnormalities in patients with or without MDS/AML will have to be carefully evaluated on the long term in large cohorts of FA patients with respect to the therapeutic options and clinical benefits. For example, a sole clonal abnormality such as 1q+ can be present in a “normal” or non-MDS hypoplastic BM, and may not necessarily predict a progression into AML in the following years.

FA is a genetic condition that strongly predisposes patients to aplasia, MDS, and AML. Follow-up of FA patients requires a specialized multidisciplinary clinical and biological expertise. To optimize treatment, it is of critical importance to understand the step-wise molecular and cellular mechanisms of the BMF and clonal evolution. This knowledge may also ultimately benefit older, non-FA patients with aplastic anemia or MDS/AML, for whom close physiopathological cellular mechanisms are likely involved, including genomic instability, cytopenia, and clonal evolution.

Thanks to the patients and families, the French Association AFMF, the physicians who take care of the patients, the Centre de Référence Aplasies médullaires constitutionnelles Saint-Louis, Robert Debré, the Réseau INCa Maladies cassantes de, l' ADN, and to the members of the team Génome et Cancer at Saint-Louis Hospital, Paris, France, especially Raphael Ceccaldi.

Conflict-of-interest disclosure: The author declares no competing financial interests. Off-label drug use: None disclosed.

Jean Soulier, Hôpital Saint-Louis, 1 Avenue Claude-Vellefaux, 75010 Paris, France; Phone: +33-14249-9891; Fax: +33-14249-4027; e-mail: jean.soulier@sls.aphp.fr.

1
Shimamura
 
A
Alter
 
BP
Pathophysiology and management of inherited bone marrow failure syndromes
Blood Rev
2010
, vol. 
24
 
3
(pg. 
101
-
122
)
2
de Winter
 
JP
Joenje
 
H
The genetic and molecular basis of Fanconi anemia
Mutat Res
2009
, vol. 
668
 
1-2
(pg. 
11
-
19
)
3
Meetei
 
AR
Levitus
 
M
Xue
 
Y
, et al. 
X-linked inheritance of Fanconi anemia complementation group B
Nat Genet
2004
, vol. 
36
 
11
(pg. 
1219
-
1224
)
4
Kutler
 
DI
Singh
 
B
Satagopan
 
J
, et al. 
A 20-year perspective on the International Fanconi Anemia Registry (IFAR)
Blood
2003
, vol. 
101
 
4
(pg. 
1249
-
1256
)
5
Rosenberg
 
PS
Greene
 
MH
Alter
 
BP
Cancer incidence in persons with Fanconi anemia
Blood
2003
, vol. 
101
 
3
(pg. 
822
-
826
)
6
Gluckman
 
E
Wagner
 
JE
Hematopoietic stem cell transplantation in childhood inherited bone marrow failure syndrome
Bone Marrow Transplant
2008
, vol. 
41
 
2
(pg. 
127
-
132
)
7
Moldovan
 
GL
D'Andrea
 
AD
How the fanconi anemia pathway guards the genome
Annu Rev Genet
2009
, vol. 
43
 (pg. 
223
-
249
)
8
Howlett
 
NG
Taniguchi
 
T
Olson
 
S
, et al. 
Biallelic inactivation of BRCA2 in Fanconi anemia
Science
2002
, vol. 
297
 
5581
(pg. 
606
-
609
)
9
Kee
 
Y
D'Andrea
 
AD
Expanded roles of the Fanconi anemia pathway in preserving genomic stability
Genes Dev
2010
, vol. 
24
 
16
(pg. 
1680
-
1694
)
10
Liu
 
T
Ghosal
 
G
Yuan
 
J
Chen
 
J
Huang
 
J
FAN1 acts with FANCI-FANCD2 to promote DNA interstrand cross-link repair
Science
2010
, vol. 
329
 
5992
(pg. 
693
-
696
)
11
Kratz
 
K
Schopf
 
B
Kaden
 
S
, et al. 
Deficiency of FANCD2-associated nuclease KIAA1018/FAN1 sensitizes cells to interstrand crosslinking agents
Cell
2010
, vol. 
142
 
1
(pg. 
77
-
88
)
12
MacKay
 
C
Declais
 
AC
Lundin
 
C
, et al. 
Identification of KIAA1018/FAN1, a DNA repair nuclease recruited to DNA damage by monoubiquitinated FANCD2
Cell
2010
, vol. 
142
 
1
(pg. 
65
-
76
)
13
Smogorzewska
 
A
Desetty
 
R
Saito
 
TT
, et al. 
A genetic screen identifies FAN1, a Fanconi anemia-associated nuclease necessary for DNA interstrand crosslink repair
Mol Cell
2010
, vol. 
39
 
1
(pg. 
36
-
47
)
14
Naim
 
V
Rosselli
 
F
The FANC pathway and BLM collaborate during mitosis to prevent micro-nucleation and chromosome abnormalities
Nat Cell Biol
2009
, vol. 
11
 
6
(pg. 
761
-
768
)
15
Fagerlie
 
SR
Diaz
 
J
Christianson
 
TA
, et al. 
Functional correction of FA-C cells with FANCC suppresses the expression of interferon gamma-inducible genes
Blood
2001
, vol. 
97
 
10
(pg. 
3017
-
3024
)
16
Bagby
 
GC
Alter
 
BP
Fanconi anemia
Semin Hematol
2006
, vol. 
43
 
3
(pg. 
147
-
156
)
17
Briot
 
D
Mace-Aime
 
G
Subra
 
F
Rosselli
 
F
Aberrant activation of stress-response pathways leads to TNF-alpha oversecretion in Fanconi anemia
Blood
2008
, vol. 
111
 
4
(pg. 
1913
-
1923
)
18
Alter
 
BP
Cancer in Fanconi anemia, 1927-2001
Cancer
2003
, vol. 
97
 
2
(pg. 
425
-
440
)
19
Rosenberg
 
PS
Socie
 
G
Alter
 
BP
Gluckman
 
E
Risk of head and neck squamous cell cancer and death in patients with Fanconi anemia who did and did not receive transplants
Blood
2005
, vol. 
105
 
1
(pg. 
67
-
73
)
20
Auerbach
 
AD
Fanconi anemia and its diagnosis
Mutat Res
2009
, vol. 
668
 
1-2
(pg. 
4
-
10
)
21
Seyschab
 
H
Friedl
 
R
Sun
 
Y
, et al. 
Comparative evaluation of diepoxybutane sensitivity and cell cycle blockage in the diagnosis of Fanconi anemia
Blood
1995
, vol. 
85
 
8
(pg. 
2233
-
2237
)
22
Shimamura
 
A
de Oca
 
RM
Svenson
 
JL
, et al. 
A novel diagnostic screen for defects in the Fanconi anemia pathway
Blood
2002
, vol. 
100
 
13
(pg. 
4649
-
4654
)
23
Lo Ten Foe
 
JR
Kwee
 
ML
Rooimans
 
MA
, et al. 
Somatic mosaicism in Fanconi anemia: molecular basis and clinical significance
Eur J Hum Genet
1997
, vol. 
5
 
3
(pg. 
137
-
148
)
24
Waisfisz
 
Q
Morgan
 
NV
Savino
 
M
, et al. 
Spontaneous functional correction of homozygous fanconi anaemia alleles reveals novel mechanistic basis for reverse mosaicism
Nat Genet
1999
, vol. 
22
 
4
(pg. 
379
-
383
)
25
Gross
 
M
Hanenberg
 
H
Lobitz
 
S
, et al. 
Reverse mosaicism in Fanconi anemia: natural gene therapy via molecular self-correction
Cytogenet Genome Res
2002
, vol. 
98
 
2-3
(pg. 
126
-
135
)
26
Auerbach
 
AD
Diagnosis of fanconi anemia by diepoxybutane analysis
Curr Protoc Hum Genet
2003
 
Chapter 8:Unit 8 7
27
Soulier
 
J
Leblanc
 
T
Larghero
 
J
, et al. 
Detection of somatic mosaicism and classification of Fanconi anemia patients by analysis of the FA/BRCA pathway
Blood
2005
, vol. 
105
 
3
(pg. 
1329
-
1336
)
28
Mankad
 
A
Taniguchi
 
T
Cox
 
B
, et al. 
Natural gene therapy in monozygotic twins with Fanconi anemia
Blood
2006
, vol. 
107
 
8
(pg. 
3084
-
3090
)
29
Pinto
 
FO
Leblanc
 
T
Chamousset
 
D
, et al. 
Diagnosis of Fanconi anemia in patients with bone marrow failure
Haematologica
2009
, vol. 
94
 
4
(pg. 
487
-
495
)
30
Quentin
 
S
Cuccuini
 
W
Ceccaldi
 
R
, et al. 
Myelodysplasia and leukemia of Fanconi anemia are associated with a specific pattern of genomic abnormalities that includes cryptic RUNX1/AML1 lesions
Blood
2011
, vol. 
117
 
15
(pg. 
e161
-
170
)
31
Faivre
 
L
Guardiola
 
P
Lewis
 
C
, et al. 
Association of complementation group and mutation type with clinical outcome in fanconi anemia. European Fanconi Anemia Research Group
Blood
2000
, vol. 
96
 
13
(pg. 
4064
-
4070
)
32
Neveling
 
K
Endt
 
D
Hoehn
 
H
Schindler
 
D
Genotype-phenotype correlations in Fanconi anemia
Mutat Res
2009
, vol. 
668
 
1-2
(pg. 
73
-
91
)
33
Castella
 
M
Pujol
 
R
Callen
 
E
, et al. 
Origin, functional role, and clinical impact of Fanconi anemia FANCA mutations
Blood
2011
, vol. 
117
 
14
(pg. 
3759
-
3769
)
34
Kalb
 
R
Neveling
 
K
Hoehn
 
H
, et al. 
Hypomorphic mutations in the gene encoding a key Fanconi anemia protein, FANCD2, sustain a significant group of FA-D2 patients with severe phenotype
Am J Hum Genet
2007
, vol. 
80
 
5
(pg. 
895
-
910
)
35
Alter
 
BP
Rosenberg
 
PS
Brody
 
LC
Clinical and molecular features associated with biallelic mutations in FANCD1/BRCA2
J Med Genet
2007
, vol. 
44
 
1
(pg. 
1
-
9
)
36
Hirsch
 
B
Shimamura
 
A
Moreau
 
L
, et al. 
Biallelic BRCA2/FANCD1 mutations: association with spontaneous chromosomal instability and solid tumors of childhood
Blood
2004
, vol. 
103
 (pg. 
2554
-
2559
)
37
Wagner
 
JE
Tolar
 
J
Levran
 
O
, et al. 
Germline mutations in BRCA2: shared genetic susceptibility to breast cancer, early onset leukemia, and Fanconi anemia
Blood
2004
, vol. 
103
 
8
(pg. 
3226
-
3229
)
38
Lensch
 
MW
Rathbun
 
RK
Olson
 
SB
Jones
 
GR
Bagby
 
GC
Selective pressure as an essential force in molecular evolution of myeloid leukemic clones: a view from the window of Fanconi anemia
Leukemia
1999
, vol. 
13
 
11
(pg. 
1784
-
1789
)
39
Li
 
X
Le Beau
 
MM
Ciccone
 
S
, et al. 
Ex vivo culture of Fancc-/- stem/progenitor cells predisposes cells to undergo apoptosis, and surviving stem/progenitor cells display cytogenetic abnormalities and an increased risk of malignancy
Blood
2005
, vol. 
105
 
9
(pg. 
3465
-
3471
)
40
Bagby
 
GC
Meyers
 
G
Bone marrow failure as a risk factor for clonal evolution: prospects for leukemia prevention
Hematology Am Soc Hematol Educ Program
2007
(pg. 
40
-
46
)
41
Ikeda
 
H
Matsushita
 
M
Waisfisz
 
Q
, et al. 
Genetic reversion in an acute myelogenous leukemia cell line from a Fanconi anemia patient with biallelic mutations in BRCA2
Cancer Res
2003
, vol. 
63
 
10
(pg. 
2688
-
2694
)
42
Ceccaldi
 
R
Briot
 
D
Larghero
 
J
, et al. 
Spontaneous abrogation of the G DNA damage checkpoint has clinical benefits but promotes leukemogenesis in Fanconi anemia patients
J Clin Invest
2011
, vol. 
121
 
1
(pg. 
184
-
194
)
43
Li
 
J
Sejas
 
DP
Zhang
 
X
, et al. 
TNF-alpha induces leukemic clonal evolution ex vivo in Fanconi anemia group C murine stem cells
J Clin Invest
2007
, vol. 
117
 
11
(pg. 
3283
-
3295
)
44
Butturini
 
A
Gale
 
RP
Verlander
 
PC
Adler-Brecher
 
B
Gillio
 
AP
Auerbach
 
AD
Hematologic abnormalities in Fanconi anemia: an International Fanconi Anemia Registry study
Blood
1994
, vol. 
84
 
5
(pg. 
1650
-
1655
)
45
Cioc
 
AM
Wagner
 
JE
MacMillan
 
ML
DeFor
 
T
Hirsch
 
B
Diagnosis of myelodysplastic syndrome among a cohort of 119 patients with fanconi anemia: morphologic and cytogenetic characteristics
Am J Clin Pathol
2010
, vol. 
133
 
1
(pg. 
92
-
100
)
46
Alter
 
BP
Caruso
 
JP
Drachtman
 
RA
Uchida
 
T
Velagaleti
 
GV
Elghetany
 
MT
Fanconi anemia: myelodysplasia as a predictor of outcome
Cancer Genet Cytogenet
2000
, vol. 
117
 
2
(pg. 
125
-
131
)
47
Maarek
 
O
Jonveaux
 
P
Le Coniat
 
M
Derre
 
J
Berger
 
R
Fanconi anemia and bone marrow clonal chromosome abnormalities
Leukemia
1996
, vol. 
10
 
11
(pg. 
1700
-
1704
)
48
Tönnies
 
H
Huber
 
S
Kuhl
 
JS
Gerlach
 
A
Ebell
 
W
Neitzel
 
H
Clonal chromosomal aberrations in bone marrow cells of Fanconi anemia patients: gains of the chromosomal segment 3q26q29 as an adverse risk factor
Blood
2003
, vol. 
101
 
10
(pg. 
3872
-
3874
)
49
O'Keefe
 
C
McDevitt
 
MA
Maciejewski
 
JP
Copy neutral loss of heterozygosity: a novel chromosomal lesion in myeloid malignancies
Blood
2010
, vol. 
115
 
14
(pg. 
2731
-
2739
)