Infant B-cell acute lymphoblastic leukemia (B-ALL) accounts for 10% of childhood ALL. The genetic hallmark of most infant B-ALL is chromosomal rearrangements of the mixed-lineage leukemia (MLL) gene. Despite improvement in the clinical management and survival (∼85-90%) of childhood B-ALL, the outcome of infants with MLL-rearranged (MLL-r) B-ALL remains dismal, with overall survival <35%. Among MLL-r infant B-ALL, t(4;11)+ patients harboring the fusion MLL-AF4 (MA4) display a particularly poor prognosis and a pro-B/mixed phenotype. Studies in monozygotic twins and archived blood spots have provided compelling evidence of a single cell of prenatal origin as the target for MA4 fusion, explaining the brief leukemia latency. Despite its aggressiveness and short latency, current progress on its etiology, pathogenesis, and cellular origin is limited as evidenced by the lack of mouse/human models recapitulating the disease phenotype/latency. We propose this is because infant cancer is from an etiologic and pathogenesis standpoint distinct from adult cancer and should be seen as a developmental disease. This is supported by whole-genome sequencing studies suggesting that opposite to the view of cancer as a “multiple-and-sequential-hit” model, t(4;11) alone might be sufficient to spawn leukemia. The stable genome of these patients suggests that, in infant developmental cancer, one “big-hit” might be sufficient for overt disease and supports a key contribution of epigenetics and a prenatal cell of origin during a critical developmental window of stem cell vulnerability in the leukemia pathogenesis. Here, we revisit the biology of t(4;11)+ infant B-ALL with an emphasis on its origin, genetics, and disease models.

B-cell acute lymphoblastic leukemia (B-ALL) is the most frequent cancer in the childhood.1  B-ALL is characterized by an uncontrolled expansion of immature B-cell precursors within the bone marrow (BM).2,3  There are different subtypes of B-ALL attending to both the specific differentiation stage where B-cell precursors are stalled and to the cytogenetic/molecular diagnosis.4,5  Over the last decades, improved understanding of the disease biology, precise risk stratification, improved supportive care, and personalized chemotherapy has increased the cure rate of childhood B-ALL close to 90%.3,6  However, leukemia in infants (<1 year of age) is rare but captures a lot of interest due to its aggressive clinical presentation in a uniquely vulnerable host. Of special interest is the infant B-ALL mediated by mixed-lineage leukemia rearrangement (MLL-r), which possesses unique clinical and biological features, representing an outlier B-ALL with unfavorable prognosis.

Germline MLL (also known as KMT2A) on chromosome 11q23 is required for normal hematopoiesis and the expression of the HOXA cluster gene.7-11  The MLL gene is rearranged most often in de novo infant leukemia and in adult secondary therapy-related leukemia.12-14  MLL-r functions as the initiating, and perhaps as the sole driving, oncogenic event by dysregulating epigenetic/transcriptional programs.15,16  The most common rearrangement is a reciprocal translocation between MLL and a partner gene resulting in a chimeric protein composed of the N terminus domain of MLL and C terminus domain of the partner gene.17,18  Many partners have been identified in MLL-r leukemia, the most common being AF4 (also known as AFF1), AF9, and ENL.19 

A biologically and clinically intriguing MLL-r is MLL-AF4 (MA4, also referred as KMT2A-AFF1), which results from t(4;11)(q21;q23) and is the hallmark genetic abnormality of infant t(4;11) pro-B/mixed B-ALL. Infant MA4+ B-ALL is associated with a dismal prognosis,20,21  therapy refractoriness, central nervous system (CNS) infiltration, and short latency.22  Our understanding of the disease pathogenesis, oncogenic insults, and cellular origin is very limited. We speculate this is largely because infant cancer is, from an etiologic and pathogenesis standpoint, very distinct to adult cancer, and it should be studied as a developmental disease.20,23  In fact, compelling evidence of a single cell of prenatal origin as the target cell for MA4 fusion has been extensively reported, with important implications in the way we understand and study developmental cancer.20,23  Despite its aggressiveness, current disease models have failed to faithfully recapitulate the disease, which may be explained from a developmental angle because some developmental cues and the prenatal nature of the target cell remain elusive. In addition, another controversial question is whether MA4 fusion alone functions as a single “big-hit” sufficient to induce an overt B-ALL.24  Whole-genome sequencing (WGS) studies reported a silent mutational landscape in MLL-r infant B-ALL,25  suggesting that a single driver mutation (MLL-r) suffices to spawn this aggressive B-ALL, thus contradicting the conventional dogma of multiple and sequential genetic changes needed for leukemia initiation. Here, we revisit the biology of normal MLL and MA4+ infant B-ALL, with an emphasis on its cellular origin, genetics, and disease models.

Infant MA4+ B-ALL is among the most vexing clinical problems in pediatric hematology/oncology. The incidence of MA4+ B-ALL in infant is between 1 and 2 logs lower than in children aged 1 to 14 years old.1  Phenotypically, MA4+ B-ALL is a CD34+CD19+ pro-B/mixed leukemia lacking CD10 and coexpressing aberrant myeloid markers such as CD15 and CD65 (Table 1).26  The majority of MA4+ B-ALLs express the neural glial marker 2 (NG2).27  Although it might be associated with the high propensity of this leukemia to infiltrate the CNS, its exact contribution to disease evolution remains a mystery.28-30 

Table 1

Main clinico-biological features of t(4;11) MA4+ B-ALL

CharacteristicComment
Incidence* Estimated 1 case/1 million newborns 
Phenotype Pro-B/mixed: CD34+CD19+CD10 with expression of the myeloid markers CD15 and CD65 and NG2. 
EFS 5-year EFS is 34% for Children’s Cancer Group protocol CCG-1953 
4-year EFS is 37% for Interfant-99 protocol 
State of the art treatment Initial glucocorticoid-based treatment (prednisone, l-asparaginase), followed by induction (dexamethasone, cytarabine, vincristine, daunorubicin, and mitoxantrone) and by myeloid-like (cytarabine, daunorubicin/mitoxantrone, and etoposide) or lymphoid-like consolidation (cyclophosphamide, cytarabine, and 6-mercaptopurine) 
CNS involvement Common: up to 50% accumulative at diagnosis and during evolution 
Age at diagnosis <1 year 
Negative prognostic factors High WBC counts (>300 000 WBC/µL) 
Age < 6 months 
CNS infiltration 
Early poor response to prednisone 
RAS mutations 
High expression level of FLT3 
Low expression level of HoxA cluster 
CharacteristicComment
Incidence* Estimated 1 case/1 million newborns 
Phenotype Pro-B/mixed: CD34+CD19+CD10 with expression of the myeloid markers CD15 and CD65 and NG2. 
EFS 5-year EFS is 34% for Children’s Cancer Group protocol CCG-1953 
4-year EFS is 37% for Interfant-99 protocol 
State of the art treatment Initial glucocorticoid-based treatment (prednisone, l-asparaginase), followed by induction (dexamethasone, cytarabine, vincristine, daunorubicin, and mitoxantrone) and by myeloid-like (cytarabine, daunorubicin/mitoxantrone, and etoposide) or lymphoid-like consolidation (cyclophosphamide, cytarabine, and 6-mercaptopurine) 
CNS involvement Common: up to 50% accumulative at diagnosis and during evolution 
Age at diagnosis <1 year 
Negative prognostic factors High WBC counts (>300 000 WBC/µL) 
Age < 6 months 
CNS infiltration 
Early poor response to prednisone 
RAS mutations 
High expression level of FLT3 
Low expression level of HoxA cluster 

EFS, event-free survival; FLT3, fms-like tyrosine kinase 3; RAS, rat sarcoma; WBC, white blood cell.

*

Estimated in Europe/United States.

Based on Interfant protocol.

In infant B-ALL, MLL-r is associated with poorer outcome. In the Children's Cancer Group protocol CCG-1953, the 5-year event-free survival (EFS) for MLL-r infants was 34% compared with 60% with germline MLL.31  In the Interfant protocol, the 4-year EFS in MLL-r and germline MLL infants was 37% and 74%, respectively (Table 1).31  Conversely, in infant acute myeloid leukemia (AML), MLL-r is not a significant risk factor. The standard care for infant MA4+ B-ALL includes intensive multiagent chemotherapy to induce remission, followed by consolidation chemotherapy (for favorable prognosis) or allogeneic hematopoietic stem cell transplantation (for unfavorable prognosis).32-35  (Table 1). The major cooperative groups conducting clinical trials for infant ALL are Interfant (Interfant-06), The Children’s Oncology Group, and the Japanese Pediatric Leukemia Study Group. All have adopted an identical induction strategy based on Interfant-99, and they use prospective risk-stratified approaches that incorporate MLL-r status. The Interfant protocol gives a 7-day single-agent prednisone before intensive induction chemotherapy and is testing whether consolidation with myeloid-like chemotherapy will prove superior to lymphoid-like consolidation in MLL-r infants. The rationale is that infant MA4+ B-ALL may derive from an early hematopoietic precursor with myeloid differentiation potential and may respond better to myeloid chemotherapy regimens. The Children’s Oncology Group is testing FLT3 tyrosine kinase inhibitors for postinduction chemotherapy (Table 1).

Even though infant MLL-r B-ALL displays, by default, an unfavorable prognosis, several additional independent factors have been identified (Table 1). The most important are age and WBC count at diagnosis, with younger infants and higher WBC count associated with poorer outcomes.21,33-37  A poor response to prednisone (≥1000 blasts/μL in peripheral blood on day 8) is also an independent negative prognostic factor.37,38  Furthermore, the presence of CNS infiltration, RAS mutations, high levels of FLT3, and low levels of HOX-A genes have also been reported negative prognostic factors in independent studies.21,39-41  A retrospective multivariate analysis encompassing both clinico-biological parameters with molecular data still needs to be undertaken in MLL-r infant B-ALL to ascertain the independent robustness of these molecular markers.

Given the large variety of MLL fusions diagnosed thus far in infant, childhood, and adult acute leukemia, one may consider the possibility that a first oncogenic hit derives from the particular truncation of the MLL protein, whereas a potential secondary hit may derive from the fused partner protein. The MLL protein is processed by Taspase1,42  resulting in 2 protein fragments (p320/p180) that bind each other, forming a molecular hub for the assembly of a large nuclear complex (Figure 1). Binding proteins are MEN1, LEDGF, GADD34, PP2A, the polymerase associated factor (PAF) elongation complex, a polycomb group complex (BMI1, HPC2, HDAC1/2, and CtBP), CYP33, CREBP, MOF, and the SET domain core proteins (WDR5, RbBP5, ASH2L, and SRY-30).11,43-52  The N-terminal portion of the MLL protein is functionally linked to bind-and-read chromatin signatures, whereas the C-terminal portion of the MLL protein associates to enzymatic functions, namely acetylation and methylation of histone core particles. Thus, the MLL complex binds to promoter regions of active genes, marking these regions by covalent histone modifications.53 

Figure 1

Proposed model for the oncogenic conversion of MLL fusions. (A) The physiologic situation of MLL functions. Taspase1-cleaved MLL is assembled into the holo-complex and binds to target promoter regions. This occurs via the N-terminally bound MEN1/LEDGF protein complex that allows binding to many transcription factors. The PHD domain is able to read histone core particles, whereas the SET domain allows it to write epigenetic signatures (H3K4me2/3). Associated CREBP and MOF are able to acetylate nucleosomes. CYP33 allows switching into the repressor mode by enabling the docking of a Polycomb group complex composed of BMI1, HPC2, CtBP, and several HDACs. This enables the removal of acetyl groups from nucleosomes or transcription factors to shut down gene transcription. (B) In the case of a chromosomal translocation, the intrinsic regulatory mechanism of MLL becomes destroyed. The disrupted MLL portions are fused to protein sequences deriving from a large amount of different partner genes (n > 80). The N-terminal portion of MLL retains the ability to bind MEN1 and LEDGF, and thus, to bind to target promoter regions. Depending on the fusion sequence (AF4, AF5, LAF4, AF9, ENL, AF10), MLL-X fusions may recruit the endogenous AF4 complex that contains P-TEFb and the histone methyltransferases DOT1L, NSD1, and CARM1. This activates gene transcription and results in enhanced epigenetic signatures (H3K79me2/3). The C-terminal portion retains CREBBP and MOF binding capacity, as well as the SET domain. In some cases (AF4, AF5, LAF4), the N-terminal fused protein sequences allow binding to P-TEFb and directly to the largest subunit of RNA polymerase II to enhance the process of transcriptional elongation. In addition, the fused protein sequences still bind NSD1 and DOT1L.

Figure 1

Proposed model for the oncogenic conversion of MLL fusions. (A) The physiologic situation of MLL functions. Taspase1-cleaved MLL is assembled into the holo-complex and binds to target promoter regions. This occurs via the N-terminally bound MEN1/LEDGF protein complex that allows binding to many transcription factors. The PHD domain is able to read histone core particles, whereas the SET domain allows it to write epigenetic signatures (H3K4me2/3). Associated CREBP and MOF are able to acetylate nucleosomes. CYP33 allows switching into the repressor mode by enabling the docking of a Polycomb group complex composed of BMI1, HPC2, CtBP, and several HDACs. This enables the removal of acetyl groups from nucleosomes or transcription factors to shut down gene transcription. (B) In the case of a chromosomal translocation, the intrinsic regulatory mechanism of MLL becomes destroyed. The disrupted MLL portions are fused to protein sequences deriving from a large amount of different partner genes (n > 80). The N-terminal portion of MLL retains the ability to bind MEN1 and LEDGF, and thus, to bind to target promoter regions. Depending on the fusion sequence (AF4, AF5, LAF4, AF9, ENL, AF10), MLL-X fusions may recruit the endogenous AF4 complex that contains P-TEFb and the histone methyltransferases DOT1L, NSD1, and CARM1. This activates gene transcription and results in enhanced epigenetic signatures (H3K79me2/3). The C-terminal portion retains CREBBP and MOF binding capacity, as well as the SET domain. In some cases (AF4, AF5, LAF4), the N-terminal fused protein sequences allow binding to P-TEFb and directly to the largest subunit of RNA polymerase II to enhance the process of transcriptional elongation. In addition, the fused protein sequences still bind NSD1 and DOT1L.

Close modal

Near the center of the MLL protein is the plant homeodomain (PHD) domain. This region is composed by PHD1-3 subdomain, a bromodomain (BD), and PHD4 subdomain. The PHD domain exhibits 2 normal PHD subdomain structures: PHD1/2 and PHD3/4. The PHD1-3 subdomains are followed by a BD necessary to stabilize the structure of the PHD3/4 domain and has no histone-acetyl reading function in the MLL protein. The PHD3/4 subdomain is required to read H3K4me2/3 signatures within the chromatin. However, when the PHD3/4 subdomain binds to CYP33/PPIE, a conformational change is catalyzed.54,55  As long as PHD3/4 subdomain is docked via a single protein helix to BD, it exhibits its reader function for nucleosomal H3K4 methylation signatures. Isomerization via CYP33/PPIE allows disconnection of PHD3/4 subdomain from the BD and interaction with the BMI1/HPC2/HDAC1-2/CtBP complex that then becomes enabled to bind to the methyl-DNA binding domain (MBD). Binding of MLL to this polycomb-group complex converts MLL into a transcriptional repressor. This defines the CYP33/PPIE isomerase as a molecular switch that triggers the MLL complex in 2 different modes of action: transcriptional activator or repressor. Nothing is known about the precise details of this molecular switch mechanism, but it is likely that it depends on the promoter context. This MLL switch is responsible for the effects on gene transcription. For instance, when Mll-knockout cells are transcriptionally profiled and compared with their isogenic wild-type cells, 66% of the differentially expressed genes become upregulated and only 33% are downregulated in the knockout cells.56  Therefore, genetic insults affecting bona fide functions of MLL may underlie leukemia initiation (Figure 1A). Compromising a domain that is responsible to perform binary decisions linked to gene transcription and epigenetic reading-writing may well explain the profound biological effect.

What does actually happen when a chromosomal translocation occurs at the MLL gene locus? Chromosomal rearrangements usually separate the MBD from the PHD domain (Figure 1B), thereby destroying the intrinsic control mechanism of the MLL protein. Even the binding of CYP33 to the PHD domain is impaired, at least when the chromosomal breakpoint localizes within MLL intron 11.57  Consequently, chromatin reading and writing functions now become independent of each other, and both portions of MLL become constitutively active, regardless of their fused protein sequences. The MLL-X fusions still bind via MEN1/LEDGF to chromatin and transcription factors in promoter regions and the PAF complex but are disabled to exert any inhibitory function. The reciprocal X-MLL fusion proteins retain the PHD3/4 chromatin reader domain, the CREBBP/MOF binding domain, and the SET domain complex. If CYP33/PPIE binds to the PHD domain of X-MLL fusions, the binding of the Polycomb complex is disabled due to the absence of MBD. This was nicely demonstrated by experiments where the PHD domain was fused to existing MLL-X fusion proteins, being sufficient to eliminate the oncogenic properties of those artificial MLL fusions, because the repressing functions are now exerted by the fused PHD domain via binding to the BMI-1 repressor complex.58,59  This model proposes that the biological functions deriving from the N- and C-terminal fused protein sequences exert only an accessory gain of function such as changing the protein interactome.

Epidemiologic and genetic studies have suggested that infant MLL-r leukemia may result from transplacental exposures during the gestation to DNA topoisomerase-II inhibitors, such as bioflavonoids.60,61  Bioflavonoid-rich dietary habits of the pregnant mothers have been linked to in utero MLL breaks.62,63  However, the wide exposure of the population to bioflavonoids contrasts with the low incidence of this infant leukemia, suggesting that the probability that DNA topoisomerase-II inhibitors hit a candidate leukemia-initiating cell (LIC) is extremely low. Importantly, etoposide, a DNA topoisomerase-II inhibitor that is used widespread in chemotherapy cocktails, can induce MLL-r in different cell types61 : in embryonic stem cells (ESCs),64,65  fetal liver-derived CD34+ hematopoietic stem cells (HSCs)66  and cord blood (CB)-derived CD34+ HSCs. Clinically, ∼5% to 10% of etoposide-treated patients develop a therapy-related leukemia characterized by the presence of MLL-r.14 

The driving genetic alterations underlying infant leukemias originate prenatally during embryonic-fetal development.67  Seminal studies in identical monozygotic twins with concordant leukemias first demonstrated the presence of a unique and common MLL-r in the leukemic cells from both siblings, which unequivocally shows that the leukemogenic event arises in utero in one of the twins and then is propagated through blood circulation within the single, monochorionic shared placenta.68  The concordance rate for this leukemia in both twins is close to ∼100%. MA4 fusion has also been found and expressed in BM stromal cells (BMSCs) from t(4;11)+ B-ALL infants, suggesting that the MLL fusion may arise in a very early prehematopoietic/mesodermal precursor.69  Large-scale CB screenings also demonstrated that chromosomal translocations can be generated during fetal development.70  Leukemia fusion genes were present in CB from healthy newborns, but overt leukemia only occurred in 1% of these newborns, suggesting that additional mutations are needed. In nontwinned infant ALL, the prenatal origin of MA4 was demonstrated by a retrospective screening of blood taken at birth on neonatal blood spots from children with B-ALL.71  Altogether, these studies established that t(4;11) arises in utero during fetal hematopoiesis.

The extremely short latency of MA4+ infant B-ALL questions whether secondary mutations are required at all to develop overt leukemia. The 2-hit model postulates that cancer results from the accumulation of DNA mutations. In the case of pediatric B-ALL, the 2-hit hypothesis postulates that the prenatal chromosomal translocation is followed by a secondary prenatal or early postnatal genetic mutation. Because infant B-ALL has a shorter natural history than common B-cell precursor ALL (cALL), it is plausible that all the necessary genetic alterations occur prenatally (Figure 2).72 

Figure 2

Two-hit cancer model in infant t(4;11)+ B-ALL. MA4 fusion is the first and driver oncogenic event. The very short latency of the disease indicates that secondary cooperating hits, if required, are expected to arise prenatally or very early after birth.

Figure 2

Two-hit cancer model in infant t(4;11)+ B-ALL. MA4 fusion is the first and driver oncogenic event. The very short latency of the disease indicates that secondary cooperating hits, if required, are expected to arise prenatally or very early after birth.

Close modal

Although extensive efforts have been directed to decipher secondary genetic alterations in MA4+ B-ALL, clinical and experimental data remain controversial. Tamai et al showed that ectopic expression of MA4 in murine HSCs is sufficient to cause B-ALL in mice and that cooperating KRAS mutations accelerate leukemic transformation.73  However, RAS mutations are subclonal, only present in ∼25% of patients, and are partially lost at relapse, suggesting they are not tumor drivers.39,74  Preliminary data from our laboratory reveal that activated KRAS cooperates with MA4 to promote CNS infiltration and extramedullary engraftment (spleen and peripheral blood) of CB-CD34+ HSPCs but is not sufficient to initiate leukemia.

The FLT3 receptor is required for normal lymphopoiesis and is expressed in almost all AML/ALLs. Gene expression profiling showed that FLT3 is highly expressed in MLL-r B-ALL,75  leading to the characterization of FLT3 mutations as potential secondary cooperating events. However, the occurrence of FLT3 mutations in MA4+ pro-B ALL remains controversial. Some studies have shown that MLL-r B-ALL harbors FLT3 mutations in 3% to 21% of cases,76-78  whereas others have shown that FLT3 mutations are not present in MLL-r B-ALL.21,40,79-81  Given that FLT3 is consistently highly expressed in MLL-r B-ALL, increased transcriptional expression of FLT3 rather than activating mutations could represent an important hallmark of MLL-r B-ALL.21,40  We recently analyzed the prognostic significance of FLT3 mutations and expression in MA4+ and MLL-germline ALL21  and found that MA4+ B-ALL exhibited higher FLT3 expression levels than normal BM, supporting that aberrantly increased transcription of FLT3, rather than activating FLT3 mutations, contributes to the pathogenesis of MA4+ B-ALL. This is supported by studies reporting FLT3 as a direct transcriptional target of MA4.82  Similar to that reported for RAS mutations,39  high FLT3 expression is an independent prognostic factor in MA4+ B-ALL patients (but not in pediatric cALL) and is associated with lower overall survival and EFS. Unfortunately, owing to the low incidence of MLL-r infant leukemias, no clear correlation between FLT3 levels and specific MLL fusions could be established.

Increasing evidence points to AF4-MLL, the reciprocal product of MA4, as a driver oncogenic event in t(4;11)+ B-ALL.83  AF4-MLL–transduced mouse HSCs developed pro-B-ALL, whereas cotransduction with MA4 and AF4-MLL resulted in mixed lineage leukemia, suggesting that the expression of the AF4-MLL protein may induce B-ALL even in the absence of MA4. Studies from Wilkinson et al demonstrated that RUNX1 is directly activated by MA4 and the RUNX1 protein interacts with the AF4-MLL chimeric protein, suggesting how these onco-proteins could cooperate at the molecular level.84 

Dysregulated immune responses to common infections have also emerged as cooperating events underlying the etiology of childhood leukemia.72  Delayed exposure of children to common pathogens may cause exacerbated T-cell responses when these infections occur later on in life.85,86  This untimely and excessive inflammatory response abolishes normal hematopoiesis promoting selective expansion of a preleukemic clone, resulting in stochastic or microenvironment-derived cooperating drivers toward overt leukemia.87,88  The delayed infection hypothesis may contribute to the etiology of pediatric leukemias with a longer latency, but it is unlikely it underlies the etiology of MA4+ infant B-ALL diagnosed during early life or even at birth, because newborns have not had natural time for exposures to infection.

Whether MA4 is sufficient for leukemogenesis or requires cooperating mutations is still somehow an open debate. Genome-wide sequencing analysis revealed genetic alterations in genes involved in B-cell development/differentiation in pediatric cALL leukemias.89  In contrast, Bardini et al described a stable genome for MA4+ infant B-ALL patients as demonstrated by the absence of copy number alterations analyzed by single nucleotide polymorphism arrays.80,81  Shortly after, Dobbins et al performed WGS of leukemic blasts at diagnosis from 3 infant MA4+ B-ALLs, and no recurrent somatic mutations were found, and the mutational landscape of patients was largely silent.90  Very recently, the whole-genome mutational landscape of MLL-r B-ALLs has been revisited in a larger patient cohort as part of the Pediatric Cancer Genome Project at St. Jude Children’s Research Hospital.25  This seminal work demonstrates that infant MLL-r B-ALL has one of the lowest frequencies of somatic mutations of any sequenced cancer, with the predominant leukemic clone carrying 1.3 nonsilent mutations. The only alterations recurrently (47% of cases) detected were mutations in the kinase-phosphatidylinositol 3-kinase-RAS pathway, and these mutations were subclonal. In contrast, noninfant MLL-r B-ALL patients displayed on average 6.5 nonsilent mutations per case, mainly affecting epigenetic regulators. Importantly, noninherited germline mutations acquired during hematopoietic development cannot be ruled out as oncogenic drivers because they have not been analyzed in the available WGS studies.

The silent mutational landscape observed in infant MLL-r B-ALL coupled to the higher number of mutations emerging in pediatric (noninfants) MLL-r B-ALL supports the developmental origin of infant MLL-r B-ALL and establishes that very few genetic alterations are required for overt leukemia. This silent mutational landscape argues that the MA4 may function as a single big hit sufficient to induce overt, short latency, and aggressive B-ALL.25  This contrasts with the 2-hit model widely proposed for other childhood leukemias. For instance, TEL-AML1–expressing t(12;21) pre-B-ALL requires additional postnatal cooperating oncogenic hits such as deletion of the germline TEL allele91  or RAG-mediated deletions.92  If MA4 is the sole initiating event, the MA4-mediated transformation would then rely on alternative epigenetic cooperating lesions occurring on a critical developmentally early window of stem cell vulnerability, explaining the low prevalence of this entity. Interestingly, Chuk et al93  reported the clearance of preleukemic MLL-AF4 cells from a healthy child. Similarly, Maia et al94  described a child where the MLL fusion was detected at birth, but there was a latency of 6 years before overt leukemia was observed. These reports present evidence against the 1-hit hypothesis (specifically an MLL-ENL fusion) by proposing the existence of MLL-r preleukemic clones that eventually disappear in the absence of cooperating proliferation/survival-promoting secondary oncogenic hits.

Mutational and epigenetic mechanisms are expected to affect gene expression, allelic imbalance, and key components of the spliceosome with the subsequent generation of alternative splicing variants commonly linked to cancer.95  DNA-Seq allows identification of potentially oncogenic silent and nonsilent single nucleotide variants and is also informative in dissecting intraclonal mutation heterogeneity.96  However, to date, RNA-Seq and whole genome epigenetic studies have not been performed on this leukemia to unravel whether epigenetic rather than genetic alterations are the missing cooperating oncogenic hits. Therefore, a multilayer-omics analysis coupled to computational integration of genome-wide DNA- and RNA-Seq and genome-wide DNA methylation of infant MA4+ B-ALL will provide a unique molecular-genetic signature of the disease onset/evolution at the genome, epigenome, and transcriptome level.97,98 

The silent mutational landscape found in infant MA4+ B-ALL does not explain the rapid onset and aggressive evolution of the disease.24  MLL is an H3K4 histone methyltransferase,10,11  and leukemia transformation by MLL fusions requires H3 lysine 79 (H3K79) methyltransferase DOT1L, which is recruited to the MLL fusion transcriptional complex53,99  (Figure 1). Epigenetic mechanisms including DNA methylation and histone modification by methylation or acetylation tightly regulate gene expression during early mammalian development and hematopoiesis.100,101  Thus, we envision a key contribution of epigenetic remodeling in the pathogenesis of this B-ALL.53  However, little information is available on epigenetic dysregulation of MA4+ B-ALL. Armstrong et al demonstrated that H3K79 methylation profiles define both murine and human MA4+ B-ALL.98,102  They showed that MA4+ B-ALLs could be distinguished from other ALLs by their H3K79 profiles. Suppression of the DOT1L inhibited expression of MA4 target genes, demonstrating that H3K79 methylation is a distinguishing feature of MA4+ B-ALLs and is key for maintenance of MA4-driven gene expression.82,103 

Ikawa et al104  reported dense methylation of regulatory regions of the CD10 gene promoter, suggesting that methylated transcription factor binding sites contribute to CD10 silencing as epigenetic mechanisms underlying the pro-B (CD10 neg) phenotype. In cancer cells, enhanced promoter methylation is typically accompanied by global loss of methylation in nonpromoter regions of the genome. However, Stumpel et al reported that MLL-r infant B-ALL cells display a global hypermethylated genomic state, both at promoter and nonpromoter regions.105  Because global hypomethylation usually leads to genomic instability linked to cancer development, this study might explain the global genomic stability/silent mutational landscape found in MA4+ infant B-ALLs and the remarkable sensitivity of MLL-r cells to demethylating agents.106 

Further studies from Stumpel et al focused on genome-wide cytosine guanine dinucleotide island methylation and promoter methylation.107  These studies revealed that different subsets of B-ALL and different types of MLL translocations are associated with distinct patterns of DNA methylation, and the degree of DNA methylation influences clinical outcome, identifying subgroups of MLL-r infant B-ALL patients that may particularly benefit from therapeutic strategies based on demethylating drugs.107,108  Cutting edge epigenetic whole-genome approaches will shortly be applied to MA4+ B-ALLs, providing the overall genome methylation picture that should be integrated with the available whole-genome DNA-Seq data to gain insights on the genomic and epigenetic aberrations underlying MA4+ B-ALLs.

No experimental model has thus far faithfully recapitulated MA4+ pro-B-ALL latency/phenotype (Table 2). Chen et al produced MA4 knock-in mice by homologous recombination in ESCs, but these mice developed mixed lymphoid/myeloid hyperplasia and B-cell lymphomas.109  Similarly, Metzler et al used the invertor conditional technology to create a mouse model in which a floxed AF4 cDNA was knocked into MLL in an opposite orientation to transcription. Then, Rag1, Lck, and CD19-Cre expression were used to drive MA4 expression in B/T progenitors, T cells, or B cells, respectively. These mice developed mature B-cell neoplasias. Although the latency varied slightly depending on the Cre line used, in all cases it was >300 days.110  Also, Kristov et al created a mouse in which conditional expression of MA4 using interferon-inducible Mx1-Cre line resulted in cALL or AML.98  Finally, Tamai and Inokuchi established a MA4 transgenic mice, but the phenotype obtained was lymphoblastic leukemia or lymphoma111  (Table 2).

Table 2

Summary of current mouse models for MA4+ B-ALL

StrategyPhenotype*LatencyTissue-specific CreReference
Constitutive Mll-AF4 knock-in mice Myeloproliferative/follicular B leukemia 520 days (17 months) NA Chen et al109  
Mll-AF4 invertor mice B-cell lineage neoplasias 317-466 days Rag-Cre Metzler et al110  
416-472 days Lck-Cre 
460-475 days CD19-Cre  
Conditional Mll-AF4 knock-in mice B-precursor ALL and AML 131 days Mx1-Cre Krivtsov et al98  
Transplant of AF4-MLL-transduced murine HSPCs AF4-MLL: pro-B ALL (63%); B/T biphenotypic (37%) AF4-MLL: 233 days NA Bursen et al83  
Double: 266 days 
Double: B/T biphenotypic (67%); pro-B ALL (33%) 
MLL-AF4 transgenic mice Lymphoblastic leukemia or lymphoma 12 months NA Tamai et al111  
StrategyPhenotype*LatencyTissue-specific CreReference
Constitutive Mll-AF4 knock-in mice Myeloproliferative/follicular B leukemia 520 days (17 months) NA Chen et al109  
Mll-AF4 invertor mice B-cell lineage neoplasias 317-466 days Rag-Cre Metzler et al110  
416-472 days Lck-Cre 
460-475 days CD19-Cre  
Conditional Mll-AF4 knock-in mice B-precursor ALL and AML 131 days Mx1-Cre Krivtsov et al98  
Transplant of AF4-MLL-transduced murine HSPCs AF4-MLL: pro-B ALL (63%); B/T biphenotypic (37%) AF4-MLL: 233 days NA Bursen et al83  
Double: 266 days 
Double: B/T biphenotypic (67%); pro-B ALL (33%) 
MLL-AF4 transgenic mice Lymphoblastic leukemia or lymphoma 12 months NA Tamai et al111  

NA, not applicable.

*

Disease phenotype observed differs from pro-B ALL (CD10).

Latency, defined as time to leukemia development, is always very protracted.

Table 3

Summary of current human models for MA4+ B-ALL

Ontogeny stageCell typePhenotypeLeukemic cooperation with FLT3/RAS* mutantsReference
Fetal-HSPC* NA — NA — 
Embryonic HSPC hESC-derived hematopoietic cells Enhances early hemato-endothelial specification No 117  
Skew toward endothelial vs hematopoietic fate 
118  
Neonatal HSPC CB-derived CD34+ HSPCs Proliferation coupled to survival advantage No 116  
Enhanced engraftment and clonogenic potential 
119  
Ontogeny stageCell typePhenotypeLeukemic cooperation with FLT3/RAS* mutantsReference
Fetal-HSPC* NA — NA — 
Embryonic HSPC hESC-derived hematopoietic cells Enhances early hemato-endothelial specification No 117  
Skew toward endothelial vs hematopoietic fate 
118  
Neonatal HSPC CB-derived CD34+ HSPCs Proliferation coupled to survival advantage No 116  
Enhanced engraftment and clonogenic potential 
119  
*

Unpublished preliminary data.

Bursen et al83  showed that retroviral expression of the reciprocal protein AF4-MLL in murine HSCs followed by transplantation induced pro-B ALL in transplanted mice with a latency of 233 days. A pro-B phenotype was not observed in 100% cases, and B/T biphenotypic acute leukemia (in the AF4-MLL group) and mixed lineage leukemia cases (in the MA4+AF4-MLL group) were also detected. Leukemia development was faster when MA4- and AF4-MLL–cotransduced cells were transplanted (Table 2). However, transplantation of MA4-transduced HSCs did not result in leukemia. Therefore, AF4-MLL seems capable of inducing ALL in mice without the requirement of MA4, suggesting that the reciprocal protein is the initiating event of MA4+ leukemia.83  However, controversy about the role of AF4-MLL as an oncogenic initiating event exists because AF4-MLL is not expressed in one third of the patients112  and by studies on t(4;11) cell lines showing that they display addiction to MA4 but not to AF4-MLL.113,114  However, a transient silent interfering RNA-mediated knockdown of AF4-MLL was not sufficient to downregulate the AF4-MLL fusion protein, likely due to the stability of the AF4-MLL protein. In addition, AF4-MLL was shown to cooperate with RUNX1 to mediate the leukemogenic process.84,115  Together, these data suggest that one cannot draw definitive conclusions from these conflicting reports and that further studies using mouse models and primary patient samples should address to what extent AF4-MLL drives leukemia initiation or maintenance.

Other modeling attempts have used prenatal/neonatal human stem cells. Montes et al explored the effect of MA4 expression in 2 types of human stem cells: CB-derived CD34+ HSPCs and human embryonic stem cells (hESCs). In vivo, MA4 increased the in vivo repopulation ability of CB-derived CD34+ progenitors and in vitro MA4 increased the clonogenic potential and the proliferation of CD34+ HSPCs. However, MA4 did not induce leukemogenesis.116  Similarly, MA4 enhanced the specification of hemogenic endothelium from hESCs but strongly impaired further hematopoietic commitment, being also insufficient for leukemogenesis in this context.117  Bueno et al also found that the cooperation between MA4 and FLT3 in hESCs and CD34+ HSPCs did not result in leukemia. In hESCs, enforced expression of FLT3-TKD/FLT3-WT abolished hematopoietic specification,118  whereas it conveyed a transient overexpansion but did not suffice to immortalize/transform MA4-expressing CB-CD34+ HSPCs119  (Table 2).

The nature of the prenatal cell initially transformed by MA4 in utero is unknown. Fluorescence in situ hybridization studies indicated that the fusion gene is present in the human primitive CD34+CD19 cells.120  Furthermore, MA4+ pro-B ALL manifests with aberrant bi-phenotypic blasts that coexpress both lymphoid and myeloid markers, with MA4 fusion being present in both lymphoid and myeloid lineages, suggesting that it may originate from immature lympho-myeloid stem/progenitor cells (HSPCs). In line with genetic evidence showing that MA4 originates in utero, Greaves and Wiemels hypothesized that this leukemia could originate from lymphoid-monocytic progenitors active in the fetal liver.67  Alternatively, because BMSCs from primary MA4+ pro-B infant patients harbor and express the MA4 fusion gene, we suggested that MA4 might arise in a population of mesodermal prehematopoietic precursors.69  This observation was restricted to MA4 because other leukemic fusion genes were absent in BMSCs. However, a similar study by Shalapour et al detected MLL-ENL and TEL-AML1 in the BMSC fraction. A variable fraction of translocated-positive BM-MSCs was found among patients when using immune fluorescence in situ hybridization.121  The association between blasts and stroma is not completely elucidated, and further investigation is required to better understand the leukemia relapse origin and the role of the microenviroment in this leukemia.

Although the identity of the cell of origin has not been addressed much in the murine models, it has been investigated in the human system using ontogenically early human stem cells such as hESC-blood derivatives and CB-HSPCs. Studies from our laboratory addressed whether these populations constitute target cells for leukemia initiation.116-119  Lentiviral-mediated expression of MA4 failed to transform either hESC blood derivatives or CB-derived CD34+ HSPCs.116,117  Other sources of developmentally early stem cells which might be potential cells of origin are fetal liver and aorta-gonad-mesonephros (AGM) cells. Human fetal liver tissue could be obtained from voluntary interruption of pregnancy, but whether human fetal liver-derived CD34+ HSPCs represent the cell of origin in this leukemia has yet to be addressed. Although conceptually the AGM region could represent the candidate mesodermal prehematopoietic precursor and therefore a possible cell of origin, no AGM-derived leukemia has been reported to date either in the human or the mouse setting. Given that genetic hallmarks in pediatric B-ALL occur in the fetus at which time B-1 B-cell progenitor numbers are high, mouse B1 progenitors have been proposed as another candidate cell-of-origin population. In fact, Montecino-Rodriguez et al showed that B-cell receptor-Abelson nonreceptor tyrosine kinase (ABL)+ B-ALL can initiate in fetal B-1 progenitors. They showed that mice transplanted with BCR-ABL–transduced fetal liver or BM B-1 progenitors became moribund more rapidly than recipients of BCR-ABL–transduced B-2 progenitors, suggesting B-1 progenitors as the target cell for disease initiation.122 

Conversely, there are no data about the phenotype of the LIC. The LIC may be distinct from the cell in which MA4 has a preleukemic impact and, this in turn, may also differ from the cell in which MA4 first arises. Two recent studies have assessed the LICs in MA4+ pro-B ALL patient samples using in vivo xenotransplantation models. Bardini et al showed that all the LIC potential is within the CD19+ fraction,29  and Aoki et al similarly showed that both CD34+CD38+CD19+ and CD34CD19+ cells contained LICs.123  Interestingly, the data of Aoki et al showed that, in MLL-AF9+ leukemia, the LICs were CD34CD19+. It also remains unknown whether this infant MA4+ B-ALL follows a hierarchical or stochastic cancer model.124  Worth mentioning, Bardini et al29  established that MA4+ infant ALL is composed of a branching subclonal architecture at diagnosis. Some MA4+ clones appear to be quiescent at diagnosis but reactivate and dominate on serial transplantation into immunodeficient mice, whereas other dominant clones at diagnosis become more quiescent, suggesting a dynamic competition between actively proliferating and quiescent subclones. They showed using paired diagnostic and relapse samples that relapses often occur from subclones already present but more quiescent at diagnosis. In sum, much information remains to be gained in both the mouse and human setting about the nature of the embryonic/fetal cell in which MA4 arises and/or exerts its transformation potential, as well as the functional LIC, which fuels leukemia maintenance/progression and therapy relapse. Identification of the cell of origin, LICs, and the subclonal architecture and dynamics in MLL+ infant leukemia should lead to improved therapeutic strategies and provide key insights for the therapy resistance and frequent relapses observed in this group of poor prognosis ALL.

This work was supported by European Research Council grant ERC-2014-CoG-646903 (to P.M.), Instituto de Salud Carlos III/Fondo Europeo de Desarrollo Regional (FEDER) Grant PI14/01191 (to C.B.), Ministerio de Economía y Competitividad (MINECO) grant SAF2013-43065 (to P.M.), The Spanish Association Against Cancer (to P.M. and C.B.), Marie Curie Career Integration grant FP7-PEOPLE-2013-CIG-631171 (to A.S.-P.), The Fundación Inocente Inocente, and The Generalitat de Catalunya grant SGR330 (to P.M.). C.B. and C.P. are supported by Miguel Servet II Contract CPII13/00011 and Ayudas Predoctorales de formación en investigación en salud (PFIS) Scholarship FI12/00468, respectively. P.M. also acknowledges financial support from The Obra Social La Caixa-Fundació Josep Carreras.

This review is dedicated to all the infants diagnosed with MLL leukemia who have participated in our research over the last years.

Contribution: All authors contributed to review preparation by gathering, reviewing and interpreting literature, and assisting with the writing process; R.W.S. and R.M. wrote parts of the review; and A.S.-P. and P.M. conceived the review, wrote the review, and put the text and figures together.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Alejandra Sanjuan-Pla, Josep Carreras Leukemia Research Institute, School of Medicine, University of Barcelona, Casanova 143, 08036 Barcelona, Spain; e-mail: asanjuan@carrerasresearch.org; or Pablo Menéndez, Josep Carreras Leukemia Research Institute, School of Medicine, University of Barcelona, Casanova 143, 08036 Barcelona, Spain; e-mail: pmenendez@carrerasresearch.org.

1
Pui
 
CH
Evans
 
WE
A 50-year journey to cure childhood acute lymphoblastic leukemia.
Semin Hematol
2013
, vol. 
50
 
3
(pg. 
185
-
196
)
2
Bhojwani
 
D
Yang
 
JJ
Pui
 
CH
Biology of childhood acute lymphoblastic leukemia.
Pediatr Clin North Am
2015
, vol. 
62
 
1
(pg. 
47
-
60
)
3
Pui
 
CH
Mullighan
 
CG
Evans
 
WE
Relling
 
MV
Pediatric acute lymphoblastic leukemia: where are we going and how do we get there?
Blood
2012
, vol. 
120
 
6
(pg. 
1165
-
1174
)
4
Ribeiro
 
RC
Pui
 
CH
Prognostic factors in childhood acute lymphoblastic leukemia.
Hematol Pathol
1993
, vol. 
7
 
3
(pg. 
121
-
142
)
5
Pui
 
CH
Evans
 
WE
Genetic abnormalities and drug resistance in acute lymphoblastic leukemia.
Adv Exp Med Biol
1999
, vol. 
457
 (pg. 
383
-
389
)
6
Yeoh
 
AE
Tan
 
D
Li
 
CK
Hori
 
H
Tse
 
E
Pui
 
CH
Asian Oncology Summit 2013
Management of adult and paediatric acute lymphoblastic leukaemia in Asia: resource-stratified guidelines from the Asian Oncology Summit 2013.
Lancet Oncol
2013
, vol. 
14
 
12
(pg. 
e508
-
e523
)
7
Jude
 
CD
Climer
 
L
Xu
 
D
Artinger
 
E
Fisher
 
JK
Ernst
 
P
Unique and independent roles for MLL in adult hematopoietic stem cells and progenitors.
Cell Stem Cell
2007
, vol. 
1
 
3
(pg. 
324
-
337
)
8
McMahon
 
KA
Hiew
 
SY
Hadjur
 
S
, et al. 
Mll has a critical role in fetal and adult hematopoietic stem cell self-renewal.
Cell Stem Cell
2007
, vol. 
1
 
3
(pg. 
338
-
345
)
9
Yu
 
BD
Hess
 
JL
Horning
 
SE
Brown
 
GA
Korsmeyer
 
SJ
Altered Hox expression and segmental identity in Mll-mutant mice.
Nature
1995
, vol. 
378
 
6556
(pg. 
505
-
508
)
10
Milne
 
TA
Briggs
 
SD
Brock
 
HW
, et al. 
MLL targets SET domain methyltransferase activity to Hox gene promoters.
Mol Cell
2002
, vol. 
10
 
5
(pg. 
1107
-
1117
)
11
Nakamura
 
T
Mori
 
T
Tada
 
S
, et al. 
ALL-1 is a histone methyltransferase that assembles a supercomplex of proteins involved in transcriptional regulation.
Mol Cell
2002
, vol. 
10
 
5
(pg. 
1119
-
1128
)
12
Joannides
 
M
Grimwade
 
D
Molecular biology of therapy-related leukaemias.
Clin Transl Oncol
2010
, vol. 
12
 
1
(pg. 
8
-
14
)
13
Sung
 
PA
Libura
 
J
Richardson
 
C
Etoposide and illegitimate DNA double-strand break repair in the generation of MLL translocations: new insights and new questions.
DNA Repair (Amst)
2006
, vol. 
5
 
9-10
(pg. 
1109
-
1118
)
14
Pui
 
CH
Relling
 
MV
Topoisomerase II inhibitor-related acute myeloid leukaemia.
Br J Haematol
2000
, vol. 
109
 
1
(pg. 
13
-
23
)
15
Yip
 
BH
So
 
CW
Mixed lineage leukemia protein in normal and leukemic stem cells.
Exp Biol Med (Maywood)
2013
, vol. 
238
 
3
(pg. 
315
-
323
)
16
Balgobind
 
BV
Zwaan
 
CM
Pieters
 
R
Van den Heuvel-Eibrink
 
MM
The heterogeneity of pediatric MLL-rearranged acute myeloid leukemia.
Leukemia
2011
, vol. 
25
 
8
(pg. 
1239
-
1248
)
17
Marschalek
 
R
Nilson
 
I
Löchner
 
K
, et al. 
The structure of the human ALL-1/MLL/HRX gene.
Leuk Lymphoma
1997
, vol. 
27
 
5-6
(pg. 
417
-
428
)
18
Marschalek
 
R
Mechanisms of leukemogenesis by MLL fusion proteins.
Br J Haematol
2011
, vol. 
152
 
2
(pg. 
141
-
154
)
19
Meyer
 
C
Hofmann
 
J
Burmeister
 
T
, et al. 
The MLL recombinome of acute leukemias in 2013.
Leukemia
2013
, vol. 
27
 
11
(pg. 
2165
-
2176
)
20
Bueno
 
C
Montes
 
R
Catalina
 
P
Rodríguez
 
R
Menendez
 
P
Insights into the cellular origin and etiology of the infant pro-B acute lymphoblastic leukemia with MLL-AF4 rearrangement.
Leukemia
2011
, vol. 
25
 
3
(pg. 
400
-
410
)
21
Chillón
 
MC
Gómez-Casares
 
MT
López-Jorge
 
CE
, et al. 
Prognostic significance of FLT3 mutational status and expression levels in MLL-AF4+ and MLL-germline acute lymphoblastic leukemia.
Leukemia
2012
, vol. 
26
 
11
(pg. 
2360
-
2366
)
22
Pui
 
CH
Acute lymphoblastic leukemia in children.
Curr Opin Oncol
2000
, vol. 
12
 
1
(pg. 
3
-
12
)
23
Mahajan
 
P
Leavey
 
PJ
Galindo
 
RL
PAX genes in childhood oncogenesis: developmental biology gone awry?
Oncogene
2015
, vol. 
34
 
21
(pg. 
2681
-
2689
)
24
Greaves
 
M
When one mutation is all it takes.
Cancer Cell
2015
, vol. 
27
 
4
(pg. 
433
-
434
)
25
Andersson
 
AK
Ma
 
J
Wang
 
J
, et al. 
St. Jude Children’s Research Hospital–Washington University Pediatric Cancer Genome Project
The landscape of somatic mutations in infant MLL-rearranged acute lymphoblastic leukemias.
Nat Genet
2015
, vol. 
47
 
4
(pg. 
330
-
337
)
26
Kalina
 
T
Flores-Montero
 
J
van der Velden
 
VH
, et al. 
EuroFlow Consortium (EU-FP6, LSHB-CT-2006-018708)
EuroFlow standardization of flow cytometer instrument settings and immunophenotyping protocols.
Leukemia
2012
, vol. 
26
 
9
(pg. 
1986
-
2010
)
27
Bueno
 
C
Montes
 
R
Martín
 
L
, et al. 
NG2 antigen is expressed in CD34+ HPCs and plasmacytoid dendritic cell precursors: is NG2 expression in leukemia dependent on the target cell where leukemogenesis is triggered?
Leukemia
2008
, vol. 
22
 
8
(pg. 
1475
-
1478
)
28
Menendez
 
P
Bueno
 
C
Expression of NG2 antigen in MLL-rearranged acute leukemias: how complex does it get?
Leuk Res
2011
, vol. 
35
 
8
(pg. 
989
-
990
)
29
Bardini
 
M
Woll
 
PS
Corral
 
L
, et al. 
Clonal variegation and dynamic competition of leukemia-initiating cells in infant acute lymphoblastic leukemia with MLL rearrangement.
Leukemia
2015
, vol. 
29
 
1
(pg. 
38
-
50
)
30
Emerenciano
 
M
Renaud
 
G
Sant’Ana
 
M
Barbieri
 
C
Passetti
 
F
Pombo-de-Oliveira
 
MS
Brazilian Collaborative Study Group of Infant Acute Leukemia
Challenges in the use of NG2 antigen as a marker to predict MLL rearrangements in multi-center studies.
Leuk Res
2011
, vol. 
35
 
8
(pg. 
1001
-
1007
)
31
van der Linden
 
MH
Valsecchi
 
MG
De Lorenzo
 
P
, et al. 
Outcome of congenital acute lymphoblastic leukemia treated on the Interfant-99 protocol.
Blood
2009
, vol. 
114
 
18
(pg. 
3764
-
3768
)
32
Mann
 
G
Attarbaschi
 
A
Schrappe
 
M
, et al. 
Interfant-99 Study Group
Improved outcome with hematopoietic stem cell transplantation in a poor prognostic subgroup of infants with mixed-lineage-leukemia (MLL)-rearranged acute lymphoblastic leukemia: results from the Interfant-99 Study.
Blood
2010
, vol. 
116
 
15
(pg. 
2644
-
2650
)
33
Hilden
 
JM
Dinndorf
 
PA
Meerbaum
 
SO
, et al. 
Children’s Oncology Group
Analysis of prognostic factors of acute lymphoblastic leukemia in infants: report on CCG 1953 from the Children’s Oncology Group.
Blood
2006
, vol. 
108
 
2
(pg. 
441
-
451
)
34
Gaynon
 
PS
Angiolillo
 
AL
Carroll
 
WL
, et al. 
Children’s Oncology Group
Long-term results of the children’s cancer group studies for childhood acute lymphoblastic leukemia 1983-2002: a Children’s Oncology Group Report.
Leukemia
2010
, vol. 
24
 
2
(pg. 
285
-
297
)
35
Salzer
 
WL
Jones
 
TL
Devidas
 
M
, et al. 
Modifications to induction therapy decrease risk of early death in infants with acute lymphoblastic leukemia treated on Children’s Oncology Group P9407.
Pediatr Blood Cancer
2012
, vol. 
59
 
5
(pg. 
834
-
839
)
36
Pieters
 
R
den Boer
 
ML
Durian
 
M
, et al. 
Relation between age, immunophenotype and in vitro drug resistance in 395 children with acute lymphoblastic leukemia--implications for treatment of infants.
Leukemia
1998
, vol. 
12
 
9
(pg. 
1344
-
1348
)
37
Pieters
 
R
Schrappe
 
M
De Lorenzo
 
P
, et al. 
A treatment protocol for infants younger than 1 year with acute lymphoblastic leukaemia (Interfant-99): an observational study and a multicentre randomised trial.
Lancet
2007
, vol. 
370
 
9583
(pg. 
240
-
250
)
38
Van der Velden
 
VH
Corral
 
L
Valsecchi
 
MG
, et al. 
Interfant-99 Study Group
Prognostic significance of minimal residual disease in infants with acute lymphoblastic leukemia treated within the Interfant-99 protocol.
Leukemia
2009
, vol. 
23
 
6
(pg. 
1073
-
1079
)
39
Driessen
 
EM
van Roon
 
EH
Spijkers-Hagelstein
 
JA
, et al. 
Frequencies and prognostic impact of RAS mutations in MLL-rearranged acute lymphoblastic leukemia in infants.
Haematologica
2013
, vol. 
98
 
6
(pg. 
937
-
944
)
40
Stam
 
RW
Schneider
 
P
de Lorenzo
 
P
Valsecchi
 
MG
den Boer
 
ML
Pieters
 
R
Prognostic significance of high-level FLT3 expression in MLL-rearranged infant acute lymphoblastic leukemia.
Blood
2007
, vol. 
110
 
7
(pg. 
2774
-
2775
)
41
Stam
 
RW
Schneider
 
P
Hagelstein
 
JA
, et al. 
Gene expression profiling-based dissection of MLL translocated and MLL germline acute lymphoblastic leukemia in infants.
Blood
2010
, vol. 
115
 
14
(pg. 
2835
-
2844
)
42
Hsieh
 
JJ
Cheng
 
EH
Korsmeyer
 
SJ
Taspase1: a threonine aspartase required for cleavage of MLL and proper HOX gene expression.
Cell
2003
, vol. 
115
 
3
(pg. 
293
-
303
)
43
Adler
 
HT
Nallaseth
 
FS
Walter
 
G
Tkachuk
 
DC
HRX leukemic fusion proteins form a heterocomplex with the leukemia-associated protein SET and protein phosphatase 2A.
J Biol Chem
1997
, vol. 
272
 
45
(pg. 
28407
-
28414
)
44
Adler
 
HT
Chinery
 
R
Wu
 
DY
, et al. 
Leukemic HRX fusion proteins inhibit GADD34-induced apoptosis and associate with the GADD34 and hSNF5/INI1 proteins.
Mol Cell Biol
1999
, vol. 
19
 
10
(pg. 
7050
-
7060
)
45
Yokoyama
 
A
Kitabayashi
 
I
Ayton
 
PM
Cleary
 
ML
Ohki
 
M
Leukemia proto-oncoprotein MLL is proteolytically processed into 2 fragments with opposite transcriptional properties.
Blood
2002
, vol. 
100
 
10
(pg. 
3710
-
3718
)
46
Xia
 
ZB
Anderson
 
M
Diaz
 
MO
Zeleznik-Le
 
NJ
MLL repression domain interacts with histone deacetylases, the polycomb group proteins HPC2 and BMI-1, and the corepressor C-terminal-binding protein.
Proc Natl Acad Sci USA
2003
, vol. 
100
 
14
(pg. 
8342
-
8347
)
47
Yokoyama
 
A
Wang
 
Z
Wysocka
 
J
, et al. 
Leukemia proto-oncoprotein MLL forms a SET1-like histone methyltransferase complex with menin to regulate Hox gene expression.
Mol Cell Biol
2004
, vol. 
24
 
13
(pg. 
5639
-
5649
)
48
Dou
 
Y
Milne
 
TA
Tackett
 
AJ
, et al. 
Physical association and coordinate function of the H3 K4 methyltransferase MLL1 and the H4 K16 acetyltransferase MOF.
Cell
2005
, vol. 
121
 
6
(pg. 
873
-
885
)
49
Dou
 
Y
Milne
 
TA
Ruthenburg
 
AJ
, et al. 
Regulation of MLL1 H3K4 methyltransferase activity by its core components.
Nat Struct Mol Biol
2006
, vol. 
13
 
8
(pg. 
713
-
719
)
50
Wysocka
 
J
Swigut
 
T
Milne
 
TA
, et al. 
WDR5 associates with histone H3 methylated at K4 and is essential for H3 K4 methylation and vertebrate development.
Cell
2005
, vol. 
121
 
6
(pg. 
859
-
872
)
51
Muntean
 
AG
Tan
 
J
Sitwala
 
K
, et al. 
The PAF complex synergizes with MLL fusion proteins at HOX loci to promote leukemogenesis.
Cancer Cell
2010
, vol. 
17
 
6
(pg. 
609
-
621
)
52
Milne
 
TA
Kim
 
J
Wang
 
GG
, et al. 
Multiple interactions recruit MLL1 and MLL1 fusion proteins to the HOXA9 locus in leukemogenesis.
Mol Cell
2010
, vol. 
38
 
6
(pg. 
853
-
863
)
53
de Boer
 
J
Walf-Vorderwülbecke
 
V
Williams
 
O
In focus: MLL-rearranged leukemia.
Leukemia
2013
, vol. 
27
 
6
(pg. 
1224
-
1228
)
54
Hom
 
RA
Chang
 
PY
Roy
 
S
, et al. 
Molecular mechanism of MLL PHD3 and RNA recognition by the Cyp33 RRM domain.
J Mol Biol
2010
, vol. 
400
 
2
(pg. 
145
-
154
)
55
Wang
 
Z
Song
 
J
Milne
 
TA
, et al. 
Pro isomerization in MLL1 PHD3-bromo cassette connects H3K4me readout to CyP33 and HDAC-mediated repression.
Cell
2010
, vol. 
141
 
7
(pg. 
1183
-
1194
)
56
Schraets
 
D
Lehmann
 
T
Dingermann
 
T
Marschalek
 
R
MLL-mediated transcriptional gene regulation investigated by gene expression profiling.
Oncogene
2003
, vol. 
22
 
23
(pg. 
3655
-
3668
)
57
Rössler
 
T
Marschalek
 
R
An alternative splice process renders the MLL protein either into a transcriptional activator or repressor.
Pharmazie
2013
, vol. 
68
 
7
(pg. 
601
-
607
)
58
Chen
 
J
Santillan
 
DA
Koonce
 
M
, et al. 
Loss of MLL PHD finger 3 is necessary for MLL-ENL-induced hematopoietic stem cell immortalization.
Cancer Res
2008
, vol. 
68
 
15
(pg. 
6199
-
6207
)
59
Muntean
 
AG
Giannola
 
D
Udager
 
AM
Hess
 
JL
The PHD fingers of MLL block MLL fusion protein-mediated transformation.
Blood
2008
, vol. 
112
 
12
(pg. 
4690
-
4693
)
60
Alexander
 
FE
Patheal
 
SL
Biondi
 
A
, et al. 
Transplacental chemical exposure and risk of infant leukemia with MLL gene fusion.
Cancer Res
2001
, vol. 
61
 
6
(pg. 
2542
-
2546
)
61
Ross
 
JA
Potter
 
JD
Reaman
 
GH
Pendergrass
 
TW
Robison
 
LL
Maternal exposure to potential inhibitors of DNA topoisomerase II and infant leukemia (United States): a report from the Children’s Cancer Group.
Cancer Causes Control
1996
, vol. 
7
 
6
(pg. 
581
-
590
)
62
Spector
 
LG
Xie
 
Y
Robison
 
LL
, et al. 
Maternal diet and infant leukemia: the DNA topoisomerase II inhibitor hypothesis: a report from the children’s oncology group.
Cancer Epidemiol Biomarkers Prev
2005
, vol. 
14
 
3
(pg. 
651
-
655
)
63
Strick
 
R
Strissel
 
PL
Borgers
 
S
Smith
 
SL
Rowley
 
JD
Dietary bioflavonoids induce cleavage in the MLL gene and may contribute to infant leukemia.
Proc Natl Acad Sci USA
2000
, vol. 
97
 
9
(pg. 
4790
-
4795
)
64
Blanco
 
JG
Edick
 
MJ
Relling
 
MV
Etoposide induces chimeric Mll gene fusions.
FASEB J
2004
, vol. 
18
 
1
(pg. 
173
-
175
)
65
Bueno
 
C
Catalina
 
P
Melen
 
GJ
, et al. 
Etoposide induces MLL rearrangements and other chromosomal abnormalities in human embryonic stem cells.
Carcinogenesis
2009
, vol. 
30
 
9
(pg. 
1628
-
1637
)
66
Moneypenny
 
CG
Shao
 
J
Song
 
Y
Gallagher
 
EP
MLL rearrangements are induced by low doses of etoposide in human fetal hematopoietic stem cells.
Carcinogenesis
2006
, vol. 
27
 
4
(pg. 
874
-
881
)
67
Greaves
 
MF
Wiemels
 
J
Origins of chromosome translocations in childhood leukaemia.
Nat Rev Cancer
2003
, vol. 
3
 
9
(pg. 
639
-
649
)
68
Ford
 
AM
Ridge
 
SA
Cabrera
 
ME
, et al. 
In utero rearrangements in the trithorax-related oncogene in infant leukaemias.
Nature
1993
, vol. 
363
 
6427
(pg. 
358
-
360
)
69
Menendez
 
P
Catalina
 
P
Rodríguez
 
R
, et al. 
Bone marrow mesenchymal stem cells from infants with MLL-AF4+ acute leukemia harbor and express the MLL-AF4 fusion gene.
J Exp Med
2009
, vol. 
206
 
13
(pg. 
3131
-
3141
)
70
Mori
 
H
Colman
 
SM
Xiao
 
Z
, et al. 
Chromosome translocations and covert leukemic clones are generated during normal fetal development.
Proc Natl Acad Sci USA
2002
, vol. 
99
 
12
(pg. 
8242
-
8247
)
71
Gale
 
KB
Ford
 
AM
Repp
 
R
, et al. 
Backtracking leukemia to birth: identification of clonotypic gene fusion sequences in neonatal blood spots.
Proc Natl Acad Sci USA
1997
, vol. 
94
 
25
(pg. 
13950
-
13954
)
72
Greaves
 
M
Infection, immune responses and the aetiology of childhood leukaemia.
Nat Rev Cancer
2006
, vol. 
6
 
3
(pg. 
193
-
203
)
73
Tamai
 
H
Miyake
 
K
Takatori
 
M
, et al. 
Activated K-Ras protein accelerates human MLL/AF4-induced leukemo-lymphomogenicity in a transgenic mouse model.
Leukemia
2011
, vol. 
25
 
5
(pg. 
888
-
891
)
74
Prelle
 
C
Bursen
 
A
Dingermann
 
T
Marschalek
 
R
Secondary mutations in t(4;11) leukemia patients.
Leukemia
2013
, vol. 
27
 
6
(pg. 
1425
-
1427
)
75
Armstrong
 
SA
Kung
 
AL
Mabon
 
ME
, et al. 
Inhibition of FLT3 in MLL. Validation of a therapeutic target identified by gene expression based classification.
Cancer Cell
2003
, vol. 
3
 
2
(pg. 
173
-
183
)
76
Armstrong
 
SA
Mabon
 
ME
Silverman
 
LB
, et al. 
FLT3 mutations in childhood acute lymphoblastic leukemia.
Blood
2004
, vol. 
103
 
9
(pg. 
3544
-
3546
)
77
Taketani
 
T
Taki
 
T
Sugita
 
K
, et al. 
FLT3 mutations in the activation loop of tyrosine kinase domain are frequently found in infant ALL with MLL rearrangements and pediatric ALL with hyperdiploidy.
Blood
2004
, vol. 
103
 
3
(pg. 
1085
-
1088
)
78
Emerenciano
 
M
Menezes
 
J
Vasquez
 
ML
Zalcberg
 
I
Thuler
 
LC
Pombo-de-Oliveira
 
MS
Brazilian Collaborative Study Group of Infant Acute Leukemia
Clinical relevance of FLT3 gene abnormalities in Brazilian patients with infant leukemia.
Leuk Lymphoma
2008
, vol. 
49
 
12
(pg. 
2291
-
2297
)
79
Stam
 
RW
den Boer
 
ML
Schneider
 
P
Meier
 
M
Beverloo
 
HB
Pieters
 
R
D-HPLC analysis of the entire FLT3 gene in MLL rearranged and hyperdiploid acute lymphoblastic leukemia.
Haematologica
2007
, vol. 
92
 
11
(pg. 
1565
-
1568
)
80
Bardini
 
M
Spinelli
 
R
Bungaro
 
S
, et al. 
DNA copy-number abnormalities do not occur in infant ALL with t(4;11)/MLL-AF4.
Leukemia
2010
, vol. 
24
 
1
(pg. 
169
-
176
)
81
Bardini
 
M
Galbiati
 
M
Lettieri
 
A
, et al. 
Implementation of array based whole-genome high-resolution technologies confirms the absence of secondary copy-number alterations in MLL-AF4-positive infant ALL patients.
Leukemia
2011
, vol. 
25
 
1
(pg. 
175
-
178
)
82
Guenther
 
MG
Lawton
 
LN
Rozovskaia
 
T
, et al. 
Aberrant chromatin at genes encoding stem cell regulators in human mixed-lineage leukemia.
Genes Dev
2008
, vol. 
22
 
24
(pg. 
3403
-
3408
)
83
Bursen
 
A
Schwabe
 
K
Rüster
 
B
, et al. 
The AF4.MLL fusion protein is capable of inducing ALL in mice without requirement of MLL.AF4.
Blood
2010
, vol. 
115
 
17
(pg. 
3570
-
3579
)
84
Wilkinson
 
AC
Ballabio
 
E
Geng
 
H
, et al. 
RUNX1 is a key target in t(4;11) leukemias that contributes to gene activation through an AF4-MLL complex interaction.
Cell Reports
2013
, vol. 
3
 
1
(pg. 
116
-
127
)
85
Kamper-Jørgensen
 
M
Woodward
 
A
Wohlfahrt
 
J
, et al. 
Childcare in the first 2 years of life reduces the risk of childhood acute lymphoblastic leukemia.
Leukemia
2008
, vol. 
22
 
1
(pg. 
189
-
193
)
86
Gilham
 
C
Peto
 
J
Simpson
 
J
, et al. 
UKCCS Investigators
Day care in infancy and risk of childhood acute lymphoblastic leukaemia: findings from UK case-control study.
BMJ
2005
, vol. 
330
 
7503
pg. 
1294
 
87
Swaminathan
 
S
Klemm
 
L
Park
 
E
, et al. 
Mechanisms of clonal evolution in childhood acute lymphoblastic leukemia.
Nat Immunol
2015
, vol. 
16
 
7
(pg. 
766
-
774
)
88
Ford
 
AM
Palmi
 
C
Bueno
 
C
, et al. 
The TEL-AML1 leukemia fusion gene dysregulates the TGF-beta pathway in early B lineage progenitor cells.
J Clin Invest
2009
, vol. 
119
 
4
(pg. 
826
-
836
)
89
Mullighan
 
CG
Goorha
 
S
Radtke
 
I
, et al. 
Genome-wide analysis of genetic alterations in acute lymphoblastic leukaemia.
Nature
2007
, vol. 
446
 
7137
(pg. 
758
-
764
)
90
Dobbins
 
SE
Sherborne
 
AL
Ma
 
YP
, et al. 
The silent mutational landscape of infant MLL-AF4 pro-B acute lymphoblastic leukemia.
Genes Chromosomes Cancer
2013
, vol. 
52
 
10
(pg. 
954
-
960
)
91
Raynaud
 
S
Cave
 
H
Baens
 
M
, et al. 
The 12;21 translocation involving TEL and deletion of the other TEL allele: two frequently associated alterations found in childhood acute lymphoblastic leukemia.
Blood
1996
, vol. 
87
 
7
(pg. 
2891
-
2899
)
92
Papaemmanuil
 
E
Rapado
 
I
Li
 
Y
, et al. 
RAG-mediated recombination is the predominant driver of oncogenic rearrangement in ETV6-RUNX1 acute lymphoblastic leukemia.
Nat Genet
2014
, vol. 
46
 
2
(pg. 
116
-
125
)
93
Chuk
 
MK
McIntyre
 
E
Small
 
D
Brown
 
P
Discordance of MLL-rearranged (MLL-R) infant acute lymphoblastic leukemia in monozygotic twins with spontaneous clearance of preleukemic clone in unaffected twin.
Blood
2009
, vol. 
113
 
26
(pg. 
6691
-
6694
)
94
Maia
 
AT
Koechling
 
J
Corbett
 
R
Metzler
 
M
Wiemels
 
JL
Greaves
 
M
Protracted postnatal natural histories in childhood leukemia.
Genes Chromosomes Cancer
2004
, vol. 
39
 
4
(pg. 
335
-
340
)
95
Bonnal
 
S
Vigevani
 
L
Valcárcel
 
J
The spliceosome as a target of novel antitumour drugs.
Nat Rev Drug Discov
2012
, vol. 
11
 
11
(pg. 
847
-
859
)
96
Greaves
 
M
Cancer stem cells: back to Darwin?
Semin Cancer Biol
2010
, vol. 
20
 
2
(pg. 
65
-
70
)
97
Arai
 
E
Sakamoto
 
H
Ichikawa
 
H
, et al. 
Multilayer-omics analysis of renal cell carcinoma, including the whole exome, methylome and transcriptome.
Int J Cancer
2014
, vol. 
135
 
6
(pg. 
1330
-
1342
)
98
Krivtsov
 
AV
Feng
 
Z
Lemieux
 
ME
, et al. 
H3K79 methylation profiles define murine and human MLL-AF4 leukemias.
Cancer Cell
2008
, vol. 
14
 
5
(pg. 
355
-
368
)
99
Chen
 
CW
Armstrong
 
SA
Targeting DOT1L and HOX gene expression in MLL-rearranged leukemia and beyond.
Exp Hematol
2015
, vol. 
43
 
8
(pg. 
673
-
684
)
100
Fernández
 
AF
Bayón
 
GF
Urdinguio
 
RG
, et al. 
H3K4me1 marks DNA regions hypomethylated during aging in human stem and differentiated cells.
Genome Res
2015
, vol. 
25
 
1
(pg. 
27
-
40
)
101
Calvanese
 
V
Fernández
 
AF
Urdinguio
 
RG
, et al. 
A promoter DNA demethylation landscape of human hematopoietic differentiation.
Nucleic Acids Res
2012
, vol. 
40
 
1
(pg. 
116
-
131
)
102
Krivtsov
 
AV
Armstrong
 
SA
MLL translocations, histone modifications and leukaemia stem-cell development.
Nat Rev Cancer
2007
, vol. 
7
 
11
(pg. 
823
-
833
)
103
Neff
 
T
Armstrong
 
SA
Recent progress toward epigenetic therapies: the example of mixed lineage leukemia.
Blood
2013
, vol. 
121
 
24
(pg. 
4847
-
4853
)
104
Ikawa
 
Y
Sugimoto
 
N
Koizumi
 
S
Yachie
 
A
Saikawa
 
Y
Dense methylation of types 1 and 2 regulatory regions of the CD10 gene promoter in infant acute lymphoblastic leukemia with MLL/AF4 fusion gene.
J Pediatr Hematol Oncol
2010
, vol. 
32
 
1
(pg. 
4
-
10
)
105
Stumpel
 
DJ
Schneider
 
P
van Roon
 
EH
Pieters
 
R
Stam
 
RW
Absence of global hypomethylation in promoter hypermethylated Mixed Lineage Leukaemia-rearranged infant acute lymphoblastic leukaemia.
Eur J Cancer
2013
, vol. 
49
 
1
(pg. 
175
-
184
)
106
Altucci
 
L
Minucci
 
S
Epigenetic therapies in haematological malignancies: searching for true targets.
Eur J Cancer
2009
, vol. 
45
 
7
(pg. 
1137
-
1145
)
107
Stumpel
 
DJ
Schneider
 
P
van Roon
 
EH
, et al. 
Specific promoter methylation identifies different subgroups of MLL-rearranged infant acute lymphoblastic leukemia, influences clinical outcome, and provides therapeutic options.
Blood
2009
, vol. 
114
 
27
(pg. 
5490
-
5498
)
108
Stumpel
 
DJ
Schotte
 
D
Lange-Turenhout
 
EA
, et al. 
Hypermethylation of specific microRNA genes in MLL-rearranged infant acute lymphoblastic leukemia: major matters at a micro scale.
Leukemia
2011
, vol. 
25
 
3
(pg. 
429
-
439
)
109
Chen
 
W
Li
 
Q
Hudson
 
WA
Kumar
 
A
Kirchhof
 
N
Kersey
 
JH
A murine Mll-AF4 knock-in model results in lymphoid and myeloid deregulation and hematologic malignancy.
Blood
2006
, vol. 
108
 
2
(pg. 
669
-
677
)
110
Metzler
 
M
Forster
 
A
Pannell
 
R
, et al. 
A conditional model of MLL-AF4 B-cell tumourigenesis using invertor technology.
Oncogene
2006
, vol. 
25
 
22
(pg. 
3093
-
3103
)
111
Tamai
 
H
Inokuchi
 
K
Establishment of MLL/AF4 transgenic mice with the phenotype of lymphoblastic leukemia or lymphoma.
J Nippon Med Sch
2013
, vol. 
80
 
5
(pg. 
326
-
327
)
112
Kowarz
 
E
Burmeister
 
T
Lo Nigro
 
L
, et al. 
Complex MLL rearrangements in t(4;11) leukemia patients with absent AF4.MLL fusion allele.
Leukemia
2007
, vol. 
21
 
6
(pg. 
1232
-
1238
)
113
Kumar
 
AR
Yao
 
Q
Li
 
Q
Sam
 
TA
Kersey
 
JH
t(4;11) leukemias display addiction to MLL-AF4 but not to AF4-MLL.
Leuk Res
2011
, vol. 
35
 
3
(pg. 
305
-
309
)
114
Sanders
 
DS
Muntean
 
AG
Hess
 
JL
Significance of AF4-MLL reciprocal fusion in t(4;11) leukemias?
Leuk Res
2011
, vol. 
35
 
3
(pg. 
299
-
300
)
115
Marschalek
 
R
It takes two-to-leukemia: about addictions and requirements.
Leuk Res
2011
, vol. 
35
 
3
(pg. 
424
-
425
)
116
Montes
 
R
Ayllón
 
V
Gutierrez-Aranda
 
I
, et al. 
Enforced expression of MLL-AF4 fusion in cord blood CD34+ cells enhances the hematopoietic repopulating cell function and clonogenic potential but is not sufficient to initiate leukemia.
Blood
2011
, vol. 
117
 
18
(pg. 
4746
-
4758
)
117
Bueno
 
C
Montes
 
R
Melen
 
GJ
, et al. 
A human ESC model for MLL-AF4 leukemic fusion gene reveals an impaired early hematopoietic-endothelial specification.
Cell Res
2012
, vol. 
22
 
6
(pg. 
986
-
1002
)
118
Bueno
 
C
Ayllon
 
V
Montes
 
R
, et al. 
 
FLT3 activation cooperates with MLL-AF4 fusion protein to abrogate the hematopoietic specification of human ESCs. Blood. 2013;121(19):3867-3878
119
Montes
 
R
Ayllón
 
V
Prieto
 
C
, et al. 
Ligand-independent FLT3 activation does not cooperate with MLL-AF4 to immortalize/transform cord blood CD34+ cells.
Leukemia
2014
, vol. 
28
 
3
(pg. 
666
-
674
)
120
Hotfilder
 
M
Röttgers
 
S
Rosemann
 
A
, et al. 
Leukemic stem cells in childhood high-risk ALL/t(9;22) and t(4;11) are present in primitive lymphoid-restricted CD34+CD19- cells.
Cancer Res
2005
, vol. 
65
 
4
(pg. 
1442
-
1449
)
121
Shalapour
 
S
Eckert
 
C
Seeger
 
K
, et al. 
Leukemia-associated genetic aberrations in mesenchymal stem cells of children with acute lymphoblastic leukemia.
J Mol Med (Berl)
2010
, vol. 
88
 
3
(pg. 
249
-
265
)
122
Montecino-Rodriguez
 
E
Li
 
K
Fice
 
M
Dorshkind
 
K
Murine B-1 B cell progenitors initiate B-acute lymphoblastic leukemia with features of high-risk disease.
J Immunol
2014
, vol. 
192
 
11
(pg. 
5171
-
5178
)
123
Aoki
 
Y
Watanabe
 
T
Saito
 
Y
, et al. 
Identification of CD34+ and CD34- leukemia-initiating cells in MLL-rearranged human acute lymphoblastic leukemia.
Blood
2015
, vol. 
125
 
6
(pg. 
967
-
980
)
124
le Viseur
 
C
Hotfilder
 
M
Bomken
 
S
, et al. 
In childhood acute lymphoblastic leukemia, blasts at different stages of immunophenotypic maturation ha7ve stem cell properties.
Cancer Cell
2008
, vol. 
14
 
1
(pg. 
47
-
58
)
Sign in via your Institution