T-cell acute lymphoblastic leukemia (T-ALL) is an aggressive malignancy caused by the accumulation of genomic lesions that affect the development of T cells. For many years, it has been established that deregulated expression of transcription factors, impairment of the CDKN2A/2B cell-cycle regulators, and hyperactive NOTCH1 signaling play prominent roles in the pathogenesis of this leukemia. In the past decade, systematic screening of T-ALL genomes by high-resolution copy-number arrays and next-generation sequencing technologies has revealed that T-cell progenitors accumulate additional mutations affecting JAK/STAT signaling, protein translation, and epigenetic control, providing novel attractive targets for therapy. In this review, we provide an update on our knowledge of T-ALL pathogenesis, the opportunities for the introduction of targeted therapy, and the challenges that are still ahead.

The characterization of chromosomal abnormalities, such as 9p deletions resulting in inactivation of CDKN2A (p16) and CDKN2B (p15) and translocations affecting the T-cell receptor (TCR) genes, has been fundamental in providing initial insights into the genetic defects present in T-cell acute lymphoblastic leukemia (T-ALL). The incorporation of gene expression profiling has provided an additional view on the subgroups present in T-ALL, each characterized by the presence of specific chromosomal aberrations leading to ectopic expression of 1 particular transcription factor such as TAL1, TLX1, TLX3, or others. Sequencing approaches focusing on candidate oncogenes or, more recently, genome-wide sequencing (exome sequencing, whole-genome sequencing, or transcriptome sequencing), have identified >100 genes that can be mutated in T-ALL. Only 2 of these genes, NOTCH1 and CDKN2A/2B, are mutated in >50% of T-ALL cases, and a large variety of genes are mutated at lower frequency (Table 1). Based on all available genomic data, we can conclude that each T-ALL case contains probably >10 biologically relevant genomic lesions, each contributing to the transformation of normal T cells into an aggressive leukemia cell with impaired differentiation, improved survival and proliferation characteristics, and altered metabolism, cell cycle, and homing properties. These changes help the T-ALL cells proliferate and survive by changing their own behavior as well as interact effectively with normal cells, which further supports their stem cell characteristics and the survival of these high numbers of abnormal T cells.

Table 1.

Mutation frequencies in adult vs pediatric T-ALL

GeneType of genetic aberrationFrequency
PediatricAdult
NOTCH1 signaling pathway    
FBXW7 Inactivating mutations 14 14 
NOTCH1 Chromosomal rearrangements/activating mutations 50 57 
Cell cycle    
CDKN2A 9p21 deletion 61 55 
CDKN2B 9p21 deletion 58 46 
RB1 Deletions 12* 
Transcription factors    
BCL11B Inactivating mutations/deletions 10 
ETV6 Inactivating mutations/deletions 14 
GATA3 Inactivating mutations/deletions 
HOXA (CALM-AF10, MLL-ENL and SET-NUP214) Chromosomal rearrangements/inversions/expression 
LEF1 Inactivating mutations/deletions 10 
LMO2 Chromosomal rearrangements/deletions/expression 13 21 
MYB Chromosomal rearrangements/duplications 17 
NKX2.1/NKX2.2 Chromosomal rearrangements/expression 8* 
RUNX1 Inactivating mutations/deletions 10 
TAL1 Chromosomal rearrangements/5′ super-enhancer mutations/deletions/expression 30 34 
TLX1 Chromosomal rearrangements/deletions/expression 20 
TLX3 Chromosomal rearrangements/expression 19 
WT1 Inactivating mutation/deletion 19 11 
Signaling    
AKT Activating mutations 
DNM2 Inactivating mutations 13 13 
FLT3 Activating mutations 
JAK1 Activating mutations 
JAK3 Activating mutations 12 
IL7R Activating mutations 10 12 
NF1 Deletions 
KRAS Activating mutations 
NRAS Activating mutations 14 
NUP214-ABL1/ ABL1 gain Chromosomal rearrangement/duplication 8* 
PI3KCA Activating mutations 
PTEN Inactivating mutations/deletion 19 11 
PTPN2 Inactivating mutations/deletion 
STAT5B Activating mutations 
Epigenetic factors    
DNMT3A Inactivating mutations 14 
EED Inactivating mutations/deletions 
EZH2 Inactivating mutations/deletions 12 12 
KDM6A/UTX Inactivating mutations/deletions 
PHF6 Inactivating mutations/deletions 19 30 
SUZ12 Inactivating mutations/deletions 11 
Translation and RNA stability    
CNOT3 Missense mutations 
mTOR Activating mutations 5* 
RPL5 Inactivating mutations 
RPL10 Missense mutations 
RPL22 Inactivating mutations/deletion 
GeneType of genetic aberrationFrequency
PediatricAdult
NOTCH1 signaling pathway    
FBXW7 Inactivating mutations 14 14 
NOTCH1 Chromosomal rearrangements/activating mutations 50 57 
Cell cycle    
CDKN2A 9p21 deletion 61 55 
CDKN2B 9p21 deletion 58 46 
RB1 Deletions 12* 
Transcription factors    
BCL11B Inactivating mutations/deletions 10 
ETV6 Inactivating mutations/deletions 14 
GATA3 Inactivating mutations/deletions 
HOXA (CALM-AF10, MLL-ENL and SET-NUP214) Chromosomal rearrangements/inversions/expression 
LEF1 Inactivating mutations/deletions 10 
LMO2 Chromosomal rearrangements/deletions/expression 13 21 
MYB Chromosomal rearrangements/duplications 17 
NKX2.1/NKX2.2 Chromosomal rearrangements/expression 8* 
RUNX1 Inactivating mutations/deletions 10 
TAL1 Chromosomal rearrangements/5′ super-enhancer mutations/deletions/expression 30 34 
TLX1 Chromosomal rearrangements/deletions/expression 20 
TLX3 Chromosomal rearrangements/expression 19 
WT1 Inactivating mutation/deletion 19 11 
Signaling    
AKT Activating mutations 
DNM2 Inactivating mutations 13 13 
FLT3 Activating mutations 
JAK1 Activating mutations 
JAK3 Activating mutations 12 
IL7R Activating mutations 10 12 
NF1 Deletions 
KRAS Activating mutations 
NRAS Activating mutations 14 
NUP214-ABL1/ ABL1 gain Chromosomal rearrangement/duplication 8* 
PI3KCA Activating mutations 
PTEN Inactivating mutations/deletion 19 11 
PTPN2 Inactivating mutations/deletion 
STAT5B Activating mutations 
Epigenetic factors    
DNMT3A Inactivating mutations 14 
EED Inactivating mutations/deletions 
EZH2 Inactivating mutations/deletions 12 12 
KDM6A/UTX Inactivating mutations/deletions 
PHF6 Inactivating mutations/deletions 19 30 
SUZ12 Inactivating mutations/deletions 11 
Translation and RNA stability    
CNOT3 Missense mutations 
mTOR Activating mutations 5* 
RPL5 Inactivating mutations 
RPL10 Missense mutations 
RPL22 Inactivating mutations/deletion 

Genes targeted by genetic alterations in >2% of T-ALL cases are shown. Calculation of the different frequencies is based on several previously published independent studies that analyzed pediatric and/or adult T-ALL patient cohorts. For frequency calculations, chromosomal rearrangements, copy-number variations and mutations were considered. For HOXA, TLX1, TLX3, TAL1, NKX2.1, NKX2.2, and LMO2, gene expression was also considered. The HOXA group includes cases that are positive for CALM-AF10, SET-NUP214, or MLL-ELN chromosomal rearrangements.

*

It was not possible to have separate numbers for pediatric and adult cases.

Careful gene expression profiling of T-ALL cases has led to the identification of subgroups of T-ALL, each characterized by a specific transcriptional profile and the ectopic expression of 1 particular transcription factor, often as a consequence of a chromosomal defect. The largest subgroup is defined by ectopic TAL1 expression (in some cases together with LMO1/LMO2), whereas other major subgroups show mutual exclusive expression of TLX1, TLX3, HOXA9/10, LMO2, or NKX2-1 (Table 2).1-3  Furthermore, the early T-cell precursor subgroup of ALL (ETP-ALL) corresponds to immature T-ALLs expressing ETP/stem cell genes and is characterized by aberrant expression of LYL12 ; hematopoietic transcription factors such as RUNX1 and ETV6 are frequently mutated in this genetically heterogeneous subgroup.1,4 

Table 2.

Chromosomal rearrangements resulting in ectopic expression of transcription factors or the generation of fusion genes with transcriptional/epigenetic activity

Chromosomal rearrangement or mutationPartner gene 1 (oncogene)Chromosome locationPartner gene 2Chromosome locationConsequence of the rearrangement/mutation
Major transcription factors      
 del(1)(p32p32) TAL1 1p32 STIL 1p32 Ectopic TAL1 expression driven by STIL promoter 
 t(1;14)(p32;q11) TCRα 14q11 Ectopic TAL1 expression driven by TCR enhancer 
 t(1;14)(p32;q11) TCRδ 14q11 Ectopic TAL1 expression driven by TCR enhancer 
 Small insertion   Ectopic TAL1 expression driven by de novo enhancer 
 t(10;14)(q24;q11) TLX1 10q24 TCRα 14q11 Ectopic TLX1 expression driven by TCR enhancer 
 t(7;10)(q34;q24) TCRβ 7q34 Ectopic TLX1 expression driven by TCR enhancer 
 t(10;14)(q24;q11) TCRδ 14q11 Ectopic TLX1 expression driven by TCR enhancer 
 t(5;14)(q35;q11) TLX3 5q35 TCRδ 14q11 Ectopic TLX3 expression driven by TCR enhancer 
 t(5;14)(q35;q32) BCL11B 14q32 Ectopic TLX3 expression driven by BCL11B 
 inv(7)(p15;q34) HOXA9/HOXA10 7p15 TCRβ 7q34 Ectopic expression of HOXA genes, predominantly HOXA9 and HOXA10, driven by TCR enhancer 
 t(10;11)(p13;q14) PICALM (CALM) 11q14 MLLT10 (AF10) 10p13 PICALM-MLLT10 fusion transcript, resulting in upregulation of HOXA genes 
 del(9)(q34;q34) SET 9q34 NUP214 9q34 SET-NUP214 fusion transcript, resulting in upregulation of HOXA genes 
 inv(14)(q11;q13) NKX2-1 14q13 TCRα 14q11 Ectopic NKX2-1 expression driven by TCR enhancer 
 T(14;20)(q11;p11) NKX2-2 20p11 TCRδ 14q11 Ectopic NKX2-2 expression driven by TCR enhancer 
 t(11;14)(p15;q11) LMO1 11p15 TCRα 14q11 Ectopic LMO1 expression, mostly together with LYL1 or TAL1 expression 
TCRδ 14q11 
 t(11;14)(p13;q11) LMO2 11p13 TCRα 14q11 Ectopic LMO2 expression, mostly together with LYL1 or TAL1 expression 
TCRδ 14q11 
 del(5)(q14;q14) MEF2C 5q14   Small deletions close to MEF2C leading to upregulation of MEF2C expression 
 t(7;19)(q34;p13) LYL1 19p13 TCRβ 7q34 Ectopic expression of LYL1 driven by TCR enhancer 
 t(11;14)(p11;q32) SPI1 11p11 BCL11B 14q32 Ectopic expression of SPI1 driven by BCL11B 
Additional transcription factor aberrations      
 dup(6)(q23;q23) MYB 6q23   Increased MYB expression due to extra copy of MYB 
 t(6;7)(q23;q34) TCRβ 7q34 Ectopic expression of MYB driven by TCR enhancer 
 Various translocations KMT2A (MLL) 11q23 Many different partners  KMT2A fusion genes 
 Mutation/deletion BCL11B 14q32   Inactivation of BCL11B 
 Mutation/deletion ETV6 12p13   Inactivation of ETV6 
 Mutation/deletion RUNX1 21q22   Inactivation of RUNX1 
 Mutation/deletion LEF1 4q25   Inactivation of LEF1 
 Mutation/deletion WT1 11p13   Inactivation of WT1 
Chromosomal rearrangement or mutationPartner gene 1 (oncogene)Chromosome locationPartner gene 2Chromosome locationConsequence of the rearrangement/mutation
Major transcription factors      
 del(1)(p32p32) TAL1 1p32 STIL 1p32 Ectopic TAL1 expression driven by STIL promoter 
 t(1;14)(p32;q11) TCRα 14q11 Ectopic TAL1 expression driven by TCR enhancer 
 t(1;14)(p32;q11) TCRδ 14q11 Ectopic TAL1 expression driven by TCR enhancer 
 Small insertion   Ectopic TAL1 expression driven by de novo enhancer 
 t(10;14)(q24;q11) TLX1 10q24 TCRα 14q11 Ectopic TLX1 expression driven by TCR enhancer 
 t(7;10)(q34;q24) TCRβ 7q34 Ectopic TLX1 expression driven by TCR enhancer 
 t(10;14)(q24;q11) TCRδ 14q11 Ectopic TLX1 expression driven by TCR enhancer 
 t(5;14)(q35;q11) TLX3 5q35 TCRδ 14q11 Ectopic TLX3 expression driven by TCR enhancer 
 t(5;14)(q35;q32) BCL11B 14q32 Ectopic TLX3 expression driven by BCL11B 
 inv(7)(p15;q34) HOXA9/HOXA10 7p15 TCRβ 7q34 Ectopic expression of HOXA genes, predominantly HOXA9 and HOXA10, driven by TCR enhancer 
 t(10;11)(p13;q14) PICALM (CALM) 11q14 MLLT10 (AF10) 10p13 PICALM-MLLT10 fusion transcript, resulting in upregulation of HOXA genes 
 del(9)(q34;q34) SET 9q34 NUP214 9q34 SET-NUP214 fusion transcript, resulting in upregulation of HOXA genes 
 inv(14)(q11;q13) NKX2-1 14q13 TCRα 14q11 Ectopic NKX2-1 expression driven by TCR enhancer 
 T(14;20)(q11;p11) NKX2-2 20p11 TCRδ 14q11 Ectopic NKX2-2 expression driven by TCR enhancer 
 t(11;14)(p15;q11) LMO1 11p15 TCRα 14q11 Ectopic LMO1 expression, mostly together with LYL1 or TAL1 expression 
TCRδ 14q11 
 t(11;14)(p13;q11) LMO2 11p13 TCRα 14q11 Ectopic LMO2 expression, mostly together with LYL1 or TAL1 expression 
TCRδ 14q11 
 del(5)(q14;q14) MEF2C 5q14   Small deletions close to MEF2C leading to upregulation of MEF2C expression 
 t(7;19)(q34;p13) LYL1 19p13 TCRβ 7q34 Ectopic expression of LYL1 driven by TCR enhancer 
 t(11;14)(p11;q32) SPI1 11p11 BCL11B 14q32 Ectopic expression of SPI1 driven by BCL11B 
Additional transcription factor aberrations      
 dup(6)(q23;q23) MYB 6q23   Increased MYB expression due to extra copy of MYB 
 t(6;7)(q23;q34) TCRβ 7q34 Ectopic expression of MYB driven by TCR enhancer 
 Various translocations KMT2A (MLL) 11q23 Many different partners  KMT2A fusion genes 
 Mutation/deletion BCL11B 14q32   Inactivation of BCL11B 
 Mutation/deletion ETV6 12p13   Inactivation of ETV6 
 Mutation/deletion RUNX1 21q22   Inactivation of RUNX1 
 Mutation/deletion LEF1 4q25   Inactivation of LEF1 
 Mutation/deletion WT1 11p13   Inactivation of WT1 

—, not applicable.

Almost all subgroups of T-ALL are characterized by the clear ectopic expression of 1 of these transcription factors, except for some immature T-ALL cases where MEF2C expression, and other sporadic translocations where SPI1, could be important.1  The oncogenic roles of TAL1 and TLX1 have been nicely illustrated by studies in the mouse, in cell lines, and in patient samples. Overexpression of TAL1 or TLX1 in developing thymocytes in the mouse results in the development of T-ALL with long latency.5-8  For TAL1, coexpression of LMO1 and ICN1 dramatically decreased disease latency, in part by increasing the number of leukemia-initiating cells.9  Mouse leukemias that eventually developed in the context of TLX1 expression often harbored mutations in Bcl11b or Notch1, as also observed in human T-ALL.7,8  These data confirm that ectopic expression of transcription factors such as TAL1 or TLX1 alone can be an initiating step in T-ALL development, but that additional mutations are required to fully transform normal T cells to leukemia cells. The ectopic expression of these transcription factors could have a strong effect on the differentiation of the cells, as suggested by the correlation with immunophenotype. Moreover, for TLX1, it was recently shown that TLX1 can bind the TCR enhancer and in this way can suppress TCRα expression and T-cell differentiation.10 

There are now 4 major mechanisms known to cause aberrant expression of transcription factors in T-ALL: (1) chromosomal translocations involving 1 of the TCR genes, (2) chromosomal rearrangements with other regulatory sequences, (3) duplication/amplification of the transcription factor, and (4) mutations or small insertions generating novel regulatory sequences acting as enhancers (Table 2). The latter mechanism was only recently identified by studies that revealed changes in chromatin structure close to the TAL1 gene, indicative for the presence of a new enhancer region. Detailed analysis of this region led to the identification of mutations that created a de novo binding site for MYB, thereby resulting in recruitment of additional transcriptional regulators and activation of TAL1 expression in cis.11,12  These data illustrate how noncoding mutations can have a strong effect on leukemia development and we can expect that more of these mutations will be identified in the future as more information comes available on the noncoding part of the genome.

The NOTCH1 signaling pathway is essential for the commitment of multipotent hematopoietic progenitors to the T-cell lineage and for further development of thymocytes.13,14  Activation of NOTCH1 signaling constitutes the most predominant oncogenic event involved in the pathogenesis of T-ALL with activating mutations in more than half of T-ALL cases (Table 1).15  Loss of function of negative regulators of NOTCH1 is an alternative mechanism leading to aberrant activation of the pathway. Mutations of FBXW7 are present in 10% to 15% of T-ALL cases and lead to increased NOTCH1 protein stability.16,17  The role of the NOTCH1 cascade in the context of T-ALL is discussed in more detail in a companion review article.18 

Interleukin 7 (IL7) signaling is essential for normal T-cell development and is triggered by the interaction of IL7 with the heterodimeric IL7 receptor (IL7R). This interaction induces reciprocal JAK1 and JAK3 phosphorylation and subsequent recruitment and activation of STAT5. Upon phosphorylation, STAT5 dimerizes and translocates to the nucleus where it regulates the transcription of many target genes, including the antiapoptotic B-cell lymphoma 2 (BCL-2) family member proteins.19,20  In addition to the JAK/STAT pathway, the RAS-MAPK and phosphatidylinositol 3-kinase (PI3K) pathways are also activated by IL2, IL7, and stem cell factor (SCF) that act on the developing T cells (Figure 1).

Figure 1.

Deregulation of the JAK-STAT signaling cascade in T-ALL. Representation of the different oncogenic mechanisms that lead to aberrant activation of the IL7 signaling in T-ALL. Interaction of IL7 with the heterodimeric IL7R induces reciprocal JAK1 and JAK3 phosphorylation and subsequent recruitment of STAT5. STAT5 dimerizes and translocates to the nucleus where it induces transcription of the prosurvival factor BCL2. IL7 also activates the RAS-MAPK and PI3K kinase pathways. IL7 signaling can indirectly be enhanced by abnormal NOTCH1 signaling, constitutive expression of ZEB2, or by increased presentation of IL7R on the cell surface of thymocytes due to impaired clathrin-dependent endocytosis caused by DNM2 mutations. Promising therapeutic agents targeting the oncogenic IL7-JAK-STAT cascade are indicated in red. *Proteins that are mutated in T-ALL.

Figure 1.

Deregulation of the JAK-STAT signaling cascade in T-ALL. Representation of the different oncogenic mechanisms that lead to aberrant activation of the IL7 signaling in T-ALL. Interaction of IL7 with the heterodimeric IL7R induces reciprocal JAK1 and JAK3 phosphorylation and subsequent recruitment of STAT5. STAT5 dimerizes and translocates to the nucleus where it induces transcription of the prosurvival factor BCL2. IL7 also activates the RAS-MAPK and PI3K kinase pathways. IL7 signaling can indirectly be enhanced by abnormal NOTCH1 signaling, constitutive expression of ZEB2, or by increased presentation of IL7R on the cell surface of thymocytes due to impaired clathrin-dependent endocytosis caused by DNM2 mutations. Promising therapeutic agents targeting the oncogenic IL7-JAK-STAT cascade are indicated in red. *Proteins that are mutated in T-ALL.

Close modal

Activating mutations in IL7R, JAK1, JAK3, and/or STAT5 are present in 20% to 30% of T-ALL cases (Table 1),4,21,22  with a higher representation within TLX+, HOXA+, and ETP-ALL patient subgroups (Figure 2).4,23  A small percentage of cases also show aberrations in phosphatases like PTPN2 and PTPRC, which, among other substrates, dephosphorylate and inactivate JAK kinases.24,25  Interestingly, the IL7R signaling cascade can also be hyperactivated in patients who do not carry any genetic aberrations in the IL7R, JAK, or STAT5 genes, indicating that still other mechanisms exist to activate this pathway.26,27  Indeed, only recently was it discovered that loss-of-function mutations in DNM2 lead to increased presentation of IL7R on the cell surface of thymocytes due to impaired clathrin-dependent endocytosis.26  In addition, a rare translocation targeting the zinc finger E-box–binding homeobox 2 locus (ZEB2) has been recently described in T-ALL, and Zeb2 overexpression was shown to increase Il7r expression and Stat5 activation, promoting T-ALL cell survival in a mouse model.27  Although most mutations seem to result in activation of JAK1 and JAK3, rare fusion transcripts involving JAK2 have also been described, and wild-type TYK2 may have an important role in activating survival pathways of T-ALL cells.28,29 

Figure 2.

Representation of the cooperation of oncogenic events. The major subclasses of T-ALL are shown based on the expression of the transcription factors TAL1, TLX1, TLX3, HOXA genes, NKX2-1, or LMO2/LYL1. For each subclass, additional genes are shown that are most frequently mutated in that subclass. The PRC2 complex contains EZH2, SUZ12, and EED.

Figure 2.

Representation of the cooperation of oncogenic events. The major subclasses of T-ALL are shown based on the expression of the transcription factors TAL1, TLX1, TLX3, HOXA genes, NKX2-1, or LMO2/LYL1. For each subclass, additional genes are shown that are most frequently mutated in that subclass. The PRC2 complex contains EZH2, SUZ12, and EED.

Close modal

Aberrant activation of the PI3K-AKT pathway results in enhanced cell metabolism, proliferation, and impaired apoptosis.30,31  Hyperactivation of the oncogenic PI3K-AKT pathway in T-ALL is mainly caused by inactivating mutations or deletions of the phosphatase and tensin homolog (PTEN), the main negative regulator of the pathway.22,32-36  In addition, some T-ALL cases show gain-of-function mutations in the regulatory and catalytic subunits of PI3K, respectively, p85 and p110, or in the downstream effectors of the cascade such as AKT and mechanistic target of rapamycin (mTOR) (Table 1).22,35,37  PI3K-AKT signaling activation is also achieved through IL7 stimulation or RAS activation.22,31,38  A recent study on 146 pediatric T-ALL cases described that almost 50% of the patients harbored at least 1 mutation in the JAK-STAT, PI3K-AKT, or RAS-MAPK pathways, underscoring the importance of activation of those cascades for the leukemic cells.22 

In addition to the activation of the JAK kinases or the PI3K pathway, activation of the ABL1 kinase is observed in up to 8% of T-ALL cases, with 6% of patients expressing the NUP214-ABL1 fusion protein.39  NUP214-ABL1 is a constitutively active tyrosine kinase that activates STAT5 and the RAS-MAPK pathway, but is much weaker as compared with BCR-ABL1.39,40  Interestingly, the kinase activity of NUP214-ABL1 is dependent on its location to the nuclear pore39,41  and its oncogenic properties rely on its interaction with MAD2L1, NUP155, and SMC4 and on the activity of the LCK, a member of the SRC family kinase.42  All data collected to date suggest that the NUP214-ABL1 fusion kinase is a weak oncogene that cooperates with additional oncogenic events to drive leukemia development. In that sense, it is of interest to note that the NUP214-ABL1 fusion is always found together with TLX1 or TLX3 expression and is often associated with loss of PTPN2, a negative regulator of NUP214-ABL1.25,39 

RAS-activating mutations were recently described in 44% of diagnosis-relapse sample pairs, indicating an overrepresentation of these defects in high-risk ALL. Interestingly, KRAS mutations rendered lymphoblasts resistant toward methotrexate, while sensitizing them to vincristine.43 

T-ALL is one of the pediatric tumor types with the highest incidence of epigenetic lesions, and 56% of samples from children with T-ALL contain mutations in this gene class.44  Also, in adult T-ALL, mutations in epigenetic factors are highly common (Table 1).45,46  For a complete overview of all epigenetic lesions identified in T-ALL, we refer to recent reviews on this topic.47,48 

The plant homeodomain protein 6 (PHF6) protein was postulated as an epigenetic factor because it contains 2 atypical plant homeodomain (PHD)-like zinc fingers. Canonical PHD domains typically bind posttranslationally modified histones. However, the C-terminal PHD-like domain in PHF6 binds double-stranded DNA. Besides DNA, PHF6 binds to a growing list of transcriptional regulators such as the nucleosome remodeling deacetylase (NuRD) complex, the RNA polymerase II–associated factor 1 (PAF1) transcriptional elongation complex, and upstream binding transcription factor, RNA polymerase I (UBTF1), a key transcriptional activator of ribosomal RNA (rRNA). The latter brings us to the nucleolar role of PHF6 and its proposed role in ribosome biogenesis. Besides interacting with UBTF1, PHF6 binds the ribosomal DNA (rDNA) promotor and rDNA-coding sequences and modulation of PHF6 levels alters rRNA synthesis rates. Todd and colleagues propose that the N-terminal PHD-like domain may mediate binding to rRNA whereas the C-terminal domain may bind rDNA. As such, PHF6 may act as a scaffold, bridging rDNA transcriptional elongation with early processing of produced rRNA. Finally, PHF6 seems involved in the DNA damage response and cell-cycle regulation: PHF6 is a substrate of the DNA damage checkpoint kinase ataxia telangiectasia mutated (ATM) and loss of PHF6 results in accumulation of phosphorylated γH2AX, a mark of DNA double-strand breaks, and in G2/M cell-cycle arrest.49  Further studies are, however, required to investigate how inactivation of PHF6 in T-ALL promotes leukemia.

Enzymes involved in regulating methylation of histone 3 lysine 27 (H3K27) are also implicated in T-ALL. Loss-of-function defects in the polycomb-repressive complex 2 (PRC2) components EZH2, SUZ12, and EED suggest a tumor suppressor role for this complex in T-ALL,4,46,50  which is supported by accelerated leukemia onset in mice upon EZH2 downregulation.46  PRC2 catalyzes H3K27 trimethylation (H3K27me3) and its tumor suppressor role may be explained by antagonism with NOTCH1: NOTCH1 activation causes eviction of PRC2 and recruitment of the lysine demethylase 6B (KDM6B, also known as JMJD3) at NOTCH1 target gene promoters, resulting in loss of the repressive H3K27me3 chromatin modification and activation of these genes.46,51  Also, the lysine demethylase 6A (KDM6A, also called UTX), which demethylates H3K27me3, shows inactivating lesions in T-ALL,50-52  and its downregulation accelerates NOTCH1-driven leukemia in mice.51,52  Interestingly, KDM6A binds TAL1, and it is recruited to TAL1 target genes to remove repressive H3K27me3 marks and activate genes. In TAL1+ leukemias, KDM6A behaves as a proto-oncogene and is required for leukemia maintenance, which may explain why KDM6A inactivation so far has not been detected in TAL1+ cases.53  The mechanism by which KDM6A plays a tumor suppressor role in TAL1 T-ALL is currently not understood. The KDM6B and KDM6A inhibitor GSKJ4 shows promising preclinical results in T-ALL. Whereas the Brand group reported specificity of this drug for the TAL1+ T-ALL subgroup, the Aifantis group observed sensitivity in all tested T-ALL cell lines and samples.51,53 

A few new oncogenes and tumor suppressors recently emerged in T-ALL. Somatic mutations in ribosomal protein genes RPL5, RPL10, and RPL22 have been described in ∼20% of T-ALL cases.50,54  Whereas RPL5 and RPL22 show heterozygous inactivating mutations and deletions, RPL10 contains an intriguing mutational hotspot at residue arginine 98 (R98), with 8% of pediatric T-ALL patients harboring an R98S RPL10 missense mutation. Inactivation of Rpl22 accelerates myristoylated Akt-driven T-cell lymphoma in mice.54,55  Mechanistically, heterozygous Rpl22 deletion activates NF-κB and induces stemness factor Lin28B.54  The roles of the RPL5 and RPL10 R98S defects in T-ALL development remain to be determined. Characterization of RPL10 R98S in yeast revealed that it impairs ribosome formation and translation fidelity. Over time, compensatory mutations are acquired, restoring ribosome biogenesis but not translation fidelity, which may drive expression of an oncogenic protein profile.56,57  Translational deregulation in T-ALL may, however, go far beyond mutations in ribosomal proteins: PHF6 has been linked to ribosome biogenesis, major T-ALL oncogenes and tumor suppressors such as PTEN, NOTCH1, and FBXW7 regulate translation, and T-ALL cells are sensitive to translation inhibitors 4EGI-1 and silvestrol.58-60 

Another intriguing novel tumor suppressor that is inactivated in 8% of adult T-ALL samples is CCR4-NOT transcription complex subunit 3 (CNOT3).50  The CCR4-NOT complex catalyzes messenger RNA (mRNA) deadenylation and may also link again to control of protein translation. In addition, roles for CNOT3 as transcriptional regulator have been proposed.61  Using studies in a fruitfly eye cancer model that is driven by Notch activation, we confirmed that NOT3 (CNOT3 homolog in the fruitfly) is a tumor suppressor gene, but more studies are required to determine how mutations in CNOT3 contribute to leukemia development. It is interesting to note that somatic mutations in CNOT3 have now also been observed in chronic lymphocytic leukemia (CLL) and in some solid tumors.62 

Inactivation of Dicer1, an essential component of the micro-RNA (miRNA) processing machinery, impairs NOTCH1-driven T-ALL development in mice and can induce regression of established NOTCH1-driven tumors, indicating a role for 1 or more miRNAs in (at least NOTCH1-driven) T-ALL.63  Indeed, many miRNAs are misexpressed in T-ALL, either due to a translocation, as key downstream targets of T-ALL oncogenes such as NOTCH1 or TAL1, or via unknown mechanisms.64-67  Oncogenic miRNAs (onco-miRs) have been identified downregulating the expression of known T-ALL–associated tumor suppressor genes. Examples are miR-19b, miR-20a, miR-26a, miR-92, and miR-223 which cooperatively downregulate IKZF1 (or IKAROS), PTEN, BIM, PHF6, NF1, and FBXW7 transcripts.68  Other miRNAs are underexpressed in T-ALL, leading to overexpression of T-ALL–associated oncogenes. An example here is miR-193b, which regulates the expression of MYB and the antiapoptotic factor MCL1.69  Many additional onco-miRs and tumor-suppressive miRNAs have been described in T-ALL. For a partial overview, we refer to Aster.66 

A related field of interest that recently emerged in T-ALL is long noncoding RNAs (lncRNAs), a heterogeneous class of transcripts defined by a minimum length of 200 nucleotides and an apparent lack of protein-coding potential. Diverse molecular mechanisms have been described by which lncNRAs control a wide variety of cellular functions and developmental processes, and promote disease pathogenesis, including cancer.70  Interestingly, different T-ALL subgroups are characterized by distinct lncRNA expression profiles.71  In addition, a series of NOTCH1-regulated lncRNAs were identified.72,73  The best characterized one is leukemia-induced noncoding activator RNA 1 (LUNAR1), which is a NOTCH1-regulated pro-oncogenic lncRNA that is required for efficient growth of T-ALL cells by maintaining high expression of insulin-like growth factor 1 receptor (IGF1R).73 

Recent sequencing studies have suggested that on average 10 to 20 protein-altering mutations are present in T-ALL cells.4,50,74  This accumulation of mutations does not occur randomly, and specific combinations of mutations are often found, suggesting that those activated oncogenes and inactivated tumor suppressor genes are physiologically interconnected and cooperate during the development and progression of the leukemia (Figure 2).75  Initial founder genomic lesions initiate a premalignant process by interacting with the existing machinery of the physiological cell state76  and additional mutations are necessary to drive transformation75,77  and lead to the development of subclonal variegation.

This has been very nicely observed in TAL1/LMO2+ T-ALL cases, where PTEN inactivation occurs most frequently. Interestingly, transplantation of TAL1 rearranged leukemia cells into immune-deficient mice allowed development of leukemic subclones with newly acquired PTEN microdeletions.32  These data suggest that there is an enormous pressure for TAL1-expressing T-ALL cells to inactivate PTEN, and that TAL1 expression and PTEN inactivation cooperate during T-cell transformation. In contrast, components of the IL7R-JAK signaling pathway are frequently mutated in immature T-ALL cases or in TLX/HOXA+ cases, but are underrepresented in TAL1/LMO2+ cases. In addition, we observed that mutations of IL7R-JAK are positively associated with mutations and deletions of members of the PRC2 complex (EZH2, SUZ12 and EED).21  Similarly, WT1 mutations are most prevalent in TLX3+ cases.78  These and other genomic data suggest that initiating lesions leading to ectopic transcription factor expression sensitize the cells to alterations of very specific pathways.

Current treatment of T-ALL consists of high-intensity combination chemotherapy and results in a very high overall survival for pediatric patients.79  Unfortunately, this treatment comes with significant short-term and long-term side effects. Especially for young children, the effects of this high-dose chemotherapy on bone development, the central nervous system, and fertility should not be underestimated.80  A complete overview of the treatment options was recently described.81,82 

In addition to the side effects, the occurrence of relapse is another important challenge as it is observed in up to 20% of pediatric and 40% of adult T-ALL.83  Relapsed T-ALL cases are often refractory to chemotherapeutics and are associated with a poor prognosis.84,85  From comparative genetic studies conducted on T-ALL samples taken at diagnosis and relapse, and studies with xenotransplantation in mice, it is clear that relapse is determined by the clonal evolution of either a minor genetic subclone present in the primary leukemia, or from the clonal expansion of an ancestral cell (prediagnosis).86-89  However, some rare T-ALL late-relapsed cases are considered to be secondary malignancies rather than originated by the initial disease.90 

The use of whole-exome sequencing has provided a way to take a detailed view of the genomes of relapsed T-ALL cases and led to the identification of activating mutations in cytosolic 5′-nucleotidase II (NT5C2) as one of the causes of relapse.86,87  NT5C2 is an enzyme responsible for the inactivation of nucleoside-analog chemotherapy drugs.91  When NT5C2 mutant proteins were expressed in T-ALL lymphoblast, increased nucleotidase activity was observed and the cells became resistant to 6-mercaptopurine (6-MP) and 6-thioguanine (6-TG), 2 chemotherapeutic drugs used in maintenance treatment of T-ALL, suggesting the important role of the identified NT5C2-activating mutations in therapy resistance.86 

To further improve the treatment of T-ALL and to reduce the toxicity of current treatment, we await the introduction of newer targeted agents. Recent drug development in both oncology and immunology has generated a spectrum of new drugs that could be useful for the treatment of specific T-ALL subsets (Table 3).

Table 3.

Promising targeted agents in T-ALL and their stage in clinical testing

CompoundSpecificityClinical phaseDisease
NOTCH1 signaling    
 MK-0752 GSI Phase 1 T-ALL and lymphoma 
 PF-03084014 GSI Phase 1 Advanced-stage cancers, T-ALL, lymphoblastic lymphoma 
 BMS-906024 (with dexamethasone) GSI Phase 1 T-ALL or T-cell lymphoblastic lymphoma 
 OMP-52M51 Notch1 inhibitory antibody Phase 1 Dose-escalation study in lymphoid malignancy 
 LY3039478 Notch1 inhibitor Phase 1 Dose-escalation study in advanced-stage cancers 
 BMS-536924 ATP-competitive IGF-1R/IR inhibitor Preclinical  
JAK-STAT inhibitors    
 Ruxolitinib JAK1/2 inhibitor Phase 1 AML 
Phase 2 B-ALL 
FDA approved Myelofibrosis 
Phase 1/2 Fallopian tube cancer, ovarian cancer, primary peritoneal cancer 
 Tofacitinib JAK3 inhibitor FDA approved Rheumatoid arthritis 
Phase 1 Systemic lupus erythematosus 
Phase 3 Juvenile idiopathic arthritis 
 Pimozide STAT5 inhibitor FDA approved Schizophrenia, psychotic disorders 
Phase 2 Amyotrophic lateral sclerosis 
 17-AAG HSP90 inhibitor Phase 3 Multiple myeloma 
Phase 2 Pancreatic cancer 
Phase 2 Advanced malignancies 
Phase 2 Ovarian cancer 
 PU-H71 HSP90 inhibitor Phase 1 Solid tumor, lymphoma 
Phase 1 Metastatic solid tumor, lymphoma, myeloproliferative neoplasms 
Inducers of apoptosis    
 ABT-263 (Navitoclax) Inhibitor of Bcl-xL, Bcl-2, and Bcl-w Phase 1 Non-small cell lung cancer 
Phase 2 Platinum-resistant or refractory ovarian cancer 
Phase 1 Hepatocellular carcinoma 
Phase 1/2 Melanoma 
 ABT-199 (Venetoclax) Bcl-2–selective inhibitor FDA approved CLL 
Phase 3 Relapsed/refractory multiple myeloma 
PI3K inhibitors    
 CAL-130 Inhibits p110γ and p110δ catalytic domains Preclinical  
 Ly294002 Inhibit PI3Kα/δ/β Phase 1 Neuroblastoma 
 Pictilisib (GDC-0941) Inhibitor of PI3Kα/δ Phase 2 Breast cancer 
Phase 2 Nonsquamous non-small cell lung cancer 
 Apitolisib (GDC-0980, RG7422) Inhibitor for PI3Kα/β/δ/γ Phase 2 Endometrial carcinoma 
Phase 1/2 Prostate cancer 
Phase 2 Renal cell carcinoma 
MEK inhibitors    
 CI-1040 ATP noncompetitive MEK1/2 inhibitor Phase 2 Breast cancer, colorectal cancer, lung cancer, pancreatic cancer 
 Selumetinib (AZD6244) MEK1 inhibitor Phase 2 Triple-negative breast cancer 
Phase 1 Lung cancer, melanoma, head and neck carcinoma, gastroesophageal cancer, breast cancer, pancreatic adenocarcinoma, colorectal cancer 
 Trametinib (GSK1120212) MEK1/2 inhibitor Phase 2 Gastrointestinal stromal tumors 
Phase 1 Melanoma 
Phase 1 Neuroblastoma 
Phase 2 Non-small cell lung cancer 
ABL inhibitors    
 Imatinib v-Abl, c-Kit, and PDGFR inhibitor FDA approved Chronic myeloid leukemia 
Phase 2 B-ALL, B lymphoblastic lymphoma, T-ALL, T lymphoblastic lymphoma 
Phase 3 Philadelphia chromosome–positive adult ALL 
 Dasatinib Abl, Src, and c-Kit inhibitor FDA approved Chronic myeloid leukemia 
Phase 2 ALL 
Phase 1 Chronic kidney disease 
 Nilotinib Abl, Src, and c-Kit inhibitor Phase 2 Glioma 
Phase 3 Philadelphia chromosome–positive adult ALL 
FDA approved Chronic myeloid leukemia 
Hedgehog inhibitors    
 Vismodegib (GDC-0449) SMO inhibitor FDA approved Basal cell carcinoma 
Phase 2 Breast cancer 
Phase 2 Non-Hodgkin lymphoma, multiple myeloma, advanced solid tumors 
 GANT61 GLI1 inhibitor Preclinical  
Translation inhibitors    
 Rapamycin Specific mTOR inhibitor Phase 2 Myelodysplastic syndrome, CLL, ALL, T lymphoblastic lymphoma, acute myelogenous leukemia, acute biphenotypic leukemia, acute undifferentiated leukemia 
Phase 1 Fallopian tube carcinoma, ovarian carcinoma, primary peritoneal carcinoma 
 4EGI-1 Competitive eIF4E/eIF4G interaction inhibitor Preclinical  
CompoundSpecificityClinical phaseDisease
NOTCH1 signaling    
 MK-0752 GSI Phase 1 T-ALL and lymphoma 
 PF-03084014 GSI Phase 1 Advanced-stage cancers, T-ALL, lymphoblastic lymphoma 
 BMS-906024 (with dexamethasone) GSI Phase 1 T-ALL or T-cell lymphoblastic lymphoma 
 OMP-52M51 Notch1 inhibitory antibody Phase 1 Dose-escalation study in lymphoid malignancy 
 LY3039478 Notch1 inhibitor Phase 1 Dose-escalation study in advanced-stage cancers 
 BMS-536924 ATP-competitive IGF-1R/IR inhibitor Preclinical  
JAK-STAT inhibitors    
 Ruxolitinib JAK1/2 inhibitor Phase 1 AML 
Phase 2 B-ALL 
FDA approved Myelofibrosis 
Phase 1/2 Fallopian tube cancer, ovarian cancer, primary peritoneal cancer 
 Tofacitinib JAK3 inhibitor FDA approved Rheumatoid arthritis 
Phase 1 Systemic lupus erythematosus 
Phase 3 Juvenile idiopathic arthritis 
 Pimozide STAT5 inhibitor FDA approved Schizophrenia, psychotic disorders 
Phase 2 Amyotrophic lateral sclerosis 
 17-AAG HSP90 inhibitor Phase 3 Multiple myeloma 
Phase 2 Pancreatic cancer 
Phase 2 Advanced malignancies 
Phase 2 Ovarian cancer 
 PU-H71 HSP90 inhibitor Phase 1 Solid tumor, lymphoma 
Phase 1 Metastatic solid tumor, lymphoma, myeloproliferative neoplasms 
Inducers of apoptosis    
 ABT-263 (Navitoclax) Inhibitor of Bcl-xL, Bcl-2, and Bcl-w Phase 1 Non-small cell lung cancer 
Phase 2 Platinum-resistant or refractory ovarian cancer 
Phase 1 Hepatocellular carcinoma 
Phase 1/2 Melanoma 
 ABT-199 (Venetoclax) Bcl-2–selective inhibitor FDA approved CLL 
Phase 3 Relapsed/refractory multiple myeloma 
PI3K inhibitors    
 CAL-130 Inhibits p110γ and p110δ catalytic domains Preclinical  
 Ly294002 Inhibit PI3Kα/δ/β Phase 1 Neuroblastoma 
 Pictilisib (GDC-0941) Inhibitor of PI3Kα/δ Phase 2 Breast cancer 
Phase 2 Nonsquamous non-small cell lung cancer 
 Apitolisib (GDC-0980, RG7422) Inhibitor for PI3Kα/β/δ/γ Phase 2 Endometrial carcinoma 
Phase 1/2 Prostate cancer 
Phase 2 Renal cell carcinoma 
MEK inhibitors    
 CI-1040 ATP noncompetitive MEK1/2 inhibitor Phase 2 Breast cancer, colorectal cancer, lung cancer, pancreatic cancer 
 Selumetinib (AZD6244) MEK1 inhibitor Phase 2 Triple-negative breast cancer 
Phase 1 Lung cancer, melanoma, head and neck carcinoma, gastroesophageal cancer, breast cancer, pancreatic adenocarcinoma, colorectal cancer 
 Trametinib (GSK1120212) MEK1/2 inhibitor Phase 2 Gastrointestinal stromal tumors 
Phase 1 Melanoma 
Phase 1 Neuroblastoma 
Phase 2 Non-small cell lung cancer 
ABL inhibitors    
 Imatinib v-Abl, c-Kit, and PDGFR inhibitor FDA approved Chronic myeloid leukemia 
Phase 2 B-ALL, B lymphoblastic lymphoma, T-ALL, T lymphoblastic lymphoma 
Phase 3 Philadelphia chromosome–positive adult ALL 
 Dasatinib Abl, Src, and c-Kit inhibitor FDA approved Chronic myeloid leukemia 
Phase 2 ALL 
Phase 1 Chronic kidney disease 
 Nilotinib Abl, Src, and c-Kit inhibitor Phase 2 Glioma 
Phase 3 Philadelphia chromosome–positive adult ALL 
FDA approved Chronic myeloid leukemia 
Hedgehog inhibitors    
 Vismodegib (GDC-0449) SMO inhibitor FDA approved Basal cell carcinoma 
Phase 2 Breast cancer 
Phase 2 Non-Hodgkin lymphoma, multiple myeloma, advanced solid tumors 
 GANT61 GLI1 inhibitor Preclinical  
Translation inhibitors    
 Rapamycin Specific mTOR inhibitor Phase 2 Myelodysplastic syndrome, CLL, ALL, T lymphoblastic lymphoma, acute myelogenous leukemia, acute biphenotypic leukemia, acute undifferentiated leukemia 
Phase 1 Fallopian tube carcinoma, ovarian carcinoma, primary peritoneal carcinoma 
 4EGI-1 Competitive eIF4E/eIF4G interaction inhibitor Preclinical  

ATP, adenosine triphosphate; FDA, US Food and Drug Administration; HSP90, heat shock protein 90; PDGFR, platelet-derived growth factor receptor; SMO, smoothened.

With NOTCH1 being the major oncogene in T-ALL and with drugs available that can interfere with the activation (cleavage) of NOTCH1 by the γ-secretase complex, a number of clinical trials have been initiated with γ-secretase inhibitors (GSIs). However, these trials led to disappointing clinical results due to the dose-limiting toxicity and low response rates. It remains to be determined how second-generation GSIs will perform, but at least some promising anecdotal responses have already been reported and with milder toxicities.92-95  Besides GSIs, other therapeutic approaches for the inhibition of NOTCH1 have been developed. Monoclonal antibodies against the NOTCH1 receptor have antitumor effects in vitro and in vivo with limited gastrointestinal toxicities.96,97  Inhibition of ADAM10 may also facilitate effective inhibition of wild-type and mutant NOTCH receptors.98  An antibody against the γ-secretase complex (A5226A) showed preclinical activity against T-ALL.99  Mastermind-inhibiting peptides that mimic the NOTCH1 interaction with MAML (SAMH1 peptides) are also being tested.100,101 

Because both JAK and ABL1 kinases are often activated in T-ALL, available JAK inhibitors (ruxolitinib and tofacitinib) and ABL1 inhibitors (imatinib, dasatinib, nilotinib) could be repurposed for the treatment of T-ALL cases with documented mutations leading to JAK/STAT or ABL1 activation. Several preclinical studies have demonstrated activity of ruxolitinib or tofacitinib for the inhibition of T-ALL cells with IL7R or JAK1/JAK3 mutations, whereas case reports have shown some activity of imatinib or dasatinib for the treatment of NUP214-ABL1+ T-ALL.4,23,102,103  The therapeutic potential of targeting the IL7Rα signaling and the use of JAK inhibitors in ALL has been recently reviewed.104,105 

Similarly, recently approved hedgehog pathway inhibitors could show activity in T-ALL. The hedgehog signaling pathway plays an important role in normal T-cell development, which is steered by the secretion of the sonic hedgehog ligand (SHH) by certain thymic epithelial cells.106  In T-ALL, the hedgehog pathway was reported to be aberrantly activated by rare mutations or ectopic expression of hedgehog pathway genes, and those cases showed response to hedgehog inhibitor treatment.50,107,108  Finally, some other attractive novel therapies have recently been described in T-ALL. Selective inhibitors of nuclear export (SINE) were shown to be highly toxic to T-ALL and acute myeloid leukemia (AML) cells in mouse xenograft models, while having little toxic effects on normal mouse hematopoietic cells.109  In addition, it was demonstrated that TCR+ T-ALL cells can be targeted by mimicking thymic-negative selection as obtained by TCR stimulation via antigen/major histocompatibility complex (MHC) presentation or via an antibody against CD3.110  The advantage of these types of therapies is that they seem independent of genetic subtypes, avoiding extensive genetic characterization before therapeutic choices are to be made.

In the past decades, enormous progress was made in our understanding of the genetics and biology of T-ALL, but there are still significant gaps in our knowledge. To date, almost all attention has gone to defects affecting protein coding genes. The recent identification of mutations in noncoding regions of the genome that result in aberrant transcription factor expression and the discovery of several miRNAs and lncRNAs with a pathogenic role in T-ALL underscore that the “noncoding genome” should not be neglected and that novel classes of oncogenes and tumor suppressor genes may remain to be discovered.

Genomic screens have identified recurrent novel oncogenes and tumor suppressors, such as PHF6, RPL10, and CNOT3, for which the exact role in the pathogenesis of T-ALL remains poorly understood. Moreover, genome-wide screening has mainly been limited to diagnostic cases and such screens of large patient cohorts are lacking for relapse T-ALL. A better understanding of the biology of relapse is an absolute requirement to improve the dismal prognosis these patients currently are facing.

Analysis of mutational patterns in large patient cohorts revealed that the mutational landscape of T-ALL is not random and that particular defects often co-occur or are mutually exclusive.21  More research is needed to understand the biology behind particular mutational patterns and to characterize the specific clinical behaviors and distinct prognosis associated with these patterns. Related to this issue, the order in which mutations are acquired in T-ALL development is also most likely not random. In addition, the cellular context and exact hematopoietic developmental stage of the cell in which a mutation arises might be very relevant for the oncogenic action of the defect. At this point, not much is known on the order in which mutations are acquired and on the cell of origin in which they need to arise. Extensive modeling in mice and single-cell sequencing will be required to answer these questions.

Finally, the interaction of the leukemia cells with their microenvironment and characterization of the T-ALL niche are also essential to fully capture T-ALL pathogenesis. In this context, recently developed advanced microscopy techniques allow in vivo imaging of interactions of leukemia cells with their environment and an essential role in T-ALL maintenance has been discovered for the CXCL12 chemokine and its receptor CXCR4.111-113 

The ultimate goal of the T-ALL research community is to translate the knowledge into highly efficient and low-toxicity targeted therapies. Whereas the current chemotherapy-based regimens in pediatric T-ALL are associated with high survival rates, the toxicities of these therapies should not be underestimated. With many new targeted agents under development for cancer and autoimmune diseases (Table 3), it should be possible to effectively repurpose the best agents for the treatment of T-ALL, and to design effective combination therapies directed against the specific sets of cooperating mutations in T-ALL.

The authors thank all researchers and clinicians for their contributions to the field and apologize to those whose work they could not describe or cite.

The laboratory of K.D.K. was supported by a European Research Council (ERC) starting grant (334946), Fonds Wetenschappelijk Onderzoek (FWO)-Vlaanderen funding (G067015N and G084013N), and a Stichting Tegen Kanker grant (2012-176). The laboratory of J.C. was supported by an ERC consolidator grant (617340), and funding from FWO-Vlaanderen (G.0683.12), Stichting Tegen Kanker (2014-120), and Kom Op Tegen Kanker. T.G. was supported by the Emmanuel van der Schueren Kom Op Tegen Kanker fellowship.

Contribution: T.G., C.V., J.C., and K.D.K. contributed to data analysis and the writing of the manuscript.

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

Correspondence: Jan Cools, KU Leuven, Campus Gasthuisberg O&N4, Herestraat 49 (Box 602), 3000 Leuven, Belgium; e-mail: jan.cools@kuleuven.be; and Kim De Keersmaecker, Department of Oncology, KU Leuven, Campus Gasthuisberg O&N1, Herestraat 49 (Box 603), 3000 Leuven, Belgium; e-mail: kim.dekeersmaecker@kuleuven.be.

1.
Homminga
I
,
Pieters
R
,
Langerak
AW
, et al
.
Integrated transcript and genome analyses reveal NKX2-1 and MEF2C as potential oncogenes in T cell acute lymphoblastic leukemia
.
Cancer Cell
.
2011
;
19
(
4
):
484
-
497
.
2.
Ferrando
AA
,
Neuberg
DS
,
Staunton
J
, et al
.
Gene expression signatures define novel oncogenic pathways in T cell acute lymphoblastic leukemia
.
Cancer Cell
.
2002
;
1
(
1
):
75
-
87
.
3.
Soulier
J
,
Clappier
E
,
Cayuela
J-M
, et al
.
HOXA genes are included in genetic and biologic networks defining human acute T-cell leukemia (T-ALL)
.
Blood
.
2005
;
106
(
1
):
274
-
286
.
4.
Zhang
J
,
Ding
L
,
Holmfeldt
L
, et al
.
The genetic basis of early T-cell precursor acute lymphoblastic leukaemia
.
Nature
.
2012
;
481
(
7380
):
157
-
163
.
5.
Condorelli
GL
,
Facchiano
F
,
Valtieri
M
, et al
.
T-cell-directed TAL-1 expression induces T-cell malignancies in transgenic mice
.
Cancer Res
.
1996
;
56
(
22
):
5113
-
5119
.
6.
Kelliher
MA
,
Seldin
DC
,
Leder
P
.
Tal-1 induces T cell acute lymphoblastic leukemia accelerated by casein kinase IIalpha
.
EMBO J
.
1996
;
15
(
19
):
5160
-
5166
.
7.
Rakowski
LA
,
Lehotzky
EA
,
Chiang
MY
.
Transient responses to NOTCH and TLX1/HOX11 inhibition in T-cell acute lymphoblastic leukemia/lymphoma
.
PLoS One
.
2011
;
6
(
2
):
e16761
.
8.
De Keersmaecker
K
,
Real
PJ
,
Gatta
GD
, et al
.
The TLX1 oncogene drives aneuploidy in T cell transformation
.
Nat Med
.
2010
;
16
(
11
):
1321
-
1327
.
9.
Gerby
B
,
Tremblay
CS
,
Tremblay
M
, et al
.
SCL, LMO1 and Notch1 reprogram thymocytes into self-renewing cells
.
PLoS Genet
.
2014
;
10
(
12
):
e1004768
.
10.
Dadi
S
,
Le Noir
S
,
Payet-Bornet
D
, et al
.
TLX homeodomain oncogenes mediate T cell maturation arrest in T-ALL via interaction with ETS1 and suppression of TCRα gene expression
.
Cancer Cell
.
2012
;
21
(
4
):
563
-
576
.
11.
Navarro
J-M
,
Touzart
A
,
Pradel
LC
, et al
.
Site- and allele-specific polycomb dysregulation in T-cell leukaemia
.
Nat Commun
.
2015
;
6
:
6094
.
12.
Mansour
MR
,
Abraham
BJ
,
Anders
L
, et al
.
Oncogene regulation. An oncogenic super-enhancer formed through somatic mutation of a noncoding intergenic element
.
Science
.
2014
;
346
(
6215
):
1373
-
1377
.
13.
Radtke
F
,
Wilson
A
,
Stark
G
, et al
.
Deficient T cell fate specification in mice with an induced inactivation of Notch1
.
Immunity
.
1999
;
10
(
5
):
547
-
558
.
14.
Radtke
F
,
MacDonald
HR
,
Tacchini-Cottier
F
.
Regulation of innate and adaptive immunity by Notch
.
Nat Rev Immunol
.
2013
;
13
(
6
):
427
-
437
.
15.
Weng
AP
,
Ferrando
AA
,
Lee
W
, et al
.
Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia
.
Science
.
2004
;
306
(
5694
):
269
-
271
.
16.
O’Neil
J
,
Grim
J
,
Strack
P
, et al
.
FBW7 mutations in leukemic cells mediate NOTCH pathway activation and resistance to gamma-secretase inhibitors
.
J Exp Med
.
2007
;
204
(
8
):
1813
-
1824
.
17.
Thompson
BJ
,
Buonamici
S
,
Sulis
ML
, et al
.
The SCFFBW7 ubiquitin ligase complex as a tumor suppressor in T cell leukemia
.
J Exp Med
.
2007
;
204
(
8
):
1825
-
1835
.
18.
Sanchez-Martin
M
,
Ferrando
A
.
The NOTCH1-MYC highway toward T-cell acute lymphoblastic leukemia
.
Blood
.
2017
;
129
(
9
):
1124
-
1133
.
19.
Takada
K
,
Jameson
SC
.
Naive T cell homeostasis: from awareness of space to a sense of place
.
Nat Rev Immunol
.
2009
;
9
(
12
):
823
-
832
.
20.
Mazzucchelli
R
,
Durum
SK
.
Interleukin-7 receptor expression: intelligent design
.
Nat Rev Immunol
.
2007
;
7
(
2
):
144
-
154
.
21.
Vicente
C
,
Schwab
C
,
Broux
M
, et al
.
Targeted sequencing identifies associations between IL7R-JAK mutations and epigenetic modulators in T-cell acute lymphoblastic leukemia
.
Haematologica
.
2015
;
100
(
10
):
1301
-
1310
.
22.
Canté-Barrett
K
,
Spijkers-Hagelstein
JA
,
Buijs-Gladdines
JG
, et al
.
MEK and PI3K-AKT inhibitors synergistically block activated IL7 receptor signaling in T-cell acute lymphoblastic leukemia
.
Leukemia
.
2016
;
30
(
9
):
1832
-
1843
.
23.
Zenatti
PP
,
Ribeiro
D
,
Li
W
, et al
.
Oncogenic IL7R gain-of-function mutations in childhood T-cell acute lymphoblastic leukemia
.
Nat Genet
.
2011
;
43
(
10
):
932
-
939
.
24.
Porcu
M
,
Kleppe
M
,
Gianfelici
V
, et al
.
Mutation of the receptor tyrosine phosphatase PTPRC (CD45) in T-cell acute lymphoblastic leukemia
.
Blood
.
2012
;
119
(
19
):
4476
-
4479
.
25.
Kleppe
M
,
Lahortiga
I
,
El Chaar
T
, et al
.
Deletion of the protein tyrosine phosphatase gene PTPN2 in T-cell acute lymphoblastic leukemia
.
Nat Genet
.
2010
;
42
(
6
):
530
-
535
.
26.
Tremblay
CS
,
Brown
FC
,
Collett
M
, et al
.
Loss-of-function mutations of Dynamin 2 promote T-ALL by enhancing IL-7 signalling
.
Leukemia
.
2016
;
30
(
10
):
1993
-
2001
.
27.
Goossens
S
,
Radaelli
E
,
Blanchet
O
, et al
.
ZEB2 drives immature T-cell lymphoblastic leukaemia development via enhanced tumour-initiating potential and IL-7 receptor signalling
.
Nat Commun
.
2015
;
6
:
5794
.
28.
Lacronique
V
,
Boureux
A
,
Valle
VD
, et al
.
A TEL-JAK2 fusion protein with constitutive kinase activity in human leukemia
.
Science
.
1997
;
278
(
5341
):
1309
-
1312
.
29.
Sanda
T
,
Tyner
JW
,
Gutierrez
A
, et al
.
TYK2-STAT1-BCL2 pathway dependence in T-cell acute lymphoblastic leukemia
.
Cancer Discov
.
2013
;
3
(
5
):
564
-
577
.
30.
Zhao
L
,
Vogt
PK
.
Class I PI3K in oncogenic cellular transformation
.
Oncogene
.
2008
;
27
(
41
):
5486
-
5496
.
31.
Cully
M
,
You
H
,
Levine
AJ
,
Mak
TW
.
Beyond PTEN mutations: the PI3K pathway as an integrator of multiple inputs during tumorigenesis
.
Nat Rev Cancer
.
2006
;
6
(
3
):
184
-
192
.
32.
Mendes
RD
,
Sarmento
LM
,
Canté-Barrett
K
, et al
.
PTEN microdeletions in T-cell acute lymphoblastic leukemia are caused by illegitimate RAG-mediated recombination events
.
Blood
.
2014
;
124
(
4
):
567
-
578
.
33.
Palomero
T
,
Sulis
ML
,
Cortina
M
, et al
.
Mutational loss of PTEN induces resistance to NOTCH1 inhibition in T-cell leukemia
.
Nat Med
.
2007
;
13
(
10
):
1203
-
1210
.
34.
Zuurbier
L
,
Petricoin
EF
III
,
Vuerhard
MJ
, et al
.
The significance of PTEN and AKT aberrations in pediatric T-cell acute lymphoblastic leukemia
.
Haematologica
.
2012
;
97
(
9
):
1405
-
1413
.
35.
Gutierrez
A
,
Sanda
T
,
Grebliunaite
R
, et al
.
High frequency of PTEN, PI3K, and AKT abnormalities in T-cell acute lymphoblastic leukemia
.
Blood
.
2009
;
114
(
3
):
647
-
650
.
36.
Silva
A
,
Yunes
JA
,
Cardoso
BA
, et al
.
PTEN posttranslational inactivation and hyperactivation of the PI3K/Akt pathway sustain primary T cell leukemia viability
.
J Clin Invest
.
2008
;
118
(
11
):
3762
-
3774
.
37.
Neumann
M
,
Vosberg
S
,
Schlee
C
, et al
.
Mutational spectrum of adult T-ALL
.
Oncotarget
.
2015
;
6
(
5
):
2754
-
2766
.
38.
Barata
JT
,
Silva
A
,
Brandao
JG
,
Nadler
LM
,
Cardoso
AA
,
Boussiotis
VA
.
Activation of PI3K is indispensable for interleukin 7-mediated viability, proliferation, glucose use, and growth of T cell acute lymphoblastic leukemia cells
.
J Exp Med
.
2004
;
200
(
5
):
659
-
669
.
39.
Graux
C
,
Cools
J
,
Melotte
C
, et al
.
Fusion of NUP214 to ABL1 on amplified episomes in T-cell acute lymphoblastic leukemia
.
Nat Genet
.
2004
;
36
(
10
):
1084
-
1089
.
40.
De Keersmaecker
K
,
Versele
M
,
Cools
J
,
Superti-Furga
G
,
Hantschel
O
.
Intrinsic differences between the catalytic properties of the oncogenic NUP214-ABL1 and BCR-ABL1 fusion protein kinases
.
Leukemia
.
2008
;
22
(
12
):
2208
-
2216
.
41.
De Keersmaecker
K
,
Rocnik
JL
,
Bernad
R
, et al
.
Kinase activation and transformation by NUP214-ABL1 is dependent on the context of the nuclear pore
.
Mol Cell
.
2008
;
31
(
1
):
134
-
142
.
42.
De Keersmaecker
K
,
Porcu
M
,
Cox
L
, et al
.
NUP214-ABL1-mediated cell proliferation in T-cell acute lymphoblastic leukemia is dependent on the LCK kinase and various interacting proteins
.
Haematologica
.
2014
;
99
(
1
):
85
-
93
.
43.
Oshima
K
,
Khiabanian
H
,
da Silva-Almeida
AC
, et al
.
Mutational landscape, clonal evolution patterns, and role of RAS mutations in relapsed acute lymphoblastic leukemia
.
Proc Natl Acad Sci USA
.
2016
;
113
(
40
):
11306
-
11311
.
44.
Huether
R
,
Dong
L
,
Chen
X
, et al
.
The landscape of somatic mutations in epigenetic regulators across 1,000 paediatric cancer genomes
.
Nat Commun
.
2014
;
5
:
3630
.
45.
Van Vlierberghe
P
,
Palomero
T
,
Khiabanian
H
, et al
.
PHF6 mutations in T-cell acute lymphoblastic leukemia
.
Nat Genet
.
2010
;
42
(
4
):
338
-
342
.
46.
Ntziachristos
P
,
Tsirigos
A
,
Van Vlierberghe
P
, et al
.
Genetic inactivation of the polycomb repressive complex 2 in T cell acute lymphoblastic leukemia
.
Nat Med
.
2012
;
18
(
2
):
298
-
301
.
47.
Van der Meulen
J
,
Van Roy
N
,
Van Vlierberghe
P
,
Speleman
F
.
The epigenetic landscape of T-cell acute lymphoblastic leukemia
.
Int J Biochem Cell Biol
.
2014
;
53
:
547
-
557
.
48.
Peirs
S
,
Van der Meulen
J
,
Van de Walle
I
, et al
.
Epigenetics in T-cell acute lymphoblastic leukemia
.
Immunol Rev
.
2015
;
263
(
1
):
50
-
67
.
49.
Todd
MA
,
Ivanochko
D
,
Picketts
DJ
.
PHF6 degrees of separation: the multifaceted roles of a chromatin adaptor protein
.
Genes (Basel)
.
2015
;
6
(
2
):
325
-
352
.
50.
De Keersmaecker
K
,
Atak
ZK
,
Li
N
, et al
.
Exome sequencing identifies mutation in CNOT3 and ribosomal genes RPL5 and RPL10 in T-cell acute lymphoblastic leukemia
.
Nat Genet
.
2013
;
45
(
2
):
186
-
190
.
51.
Ntziachristos
P
,
Tsirigos
A
,
Welstead
GG
, et al
.
Contrasting roles of histone 3 lysine 27 demethylases in acute lymphoblastic leukaemia
.
Nature
.
2014
;
514
(
7523
):
513
-
517
.
52.
Van der Meulen
J
,
Sanghvi
V
,
Mavrakis
K
, et al
.
The H3K27me3 demethylase UTX is a gender-specific tumor suppressor in T-cell acute lymphoblastic leukemia
.
Blood
.
2015
;
125
(
1
):
13
-
21
.
53.
Benyoucef
A
,
Palii
CG
,
Wang
C
, et al
.
UTX inhibition as selective epigenetic therapy against TAL1-driven T-cell acute lymphoblastic leukemia
.
Genes Dev
.
2016
;
30
(
5
):
508
-
521
.
54.
Rao
S
,
Lee
S-Y
,
Gutierrez
A
, et al
.
Inactivation of ribosomal protein L22 promotes transformation by induction of the stemness factor, Lin28B
.
Blood
.
2012
;
120
(
18
):
3764
-
3773
.
55.
Rao
S
,
Cai
KQ
,
Stadanlick
JE
, et al
.
Ribosomal protein Rpl22 controls the dissemination of T-cell lymphoma
.
Cancer Res
.
2016
;
76
(
11
):
3387
-
3396
.
56.
Sulima
SO
,
Patchett
S
,
Advani
VM
,
De Keersmaecker
K
,
Johnson
AW
,
Dinman
JD
.
Bypass of the pre-60S ribosomal quality control as a pathway to oncogenesis
.
Proc Natl Acad Sci USA
.
2014
;
111
(
15
):
5640
-
5645
.
57.
De Keersmaecker
K
,
Sulima
SO
,
Dinman
JD
.
Ribosomopathies and the paradox of cellular hypo- to hyperproliferation
.
Blood
.
2015
;
125
(
9
):
1377
-
1382
.
58.
Schwarzer
A
,
Holtmann
H
,
Brugman
M
, et al
.
Hyperactivation of mTORC1 and mTORC2 by multiple oncogenic events causes addiction to eIF4E-dependent mRNA translation in T-cell leukemia
.
Oncogene
.
2015
;
34
(
27
):
3593
-
3604
.
59.
Wolfe
AL
,
Singh
K
,
Zhong
Y
, et al
.
RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer
.
Nature
.
2014
;
513
(
7516
):
65
-
70
.
60.
Girardi
T
,
De Keersmaecker
K
.
T-ALL: ALL a matter of translation?
Haematologica
.
2015
;
100
(
3
):
293
-
295
.
61.
Shirai
Y-T
,
Suzuki
T
,
Morita
M
,
Takahashi
A
,
Yamamoto
T
.
Multifunctional roles of the mammalian CCR4-NOT complex in physiological phenomena
.
Front Genet
.
2014
;
5
:
286
.
62.
Puente
XS
,
Beà
S
,
Valdés-Mas
R
, et al
.
Non-coding recurrent mutations in chronic lymphocytic leukaemia
.
Nature
.
2015
;
526
(
7574
):
519
-
524
.
63.
Junker
F
,
Chabloz
A
,
Koch
U
,
Radtke
F
.
Dicer1 imparts essential survival cues in Notch-driven T-ALL via miR-21-mediated tumor suppressor Pdcd4 repression
.
Blood
.
2015
;
126
(
8
):
993
-
1004
.
64.
Li
X
,
Sanda
T
,
Look
AT
,
Novina
CD
,
von Boehmer
H
.
Repression of tumor suppressor miR-451 is essential for NOTCH1-induced oncogenesis in T-ALL
.
J Exp Med
.
2011
;
208
(
4
):
663
-
675
.
65.
Mavrakis
KJ
,
Wolfe
AL
,
Oricchio
E
, et al
.
Genome-wide RNA-mediated interference screen identifies miR-19 targets in Notch-induced T-cell acute lymphoblastic leukaemia
.
Nat Cell Biol
.
2010
;
12
(
4
):
372
-
379
.
66.
Aster
JC
.
Dicing up T-ALL
.
Blood
.
2015
;
126
(
8
):
929
-
930
.
67.
Mansour
MR
,
Sanda
T
,
Lawton
LN
, et al
.
The TAL1 complex targets the FBXW7 tumor suppressor by activating miR-223 in human T cell acute lymphoblastic leukemia
.
J Exp Med
.
2013
;
210
(
8
):
1545
-
1557
.
68.
Mavrakis
KJ
,
Van Der Meulen
J
,
Wolfe
AL
, et al
.
A cooperative microRNA-tumor suppressor gene network in acute T-cell lymphoblastic leukemia (T-ALL) [published correction appears in Nat Genet. 2011;43(8):815]
.
Nat Genet
.
2011
;
43
(
7
):
673
-
678
.
69.
Mets
E
,
Van der Meulen
J
,
Van Peer
G
, et al
.
MicroRNA-193b-3p acts as a tumor suppressor by targeting the MYB oncogene in T-cell acute lymphoblastic leukemia
.
Leukemia
.
2015
;
29
(
4
):
798
-
806
.
70.
Schmitz
SU
,
Grote
P
,
Herrmann
BG
.
Mechanisms of long noncoding RNA function in development and disease
.
Cell Mol Life Sci
.
2016
;
73
(
13
):
2491
-
2509
.
71.
Wallaert
A
,
Durinck
K
,
Van Loocke
W
, et al
.
Long noncoding RNA signatures define oncogenic subtypes in T-cell acute lymphoblastic leukemia
.
Leukemia
.
2016
;
30
(
9
):
1927
-
1930
.
72.
Durinck
K
,
Wallaert
A
,
Van de Walle
I
, et al
.
The Notch driven long non-coding RNA repertoire in T-cell acute lymphoblastic leukemia
.
Haematologica
.
2014
;
99
(
12
):
1808
-
1816
.
73.
Trimarchi
T
,
Bilal
E
,
Ntziachristos
P
, et al
.
Genome-wide mapping and characterization of Notch-regulated long noncoding RNAs in acute leukemia
.
Cell
.
2014
;
158
(
3
):
593
-
606
.
74.
Holmfeldt
L
,
Wei
L
,
Diaz-Flores
E
, et al
.
The genomic landscape of hypodiploid acute lymphoblastic leukemia
.
Nat Genet
.
2013
;
45
(
3
):
242
-
252
.
75.
Vogelstein
B
,
Papadopoulos
N
,
Velculescu
VE
,
Zhou
S
,
Diaz
LA
Jr
,
Kinzler
KW
.
Cancer genome landscapes
.
Science
.
2013
;
339
(
6127
):
1546
-
1558
.
76.
Speck
NA
,
Gilliland
DG
.
Core-binding factors in haematopoiesis and leukaemia
.
Nat Rev Cancer
.
2002
;
2
(
7
):
502
-
513
.
77.
Knudson
AG
Jr
.
Mutation and cancer: statistical study of retinoblastoma
.
Proc Natl Acad Sci USA
.
1971
;
68
(
4
):
820
-
823
.
78.
Tosello
V
,
Mansour
MR
,
Barnes
K
, et al
.
WT1 mutations in T-ALL
.
Blood
.
2009
;
114
(
5
):
1038
-
1045
.
79.
Hunger
SP
,
Lu
X
,
Devidas
M
, et al
.
Improved survival for children and adolescents with acute lymphoblastic leukemia between 1990 and 2005: a report from the children’s oncology group
.
J Clin Oncol
.
2012
;
30
(
14
):
1663
-
1669
.
80.
den Hoed
MA
,
Pluijm
SM
,
te Winkel
ML
, et al
.
Aggravated bone density decline following symptomatic osteonecrosis in children with acute lymphoblastic leukemia. Haematologica. 2015;100(12):1564-1570
.
81.
Litzow
MR
,
Ferrando
AA
.
How I treat T-cell acute lymphoblastic leukemia in adults
.
Blood
.
2015
;
126
(
7
):
833
-
841
.
82.
Pui
C-H
,
Mullighan
CG
,
Evans
WE
,
Relling
MV
.
Pediatric acute lymphoblastic leukemia: where are we going and how do we get there?
Blood
.
2012
;
120
(
6
):
1165
-
1174
.
83.
Pui
C-H
,
Robison
LL
,
Look
AT
.
Acute lymphoblastic leukaemia
.
Lancet
.
2008
;
371
(
9617
):
1030
-
1043
.
84.
Nguyen
K
,
Devidas
M
,
Cheng
S-C
, et al
;
Children’s Oncology Group
.
Factors influencing survival after relapse from acute lymphoblastic leukemia: a Children’s Oncology Group study
.
Leukemia
.
2008
;
22
(
12
):
2142
-
2150
.
85.
Bhojwani
D
,
Pui
C-H
.
Relapsed childhood acute lymphoblastic leukaemia
.
Lancet Oncol
.
2013
;
14
(
6
):
e205
-
e217
.
86.
Tzoneva
G
,
Perez-Garcia
A
,
Carpenter
Z
, et al
.
Activating mutations in the NT5C2 nucleotidase gene drive chemotherapy resistance in relapsed ALL
.
Nat Med
.
2013
;
19
(
3
):
368
-
371
.
87.
Kunz
JB
,
Rausch
T
,
Bandapalli
OR
, et al
.
Pediatric T-cell lymphoblastic leukemia evolves into relapse by clonal selection, acquisition of mutations and promoter hypomethylation
.
Haematologica
.
2015
;
100
(
11
):
1442
-
1450
.
88.
Mullighan
CG
,
Phillips
LA
,
Su
X
, et al
.
Genomic analysis of the clonal origins of relapsed acute lymphoblastic leukemia
.
Science
.
2008
;
322
(
5906
):
1377
-
1380
.
89.
Clappier
E
,
Gerby
B
,
Sigaux
F
, et al
.
Clonal selection in xenografted human T cell acute lymphoblastic leukemia recapitulates gain of malignancy at relapse
.
J Exp Med
.
2011
;
208
(
4
):
653
-
661
.
90.
Szczepanski
T
,
van der Velden
VH
,
Waanders
E
, et al
.
Late recurrence of childhood T-cell acute lymphoblastic leukemia frequently represents a second leukemia rather than a relapse: first evidence for genetic predisposition
.
J Clin Oncol
.
2011
;
29
(
12
):
1643
-
1649
.
91.
Brouwer
C
,
Vogels-Mentink
TM
,
Keizer-Garritsen
JJ
, et al
.
Role of 5′-nucleotidase in thiopurine metabolism: enzyme kinetic profile and association with thio-GMP levels in patients with acute lymphoblastic leukemia during 6-mercaptopurine treatment
.
Clin Chim Acta
.
2005
;
361
(
1-2
):
95
-
103
.
92.
Papayannidis
C
,
DeAngelo
DJ
,
Stock
W
, et al
.
A Phase 1 study of the novel gamma-secretase inhibitor PF-03084014 in patients with T-cell acute lymphoblastic leukemia and T-cell lymphoblastic lymphoma
.
Blood Cancer J
.
2015
;
5
(
9
):
e350
.
93.
Zweidler-McKay
P
,
DeAngelo
DJ
,
Douer
D
, et al
.
The safety and activity of BMS-906024, a gamma secretase inhibitor (GSI) with anti-Notch activity, in patients with relapsed T-cell acute lymphoblastic leukemia (T-ALL): initial results of a phase 1 trial [abstract]
.
Blood
.
2014
;
124
(
21
):
968
.
94.
Messersmith
WA
,
Shapiro
GI
,
Cleary
JM
, et al
.
A phase I, dose-finding study in patients with advanced solid malignancies of the oral γ-secretase inhibitor PF-03084014
.
Clin Cancer Res
.
2015
;
21
(
1
):
60
-
67
.
95.
Knoechel
B
,
Bhatt
A
,
Pan
L
, et al
.
Complete hematologic response of early T-cell progenitor acute lymphoblastic leukemia to the γ-secretase inhibitor BMS-906024: genetic and epigenetic findings in an outlier case
.
Cold Spring Harb Mol Case Stud
.
2015
;
1
(
1
):
a000539
.
96.
Aste-Amézaga
M
,
Zhang
N
,
Lineberger
JE
, et al
.
Characterization of Notch1 antibodies that inhibit signaling of both normal and mutated Notch1 receptors
.
PLoS One
.
2010
;
5
(
2
):
e9094
.
97.
Wu
Y
,
Cain-Hom
C
,
Choy
L
, et al
.
Therapeutic antibody targeting of individual Notch receptors
.
Nature
.
2010
;
464
(
7291
):
1052
-
1057
.
98.
Sulis
ML
,
Saftig
P
,
Ferrando
AA
.
Redundancy and specificity of the metalloprotease system mediating oncogenic NOTCH1 activation in T-ALL
.
Leukemia
.
2011
;
25
(
10
):
1564
-
1569
.
99.
Hayashi
I
,
Takatori
S
,
Urano
Y
, et al
.
Neutralization of the γ-secretase activity by monoclonal antibody against extracellular domain of nicastrin
.
Oncogene
.
2012
;
31
(
6
):
787
-
798
.
100.
Moellering
RE
,
Cornejo
M
,
Davis
TN
, et al
.
Direct inhibition of the NOTCH transcription factor complex
.
Nature
.
2009
;
462
(
7270
):
182
-
188
.
101.
Liu
R
,
Li
X
,
Tulpule
A
, et al
.
KSHV-induced notch components render endothelial and mural cell characteristics and cell survival
.
Blood
.
2010
;
115
(
4
):
887
-
895
.
102.
Degryse
S
,
de Bock
CE
,
Cox
L
,
Demeyer
S
,
Gielen
O
.
JAK3 mutants transform hematopoietic cells through JAK1 activation, causing T-cell acute lymphoblastic leukemia in a mouse model
.
Blood
.
2014
;
124(20):3092-3100
.
103.
Maude
SL
,
Dolai
S
,
Delgado-Martin
C
, et al
.
Efficacy of JAK/STAT pathway inhibition in murine xenograft models of early T-cell precursor (ETP) acute lymphoblastic leukemia
.
Blood
.
2015
;
125
(
11
):
1759
-
1767
.
104.
Cramer
SD
,
Aplan
PD
,
Durum
SK
.
Therapeutic targeting of IL-7Rα signaling pathways in ALL treatment
.
Blood
.
2016
;
128
(
4
):
473
-
478
.
105.
Degryse
S
,
Cools
J
.
JAK kinase inhibitors for the treatment of acute lymphoblastic leukemia
.
J Hematol Oncol
.
2015
;
8
:
91
.
106.
Crompton
T
,
Outram
SV
,
Hager-Theodorides
AL
.
Sonic hedgehog signalling in T-cell development and activation
.
Nat Rev Immunol
.
2007
;
7
(
9
):
726
-
735
.
107.
Dagklis
A
,
Pauwels
D
,
Lahortiga
I
, et al
.
Hedgehog pathway mutations in T-cell acute lymphoblastic leukemia
.
Haematologica
.
2015
;
100
(
3
):
e102
-
e105
.
108.
Dagklis
A
,
Demeyer
S
,
De Bie
J
, et al
.
Hedgehog pathway activation in T-cell acute lymphoblastic leukemia predicts response to SMO and GLI1 inhibitors. Blood. 2016;128(23):2642-2654
.
109.
Etchin
J
,
Sanda
T
,
Mansour
MR
, et al
.
KPT-330 inhibitor of CRM1 (XPO1)-mediated nuclear export has selective anti-leukaemic activity in preclinical models of T-cell acute lymphoblastic leukaemia and acute myeloid leukaemia
.
Br J Haematol
.
2013
;
161
(
1
):
117
-
127
.
110.
Trinquand
A
,
Dos Santos
NR
,
Tran Quang
C
, et al
.
Triggering the TCR developmental checkpoint activates a therapeutically targetable tumor suppressive pathway in T-cell leukemia
.
Cancer Discov
.
2016
;
6
(
9
):
972
-
985
.
111.
Hawkins
ED
,
Duarte
D
,
Akinduro
O
, et al
.
T-cell acute leukaemia exhibits dynamic interactions with bone marrow microenvironments
.
Nature
.
2016
;
538
(
7626
):
518
-
522
.
112.
Passaro
D
,
Irigoyen
M
,
Catherinet
C
, et al
.
CXCR4 is required for leukemia-initiating cell activity in T cell acute lymphoblastic leukemia
.
Cancer Cell
.
2015
;
27
(
6
):
769
-
779
.
113.
Pitt
LA
,
Tikhonova
AN
,
Hu
H
, et al
.
CXCL12-producing vascular endothelial niches control acute T cell leukemia maintenance
.
Cancer Cell
.
2015
;
27
(
6
):
755
-
768
.
Sign in via your Institution