Abstract
Acute myelogenous leukemia (AML) can progress quickly and without treatment can become fatal in a short period of time. However, over the last 30 years fine-tuning of therapeutics have increased the rates of remission and cure. Cytogenetics and mutational gene profiling, combined with the option of allogeneic hematopoietic stem cell transplantation offered in selected patients have further optimized AML treatment on a risk stratification basis in younger adults. However there is still an unmet medical need for effective therapies in AML since disease relapses in almost half of adult patients becoming refractory to salvage therapy. Improvements in the understanding of molecular biology of cancer and identification of recurrent mutations in AML provide opportunities to develop targeted therapies and improve the clinical outcome. In the spectrum of identified gene mutations, primarily targetable lesions are gain of function mutations of tyrosine kinases FLT3, JAK2 and cKIT for which specific, dual and multi-targeted small molecule inhibitors have been developed. A number of targeted compounds such as sorafenib, quizartinib, lestaurtinib, midostaurin, pacritinib, PLX3397 and CCT137690 are in clinical development. For loss-of-function gene mutations, which are mostly biomarkers of favorable prognosis, combined therapeutic approaches can maximize the therapeutic efficacy of conventional therapy. Apart from mutated gene products, proteins aberrantly overexpressed in AML appear to be clinically significant therapeutic targets. Such a molecule for which targeted inhibitors are currently in clinical development is PLK1. We review characteristic gene mutations, discuss their biological functions and clinical significance and present small molecule compounds in clinical development, which are expected to have a role in treating AML subtypes with characteristic molecular alterations.
Keywords: Acute myeloid leukemia, targeted therapy, mutation, FLT3, NPM1, CEBPA, JAK2
Introduction
Acute myelogenous leukemia (AML) can progress quickly and without treatment can become fatal in a short period of time. However, over the last 30 years fine-tuning of therapies and therapeutic schemes have increased the rates of remission and cure [1]. Currently certain karyotype abnormalities and gene mutations are being taken into consideration to guide treatment and in particular the therapeutic use of allogeneic hematopoietic stem cell transplantation in non-elderly patients. However, AML remains incurable for a significant proportion of adult patients [2-4], while no viable therapeutic option exists for patients with relapsed and refractory AML [5]. In this context it is crucial to develop novel targeted therapies that could improve the clinical outcome in subsets of AML [6].
A better understanding of the molecular basis of cancer during the last two decades has contributed to the development of drugs that target protein products of mutated or chimeric genes, which are linked to various cancers [7-9]. Pivotal example of small-molecule kinase inhibitors that bind to driver oncoprotein and block its function on a potentially curative intent are the BCR-ABL kinase inhibitors, in use for chronic myeloid leukemia (CML) [10]. Acute Promyelocytic Leukemia (APL) is another example where effective targeted therapies, such as all-trans retinoic acid (ATRA) and arsenic trioxide are used and can reinstall differentiation of leukemic promyelocytes by targeting the culprit PML-RARa fusion protein [11,12]. However, in non-APL AML, despite the identification of a handful recurrent chimeric genes and gene mutations, development of targeted therapies has been notably sluggish. Here we review characteristic gene mutations, discuss their biological functions and clinical significance and present small molecule compounds in clinical development, which are expected to have a role in treating non-APL AML subtypes with characteristic druggable mutations.
Gene mutations of good prognosis
NPM1 mutations in AML (NPMc+ AML)
Nucleophosmin (NPM1) is a ubiquitously expressed phosphoprotein that belongs to the nucleoplasmin family of nuclear chaperones. It is encoded by the NPM1 gene located at 5q35.1 that produces 3 isoforms through alternative splicing, of which NPM1 (or nucleolar phosphoprotein B23.1) is a 294-amino acid [13,14]. NPM1 is a pleiotropic nucleolar protein that shuttles across cytoplasm and nucleoplasm and regulates among others centrosome maturation and the ARF/p53 pathway [15-18].
NPM1 protein as a product of mutated or fused NPM1 has been associated with certain blood cancers. In Anaplastic Large Cell Lymphoma (ALCL) the t(2;5) translocation results to NPM1-ALK fusion oncoprotein, which anchors to cell cytoplasm and is detected with antibodies against the NPM1 N-terminus [19]. In 2005, it was first discovered that about 35% of adult AML had aberrant nucleophosmin expression in leukemic cell cytoplasm, as a result of NPM1 gene insertions at exon-12 [20]. In about 80% of AML with mutated NPM1, the mutation is a duplication of the 4-base sequence TCTG at positions 956-959 of the NPM1 gene, the socalled mutation A, which results in a slightly longer protein with a different C-terminal amino-acidic sequence [21]. These C-terminal changes are responsible for cytoplasmic localization of the NPM1 leukemic mutants through generation of new nuclear export signal (NES) motifs and loss of the two tryptophan residues 288 and 290 which cause the unfolding of the C-terminal domain and thus loss of binding capacity to the nucleolus [21,22]. NPM1 mutations may rarely occur at exon-9 and exon-11 and these mutants also localize in the cytoplasm through the same mechanism that exon-12 NPM1 mutants dislocate [23,24].
Cytoplasmic mutant NPM1 contributes to AML development by inactivating p19Arf through delocalization of the tumor suppressor protein. This results in reduced p19Arf activities, both p53-dependent (Mdm2 and p21cip1 induction) and p53-independent (sumoylation of NPM). p19Arf stability is compromised when coupled with NPM1 mutant, which may lead to weaker control of the p53-dependent cell-cycle arrest [25,26]. Mutated NPM1 bounds to NF-kappaB and dislocates it in the cytoplasm, leading to its inactivation. This inactivation of NF-kappaB is thought to be responsible for the high response rates of AML with NPM1mutant to chemotherapy [27,28] NPM1c+ (cytoplasmic positive) AML is closely associated with normal karyotype and represent a provisional entity in the WHO 2008 classification.
NPM1 targeted therapy
There are two key points that prompt consideration of nucleophosmin as a therapeutic target: a) NPM1 mutation is one of the most common recurring genetic lesions in AML with a prevalence of 27%-35% in adult AML and 45%-64% in adult AML with a normal karyotype and b) normal karyotype AML and the genotype ‘mutant NPM1 without FLT3-ITD’ carry a most favorable prognosis when treated with intensive chemotherapy [29-32]. This data indicate that NPM1 mutation behaves as a founder genetic lesion in a fraction of AML patients, which makes it an attractive target for therapeutic intervention, primary aiming to increase chemotherapy efficacy [33]. Interestingly it has been shown that the favorable outcome of chemotherapy in NPM1 mutated non FLT3-ITD AML can be improved by incorporating all-trans retinoic acid (ATRA) [34]. Moreover specific inhibitors of NPM1 oligomerization such as NSC348884 may further sensitize leukemic cells of this genotype to apoptosis when exposed to the ATRA plus cytarabine combination [35].
CEBPA mutations in AML
CCAAT/enhancer binding protein alpha (CEBP-alpha, CEBPA) is an intronless gene located at chromosome 19q13.1 that encodes for a basic region leucine zipper (bZIP) transcription factor, which can bind as a homodimer to certain promoters and enhancers but can also form heterodimers with the related proteins CEBP-beta and CEBP-gamma [36]. CEBPA functions as key regulator of granulocytic differentiation [37].
CEBPA mutations contribute to leukemogenesis by promoting proliferation and blocking differentiation of myeloid lineage [38,39]. The two most frequent mutations are: a) N-terminal frame-shift mutations that truncate the p42 form while preserving the p30 form which inhibits the remaining wild-type CEBPA p42 protein in a dominant-negative manner and b) C-terminal in-frame insertions or deletions that disrupt the basic zipper region, thus affecting DNA binding [40]. Most cases carry both types of CEBPA mutations: a N-terminal frame-shift mutation and a C-terminal in-frame mutation, with the two mutations typically being located on different alleles [41,42].
CEBPA-mutated AML usually displays classical features of AML with or without cell maturation but some cases may show monocytic or monoblastic features. Myeloid-associated antigens HLA-DR and CD34 are usually expressed, as is CD7 in a significant proportion of patients. About 70% of cases have normal karyotype and approximately 25% carry concomitant FLT3-ITD mutations [43].
Prognosis of cytogenetically normal (CN)-AML patients with CEBPA mutations in the absence of an FLT3-ITD or NPM1 mutation, is favorable, similar to AML with inv(16)(p13.1q22) or t(8; 21)(q22; q22) [43-45]. However only patients with double CEBPA mutations have favorable clinical course, whereas single CEBPA mutations not only do they not differ from CEBPA wild-type patients but also they have a tendency toward high-risk FLT3-ITD mutations [46]. However, coexistence of NPM1 mutations with monoallelic CEBPA mutations was shown to be associated with prolonged survival in CN-AML patients [47]. Hereditary predisposition is a noteworthy point related to CEBPA. Germ-cell mutations appear to occur in 7% of patients with CN AML and myeloid precursor cells from healthy individuals carrying single germ-line CEBPA mutation may evolve to overt AML by acquiring a second sporadic CEBPA mutation [46,48]. Adult AML with CEBPA mutation is also a provisional entity in the current WHO classification.
CEBPA targeted therapy
Restoring function of particular dysregulated transcription factors appears to be a reasonable target for novel therapeutic strategies in AML [49]. However, no therapies to restore CEBPA function in dysregulated CEBPA CN-AML cases have been currently developed.
Gene mutations of poor prognosis
FLT3 mutations in AML
FLT3 (Fms-like tyrosine kinase 3, CD135) is a member of class III tyrosine kinase (RTKIII) receptor family, which also includes c-FMS, c-KIT, and PDGFR. The FLT3 gene encodes a 993-amino acid protein in humans, which is composed of an immunoglobulin-like extracellular ligand-binding domain, a transmembrane domain, a Juxtamembrane dimerization domain and a cytoplasmic domain with a split tyrosine kinase motif. It is expressed in immature hematopoietic cells, placenta, gonads, brain, and in lymphohematopoietic organs such as the liver, spleen and the thymus [50].
FLT3 receptor exists in a monomeric unphosphorylated status and turns activated when bound by its FLT3 ligand (FL), which promotes its unfolding and homodimerization. Homodimerization of FLT3 switches on its tyrosine kinase activity and recruits a number of intracellular proteins [SHC proteins, GRB2, GRB2-associated binder 2 (GAB2), SHIP, CBL, CBLB (CBLB-related protein)] to its intracellular domain. Each protein becomes activated and a phosphorylation cascade starts resulting in activation of secondary mediators (MAP kinase, STAT, and AKT/PI3 kinase signal transduction pathways), which are transported to the nucleus by HSP90, where they regulate transcription of several genes, which participate in differentiation, proliferation, and apoptosis [51,52].
FLT3 expression in the normal bone marrow is restricted to early progenitors, including CD34+ cells with high levels of expression of CD117 (c-KIT), and committed myeloid and lymphoid progenitors with variable expression in the more mature monocytic lineage [53]. It is also expressed at high levels in many hematologic malignancies including most of AML subtypes, B-precursor cell acute lymphoblastic leukemia (ALL), some T-cell ALLs, and CML in blast crisis [54,55].
Mutations of the FLT3 are of major clinical relevance in AML because they commonly guide treatment decisions as independent indicators of poor prognosis [2,3].
FLT3 internal tandem duplications (FLT3-ITD)
The most common mutation of FLT3 in AML is internal tandem duplications (FLT3-ITD). FLT3-ITD results from a duplication of a fragment within the juxtamembrane domain coding region (encoded by exons 14 and 15) of FLT3. It is one of the most common mutations in hematologic malignancies, occurring in CML (5-10%), MDS (5-10%), and AML (15-35%) patients [56]. Nakao et al. first described FLT3-ITD in a high proportion of patients with AML [56]. FLT3-ITD is rare in infant AML, but increases to 5% to 10% in age 5 to 10 years, 20% in young adults, and >35% in AML patients older than 55 years [57]. FLT3-ITD mutations vary in size and region of ITD involvement that ranges from 3 to more than 400 base pairs [58].
Segmental duplication of the Juxtamembrane (JM) domain of FLT3 promotes auto-dimerization and autophosphorylation of the receptor, which turns it constitutively phosphorylated and activating AKT [59,60]. Some of the effects of FLT3-ITDs are unique to the mutated receptor: cellular proliferation of FLT3-ITD transduced cells is mediated by RAS and STAT5 pathways, while ligand-induced FLT3-WT activation does not lead to STAT5 activation and no STAT5 DNA binding [61].
FLT3 tyrosine kinase domain mutations (FLT3/TKD)
Missense mutations have also been described in the activation loop domain of the tyrosine kinase of FLT3 (FLT3 activation loop mutation, FLT3/ALM, or FLT3 Tyrosine Kinase Domain mutation, FLT3/TKD) [62]. FLT3/TKD are the second most common type of FLT3 mutations found in 5-10% of AML and they can rarely coexist with FLT3-ITD. The majority of the FLT3/TKD occur in codon 835 with a change of an aspartic acid to tyrosine (D835Y or Asp835Tyr), however, other point mutations, deletions, and insertions within codon D835 (Asp835) and its surrounding codons have been described [60,62-64] FLT3/TKD promotes ligand-independent proliferation through autophosphorylation and constitutive receptor activation, similar to that of FLT3-ITD but there are significant biological differences between the two types of FLT3 mutations. They promote activation of different downstream effectors, and trigger different biological responses [65,66].
Prognostic significance of FLT3-ITD
Many large studies have shown that presence of FLT3-ITD is an independent prognostic factor for poor outcome in AML [63,67]. Kottaridis et al [68] examined the prevalence and prognostic significance of FLT3-ITD in a cohort of more than 850 adult AML patients. They found FLT3-ITD in 27% of patients and confirmed previous studies showing that FLT3-ITDs were associated with leukocytosis and normal cytogenetics. In their study, AML patients with FLT3-ITD had a lower remission rate, higher relapse rate (RR), and worse survival. Multivariate analyses showed that FLT3-ITD was the most significant prognostic factor with respect to RR and disease free survival (DFS) [68]. In other studies, survival for patients with FLT3-ITD was 20% to 30% compared to 50% for those without FLT3-ITD and allelic variation (mutant to wild-type ratio) in patients with FLT3-ITD seemed to influence outcome. Various thresholds of FLT3/ITD allelic ratio established an allelic ratio threshold that demarcated patients with FLT3-ITD at high risk of relapse [57]. Similar work in other studies has shown differences in clinical outcome for those with differing allelic ratios [69].
FLT3 targeted therapies
FLT3 tyrosine kinase is thought to be the most reasonable targetable protein in AML [70,71]. Several potent FLT3 kinase inhibitors are currently in development for AML that harbors FLT3-ITD mutations and first results of FLT3 inhibitors in clinical development have already produced encouraging and clinical relevant activity (Table 1) [71-73].
Table 1.
Mutated Gene | Gene pruduct function | Compound | Targets | trial phases | combinations | Trials Registered Clinicaltrials.gov (Nov 2012) | Refs |
---|---|---|---|---|---|---|---|
FLT3 | This gene encodes a tyrosine-protein kinase that acts as cell-surface receptor for the cytokine (fms-related tyrosine kinase 3 ligand) FLT3LG and regulates differentiation, proliferation and survival of hematopoietic progenitor cells and of dendritic cells. Activation of wild-type FLT3 causes only marginal activation of proliferation, and differentiation of hematopoietic cells in bone marrow. Mutations that cause constitutive kinase activity promote cell proliferation and resistance to apoptosis via the activation of multiple signaling pathways | Lestaurtinib | FLT3, JAK2 | 1,2,3 | chemotherapy | 4 active 4 completed | [89,211] |
PLX3397 | FMS, KIT & FLT3 | 1 | none | 1 active | [96] | ||
Quizartinib (AC220) | FLT3, KIT, PDGFRA, PDGFRB, RET | 1,2 | chemotherapy | 6 active 1 completed | [86] | ||
Sorafenib | FLT3, cKIT | 1, 2 | 11 active 6 completed | [212-215] | |||
Midostaurin | Multiple kinases | 1,2,3 | ATRA, Bortezomib, azacytidine, decitabine, chemotherapy | 9 active 4 completed | [90] | ||
NPM1 | A nucleolar phosphoprotein that shuttles across cytoplasm and nucleus and regulates centrosome maturation and the ARF/p53 pathway. NPM1 leukemic mutants are characterized by insertions that cause a reading frame-shift and result in a longer protein with a different C-terminal, which is responsible for lossing the binding capacity to the nucleolus. This may contribute to leukemogenesis by inactivating p19Arf through delocalization of the tumor suppressor protein and dislocating NF-kappaB to the cytoplasm. | ATRA | Induces p53 and p21 | 1 | 2 active | [34] | |
PLK1 | PLK1 is the most well characterized member of four serine/threonine protein kinases, which and strongly promotes the progression of cells through mitosis. PLK1 performs several important functions throughout M phase of the cell cycle, including the regulation of centrosome maturation and spindle assembly, the removal of cohesins from chromosome arms, the inactivation of anaphase-promoting complex/cyclosome (APC/C) inhibitors, and the regulation of mitotic exit and cytokinesis. It plays a key role in centrosome functions and the assembly of bipolar spindles. It als acts as a negative regulator of p53 family members leading to ubiquitination and subsequent degradation of p53/TP53, inhibition of the p73/TP73 mediated pro-apoptotic functions and phosphorylation/degradation of BORA. During the various stages of mitosis PLK1 localizes to the centrosomes, kinetochores and central spindle. Plk1 is found aberrantly overexpressed in a variety of human cancers and in AML, correlated with cellular proliferation and poor prognosis. | Volasertib | PLK1 | 1,2,3 | chemotherapy | 2 active | [216] |
BI 2536 | PLK1 | 1,2 | chemotherapy | 1 completed | [207] | ||
MLL | The MLL gene (mixed-lineage leukemia) encodes a protein plays an essential role in early development and hematopoiesis by acting as a histone methyltransferase and transcriptional co-activator. Among others it activates aberrant trascription DOT1L, which is considered a driver of leukemogenesis | EPZ-5676 | DOT1L | 1 | 1 active | [175] | |
JAK2 | The Janus-kinase-2 gene (JAK2) encodes a non-receptor tyrosine kinase involved in relaying signals for hemopoietic cell growth, development and differentiation crossactivated by type I/II cytokine receptors | Ruxolitinib | JAK2 | 1,2 | 1 active | [217] | |
Pacritinib | JAK2, FLT3 | 1 | 1 completed | [82] | |||
Lestaurtinib (CEP-701) | JAK2, FLT3 | 1,2 | 1 active 3 completed | [218] |
Sorafenib is one of the most extensively investigated first generation FLT3 inhibitors. It has shown to specifically reduce the percentage of leukemia blasts in the peripheral blood (7.5% from 81%) and the bone marrow (34% from 75.5%) of AML patients with FLT3-ITD but not in patients without this mutation [74]. It has also shown activity in FLT3-ITD-positive AML relapsing patients after allogeneic stem cell transplantation [75]. However, development of resistance to TKIs is a well known therapeutic problem [76,77] and in the case of FLT3-ITD+ AML, it appears that stromal niche cells offer sanctuary to early leukemic stem/progenitor cells protecting them from eradication by firstgeneration inhibitors [78]. Several investigators focus their attempts in developing strategies to prevent or reverse ‘acquired’ resistance to TKIs. Recent in vitro studies have shown that the anti-leukemic activity of TKIs can be increased when combined with the proapoptotic small molecule Nutlin-3, which inhibits the MDM2/p53 interaction [79,80]. Moreover fluvastatin, a drug in use for the treatment of hypercholesterolemia, has shown potency to reverse resistance and increase activity of sorafenib [81,82].
Upregulation of JAK2 in FLT3-TKI-resistant AML cells appears to be a potential mechanism of resistance to selective FLT3 inhibition [83,84]. Second-generation potent multi-targeted FLT3/JAK2 inhibitors are thought to address this important therapeutic issue. A number of such compounds, such as quizartinib, lestaurtinib and midostaurin are currently in early phases of clinical development.
Quizartinib (AC220) is such a second-generation FLT3 inhibitor, which exhibits low nanomolar potency, good bioavailability and exceptional kinase selectivity [85]. Early clinical results of quizartinib were promising. They showed meaningful reductions in marrow blasts in a substantial proportion of patients with both refractory and relapsed FLT3-ITD+ AML [86]. Lestaurtinib (CEP701) a dual FLT3 and JAK2 inhibitor has shown activity as monotherapy in AML, but although it produced high remission rates, it failed to increase survival in combination with cytarabine and idarubicin in young patients with relapsed or refractory AML [87-89]. Midostaurin (PKC412), a semi-synthetic multitargeted tyrosine kinase inhibitor, has demonstrated activity as monotherapy in patients with FLT3-mutant and wild-type AML and high complete response and survival rates when given in combination with standard chemotherapy in newly diagnosed young adults with AML [90,91]. Pacritinib (SB1518) is another novel potent JAK2/FLT3 inhibitor, which has demonstrated promising activity and clinical benefits in refractory AML patents treated in a phase I trial [82]. Pacritinib in combination with pracinostat (SB939), an oral HDAC inhibitor showed synergy in reduction of tumor growth and JAK2 and FLT3 signaling [92]. Another oral multikinase inhibitor that has showed antileukemic activity in preclinical models is TG02 that inhibits CDKs 1, 2, 7 and 9 along with JAK2 and FLT3 [93].
In addition more specific and potent anti-FLT3 compounds such as PLX3397 and FLT3-Aurora kinase inhibitor CCT137690, are in early phases of clinical development [94-96] and others such as DCC2036, CCT241736 have produced in vitro very promising data for the treatment of FLT3-ITD+ AML [97].
KIT mutations in core binding factor leukemia (CBF) AML
Core-binding factor (CBF) AML patients when compared to other cytogenetic groups have a favorable prognosis, particularly when treated with high-dose cyrabine consolidation regimens and do not require stem cell transplantation. However, relapses do occur and approximately 50% of patients with these cytogenetic abnormalities are alive at 5 years [98]. Mutations that have been found in this group of patients and have been correlated with adverse outcome are related to cKIT and JAK2 genes [99-101].
The c-KIT gene (stem cell factor) encodes for a tyrosine kinase with a structure similar to platelet growth factor and is expressed in hematopoietic progenitor cells and AML blasts [102]. Upon binding of the ligand stem cell factor to c-kit, phosphorylation of several cytoplasmic proteins occurs and pertinent downstream pathways get activated. Those include the JAK/STAT pathway, the PI-3 kinase pathway and the MAP kinase pathway [103]. Mutations in c-KIT receptor result in constitutive phosphorylation and activation of the receptor in absence of the ligand. Mutations in the KIT and FLT3 genes are associated with unfavourable prognosis in AML patients with t(8; 21). In particular, patients with c-KIT mutated have been reported to have a higher incidence of relapse (80 versus 13.5%) and a lower 6-year progression free survival (PFS) compared to unmutated [104]. Patients with t(8; 21), but not those with inv(16) have a shorter relapse free survival when harbouring a mutated c-kit [105]. KIT mutation was recently found to be related to PFS in patients with inv(16) or t(16; 16) AML [106]. The prognostic impact of c-KIT mutations in patients with inv(16) remains controversial since some studies failed to establish a link [105], while others found that exon 8 mutations increased the relapse rate but did not affect overall survival (OS) [107].
KIT targeted therapies
Multi-kinase inhibitors imatinib and sunitinib beside their indications for the treatment of CML and renal cancer respectively, have also been licensed for the treatment of gastrointestinal stromal tumors, because they effectively inhibit mutated c-KIT, which is the characteristic molecular abnormality in these tumors [108,109]. However, not all c-KIT mutations respond to the same agent. Exon 8 and the exon 17 N822 c-KIT mutations but not the D816 are sensitive to imatinib in vitro, therefore assessment of the exact c-KIT mutational status is important and may have direct therapeutic consequences. Initial clinical studies with imatinib in a small number of patients with refractory AML did not show beneficial results [110], however, when tested in c-KIT positive AML patients results were more promising [111]. Several studies have investigated the activity of imatinib alone or in combination with chemotherapy in c-KIT positive AML patients and results are awaited.
Small molecules such as SU5416 and SU6668 have activity against c-KIT [109] although neither is selective. Both were developed as angiogenesis inhibitors and also inhibit FLT3, KDR and FGFR [109]. In addition SU5416 inhibits VEGFR2, while SU6668 inhibits PDGFR. SU6668 has shown antiangiogenic properties and inhibition of c-KIT in preclinical models, whereas SU5416 reached later stages of drug development; however it showed modest activity in patients with relapsed/refractory AML or MDS [112]. Further investigating these molecules has been halted. APcK110 is a novel KIT inhibitor with potent proapoptotic and antiproliferative activity in AML cell lines and primary samples whereas in an AML xenograft mouse model it was shown to extend survival [113].
JAK2 mutations in CBF AML
The Janus-kinase-2 gene (JAK2) encodes a non-receptor tyrosine kinase involved in relaying signals for hemopoietic cell growth, development and differentiation [114]. JAK proteins consist a family of four non-receptor tyrosine kinases (JAK1, JAK2, JAK3 and Tyk2) that are closely associated with type I/II cytokine receptors. When activated via association to cell surface receptors they phosphorylate and translocate STATs to the nucleus to activate gene transcription [115,116]. Among the family members JAK2 associates with the IFN-1, IL-6, 12/23 cytokine and EPO receptors [116].
JAK2 is commonly mutated in myeloid neoplasias. The JAK2V617F gain of function mutation in the cytoplasmic tyrosine kinase domain is a common finding in myeloproliferative neoplasms [117]. The same mutation has been found in a small number of AML patients, more commonly in t(8; 21) AML [100,101,118]. AML t(8; 21) patients harbouring JAK2V617F in addition to KIT and FLT3 mutations have poorer disease-free survival compared to wild type JAK2 [119-121]. Moreover activating JAK2 gene fusions with the TEL(ETV6) (TEL-JAK2) and PCM1 genes have been found in leukemia patients [122-124]. Beside the detected mutations, a recent immunohistochemical study found JAK2 to be invariably activated (phosphorylated) in AML, while high of p-JAK2 levels were found to be a predictor of poor response to chemotherapy (45% in patients with high p-JAK2 vs. 78% in patients with low p-JAK2, p < 0.003) and a factor of poor prognosis (p=0.023) which justifies its consideration as a therapeutic target in AML [125].
JAK2 targeted therapy in AML
JAK inhibitors constitute a new class of drugs with activity in a wide range of diseases, primarily in myeloproliferative neoplasias and autoimmune disorders [126]. Ruxolitinib, the first JAK inhibitor that recently received marketing authorization by FDA and EMA for the treatment of myelofibrosis, is now investigated in patients with relapsed or refractory acute leukemia (ClinicalTrials.gov Identifier: NCT01251965) [127].
Following ruxolitinib approval, several highly potent next generation JAK2/FLT3 inhibitors, such as pacritinib and lestaurtinib, entered clinical evaluation for patients with advanced myeloid malignancies (NCT00719836, NCT00469859) [126]. First available data suggest that blockade of JAK2 in conjunction with FLT3 can enhance clinical benefit for AML patients harboring a FLT3-ITD mutation and provide a strong basis for a clinical evaluation of these targeted small molecule therapeutics in AML patients particularly to those who are resistant to FLT3 directed TKI therapy [82].
Gene mutations of unclear prognostic value
RAS mutations in AML
RAS proto-oncogene belongs to the GTPase family and has 3 isoforms: N-Ras, K-Ras, and H-Ras. Mutant RAS isoforms are found in various types of tumors and leukemia [128]. Point mutations are mostly found at codons 12, 13, and 61 of RAS proto-oncogene. De novo AML patients harbour activating mutations in the RAS proto-oncogenes (N-RAS and K-RAS) in about 25% of cases [129]. HRAS mutations are extremely rare in myeloid leukemia [130]. RAS mutations seem to contribute to leukemogenesis (class I mutations). Several reports have suggested that AML patients harboring RAS mutations have worse, similar or more favourable clinical outcomes than those with wildtype RAS genes [129,131,132]. The presence of RAS mutations seems to sensitize AML cells to high-dose cytarabine therapy in vivo and these patients when treated with chemotherapy alone probably benefit from high-dose cytarabine postremission treatment [133]. NRAS mutations are frequently detected in patients with inv(16)/t(16; 16) [133,134].
RAS targeted therapies
The product of mutated RAS gene is an abnormal Ras protein that is constitutively active. Activated Ras anchores on the cell membrane and stimulates a critical network of signal transduction pathways involved in cellular proliferation, survival and differentiation. Wild type Ras proteins require post-translational modifications by farnesyltransferase (FTase) to get attached to binding sites in the cell membrane to become biologically active. Farnesyl transferase inhibitors (FTIs) are the best-studied class of Ras inhibitors in hematologic malignancies. However Ras can escape FTI suppression and become activated through geranylgeranylation [98]. Tipifarnib), is the main FTI tested in patients with AML. However, inceased toxicity and suboptimal activity in elderly patients did not justify further investigation of this drug [135-137]. The same drug was also proven inactive in young AML patients [138]. Negative was also a phase 2 trial of lonafarnib, which is another FTI in patients with MDS or secondary AML [139].
Gene mutations in epigenetic modifiers
IDH mutations in AML
Isocitrate dehydrogynase (IDH) isoenzymes catalyse an essential step in the Krebs cycle that catalyzes conversion of isocitrate to α-ketoglutarate [140]. In mammalian cells three classes of IDH exist: nicotinamide adenine dinucleotide (NAD)-dependent IDH, mitochondrial nicotinamide adenine dinucleotide (NADP)-dependent IDH, and cytosolic NADP-dependent IDH [141]. IDH1 gene is located at chromosome band 2q33.3 and its product is NADP-dependent and localized in cytoplasm and peroxisomes while IDH2 gene is located at chromosome band 15q26.1 and encodes the mitochondrial NADP-dependent IDH2 enzyme [142,143].
Recurring mutations in IDH1 and IDH2 have been described in more than 70% of World Health Organization grade 2 and 3 astrocytomas, oligodendrogliomas, and glioblastomas [144-146] and in approximately 30% of patients with normal karyotype AML [147-150]. Mutations in the IDH1 occur at R132 while at IDH2 at R172. The mechanisms underlying causal association of mutated IDH with cancer pinpoint to deleterious metabolic alterations although intervention with epigenetic homeostasis through remodeling of the methylome has also been suggested [151-154]. Dang et al showed that IDH1 mutation leads to the overproduction of 2-hydroxyglutarate, a putative oncometabolite that has been associated with a high risk of brain tumors in patients with inborn errors [147,155] and Zhao et al found that mutant IDH1 contributes to tumor growth by activating hypoxia-inducible factor-1α [156].
Somatic mutation at IDH1 R132 was originally described by Mardis et al., who sequenced the entire genome of a CN AML. They subsequently screened 187 AML cases and showed a heterozygous IDH1R132 mutation in 15 cases (8.0%) [147]. In AML, mutant IDH enzyme activity converts α-ketoglutarate to 2-hydroxyglutarate (2-HG), which leads in accumulation of the cancer-associated metabolite 2-hHG [153,157,158]. Several studies have studied the IDH1/2 mutational status in patients with AML and a statistically significant co-occurrence with NPM1 mutations has been reported [159,160]. In the two largest studies correlations IDH1/2 mutations with outcome in AML by the UK MRC, except for the IDH1/2 mutation enrichment in the NPM mutant group, it was reported that patients with the IDH R140Q mutation had an improved OS and decreased response rates. In contrast, IDHR172 mutations did not correlate to outcome or response to therapy, whereas presence of the IDHR132 mutation had an impact on worsened outcome in patients with the FLT3 wild type genotype [160,161]. It becomes obvious that since the number of cooccurring mutations increases, further investigation is needed to better define the prognostic impact of the IDH1/2 mutations in patients with AML.
IDH targeted therapy
It is thought that smallmolecule inhibitors with a potential to block the synthesis of 2-HG could be developed given that IDH mutations lead to a gain-of-function mutation but to date no such therapies have been discovered [162]. However it has been observed that IDH-mutant AMLs have a unique methylation profile characterized by global promoter hypermethylation, which renders these cases reasonable candidates for demethylation therapies [163].
MLL mutations in AML
The MLL gene (mixed-lineage leukemia) encodes a protein that plays an essential role in early development and hematopoiesis by acting as a histone methyltransferase and transcriptional co-activator. One of its domains, the SET domain, mediates methylation of ‘Lys-4’ of histone H3 (H3K4me) complex and acetylation of ‘Lys-16’ of histone H4 (H4K16ac). H3K4me mediates epigenetic transcriptional activation of specific target genes, including many of the HOX genes [164,165].
Aberrant expression of MLL is usually associated with leukemogenesis [166,167]. Mutations and chromosomal translocations involving the MLL gene identify a unique group of acute leukemias, and often predict a poor prognosis. Partial tandem duplication of the MLL gene (MLL-PTD) was the first mutation observed in de novo AML with a normal karyotype or trisomy [168]. These duplications consist of an in-frame duplication of MLL exons. MLL-PTDs are named according to the fused exons e.g. e9/e3. Some, PTD seem to be generated by mispairing of Alu elements, which are repetitive regions with high homology [169].
The incidence of MLL-PTD is around 6% in unselected AML cases but it is higher in cases with normal karyotype (up to 8%) and even higher in cases with trisomy 11 (up to 25%), while favorable karyotypes (e.g. t(8; 21), t(15; 17), inv(16)/t(16; 16)) are MLL-PTD negative [170,171]. MLL-PTD may also be associated with FLT3-ITD and FLT3 point mutations [171,172]. Patients with MLL-PTD expression seem to have shortened remission duration and shorter disease-free survival (DFS) [168].
MLL targeted therapy
MLL specific therapies are optimally targeting mislocated enzymatic activity of DOT1L, which is considered a driver of leukemogenesis in aberrantly expressing MLL leukemias. DOT1L is a histone methyltransferase recruited by rearranged/mutated MLL that methylates lysine-79 of histone H3 and drives expression of the leukemia-causing genes HOXA9 and MEIS1 [93,173,174] The first small molecule inhibitor of DOT1L that entered human clinical development just recently is EPZ-5676 [175].
EZH2 mutations
EZH2 is the enzymatic component of the the Polycomb repressive complex (PRC) components and is an Histone 3 Lysine 27 (H3K27) methyltransferase. Overexpression of EZH2 has been reported in both solid tumors and blood cancers [176,177] and has been shown to be due to, at least in part, the loss of transcriptional repression of specific microRNAs [177]. Missense, nonsense and frameshift mutations have been reported mainly in MDS [178,179], while recently it was shown that almost half cases of early T-cell precursor acute lymphoblastic leukemia present mutations in histone-modifying genes, including EZH2 [180]. In AML, EZH2 mutations have been described in a single case of acute myelomonocytic leukemia out of 143 cases screened [181], in a case with childhood AML [182] and recently in a male with CN-AML out of 50 screened [183]. The contradictory findings of overexpression of EZH2 in epithelial cancers and lymphomas and inactivating mutations in myeloid malignancies raises the possibilty that alterations affecting the methylation of H3K27 may be tumor specific. The effects of EZH2 mutations are still unknown and have only recently started to be under investigation. Initial findings though suggest that except for histone modifications, DNA methylation might also be affected, since EZH2 serves as a recrutiment platform for DNA methyltransferases and seems to be a prerequisite for DNA promoter methylation [184].
EZH2 targeted therapy
Development of selective inhibitors of histone methyltransferases, such as EZH2 have only recently begun. An S-adenosylhomocysteine hydrolase inhibitor named 3-Deazaneplanocin A (DZNep) has been shown to induce efficient apoptotic cell death in cancer cells and not in normal cells and to effectively deplete cellular levels of PRC2 components such as EZH2 while inhibiting associated histone H3K27 methylation [185,186]. Combined DZNep and panobinostat treatment induced more depletion of EZH2 and more apoptosis in AML cells compared to normal CD34(+) bone marrow progenitor cells [187]. This compound has not reached yet the clinical trial setting.
Mutations lacking targeted therapy
It should be noted that not all known recurrently mutated in AML genes have been considered as possible targets for developing novel targeted therapeutics. A number of clinically relevant AML related mutations such of TET2, ASXL1, WT1, p53 and BCOR, although of prognostic significance, are currently lacking known drug discovery activities [183,188-190].
Other novel targeted therapies of interest
PLK1 aberrations in AML
Polo like kinases (PLK) are a family of four serine/threonine protein kinases that are critical regulators of cell cycle progression, mitosis, cytokinesis, DNA damage response and apoptosis [191]. They bind and phosphorylate proteins are that already phosphorylated on a specific motif recognized by the POLO box domains and interplay with Aurora kinases [192,193].
PLK1 is the most well characterized member of PLK1 family and considered to be a master player of cell-cycle regulation during mitosis strongly promoting the progression of cells through mitosis. Characteristically, PLK1 regulates the mitotic licensing of centriole duplication in human cells and also DNA replication under stressful conditions, and anti-apoptotic activity through phosphorylation of Bcl-x(L) [194-196].
Overexressed PLK1 is thought to behave as oncoprotein [197]. PLK1 is commonly found overexpressed in a majority of samples from patients with acute myeloid leukemia compared with normal progenitors [198].
PLK1 targeted therapy
Early observations that PLK1 depletion could induce apoptosis in cancer cells led to discovery and development of PLK1 inhibitors with potent antitumor activity against solid and blood cancers [199-204]. PLK inhibition is now considered a promising strategy for the treatment of AML preferably combined with conventional antileukemic chemotherapy [205,206]. First PLK1 inhibitors are currently in early clinical development in AML with promising early results. The first PLK1 inhibitor BI 2536 showed interesting clinical activity in patients with relapsed and treatment refractory AML in an early clinical trial [207]. Its successor volasertib (BI 6727) demonstrated more favorable toxicity profile and potent antileukemic activity as monotherapy and in combination with low dose aracytin in heavily pretreated AML patients and was taken to a current phase III clinical investigation [208,209].
Conclusions
AML is a highly agressive heterogenous malignant disease, classified by recurrent genetic abnormalities that define subgroups of distinct biological and clinical features. However, therapeutic approaches have stuck to “one-size fits all” conventional chemotherapy because of lack of targeted therapeutic options. Although in solid cancers a few targeted therapies have advanced to the clinical practice during the last decade, AML has notoriously been left behind despite the fact that this disease was the first human cancer genome to be sequenced and molecularly characterised. Advancements of applied technologies in molecular biology and drug discovery offer hopes that progress will be made towards more rational therapeutic approaches in AML patients. This milestone in AML therapy can only be reached through well-designed clinical trials conducted by expert teams and targeted to well characterized disease subsets. Such studies must follow resource sparing approaches because of the rarity of target patient subgroups and the highly demanding nature of such trials [210].
Acknowledgments
Eleftheria Hatzimichael is a scholar of the Hellenic Society of Haematology Foundation (2012).
Conflict of interest
The authors have no conflict of interest to declare.
References
- 1.Erba HP. Has there been progress in the treatment of older patients with acute myeloid leukemia? Best Pract Res Clin Haematol. 2010;23:495–501. doi: 10.1016/j.beha.2010.09.012. [DOI] [PubMed] [Google Scholar]
- 2.Schlenk RF, Dohner K, Krauter J, Frohling S, Corbacioglu A, Bullinger L, Habdank M, Spath D, Morgan M, Benner A, Schlegelberger B, Heil G, Ganser A, Dohner H. Mutations and treatment outcome in cytogenetically normal acute myeloid leukemia. N Engl J Med. 2008;358:1909–1918. doi: 10.1056/NEJMoa074306. [DOI] [PubMed] [Google Scholar]
- 3.Patel JP, Gonen M, Figueroa ME, Fernandez H, Sun Z, Racevskis J, Van Vlierberghe P, Dolgalev I, Thomas S, Aminova O, Huberman K, Cheng J, Viale A, Socci ND, Heguy A, Cherry A, Vance G, Higgins RR, Ketterling RP, Gallagher RE, Litzow M, van den Brink MR, Lazarus HM, Rowe JM, Luger S, Ferrando A, Paietta E, Tallman MS, Melnick A, Abdel-Wahab O, Levine RL. Prognostic relevance of integrated genetic profiling in acute myeloid leukemia. N Engl J Med. 2012;366:1079–1089. doi: 10.1056/NEJMoa1112304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Dohner H, Estey EH, Amadori S, Appelbaum FR, Buchner T, Burnett AK, Dombret H, Fenaux P, Grimwade D, Larson RA, Lo-Coco F, Naoe T, Niederwieser D, Ossenkoppele GJ, Sanz MA, Sierra J, Tallman MS, Lowenberg B, Bloomfield CD. Diagnosis and management of acute myeloid leukemia in adults: recommendations from an international expert panel, on behalf of the European LeukemiaNet. Blood. 2010;115:453–474. doi: 10.1182/blood-2009-07-235358. [DOI] [PubMed] [Google Scholar]
- 5.Litzow MR. Progress and strategies for patients with relapsed and refractory acute myeloid leukemia. Curr Opin Hematol. 2007;14:130–137. doi: 10.1097/MOH.0b013e32801684ef. [DOI] [PubMed] [Google Scholar]
- 6.Stavropoulou V, Brault L, Schwaller J. Insights into molecular pathways for targeted therapeutics in acute leukemia. Swiss Med Wkly. 2010;140:w13068. doi: 10.4414/smw.2010.13068. [DOI] [PubMed] [Google Scholar]
- 7.Flaherty KT, Infante JR, Daud A, Gonzalez R, Kefford RF, Sosman J, Hamid O, Schuchter L, Cebon J, Ibrahim N, Kudchadkar R, Burris HA 3rd, Falchook G, Algazi A, Lewis K, Long GV, Puzanov I, Lebowitz P, Singh A, Little S, Sun P, Allred A, Ouellet D, Kim KB, Patel K. Combined BRAF and MEK inhibition in melanoma with BRAF V600 mutations. N Engl J Med. 2012;367:1694–1703. doi: 10.1056/NEJMoa1210093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Sasaki T, Rodig SJ, Chirieac LR, Janne PA. The biology and treatment of EML4-ALK nonsmall cell lung cancer. Eur J Cancer. 2010;46:1773–1780. doi: 10.1016/j.ejca.2010.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Roukos DH, Murray S, Briasoulis E. Molecular genetic tools shape a roadmap towards a more accurate prognostic prediction and personalized management of cancer. Cancer Biol Ther. 2007;6:308–312. doi: 10.4161/cbt.6.3.3994. [DOI] [PubMed] [Google Scholar]
- 10.Sattler M, Scheijen B, Weisberg E, Griffin JD. Mutated tyrosine kinases as therapeutic targets in myeloid leukemias. Adv Exp Med Biol. 2003;532:121–140. doi: 10.1007/978-1-4615-0081-0_11. [DOI] [PubMed] [Google Scholar]
- 11.Quignon F, Chen Z, de The H. Retinoic acid and arsenic: towards oncogene-targeted treatments of acute promyelocytic leukaemia. Biochim Biophys Acta. 1997;1333:M53–61. doi: 10.1016/s0304-419x(97)00025-5. [DOI] [PubMed] [Google Scholar]
- 12.Tallman MS. Acute promyelocytic leukemia as a paradigm for targeted therapy. Semin Hematol. 2004;41:27–32. doi: 10.1053/j.seminhematol.2004.02.004. [DOI] [PubMed] [Google Scholar]
- 13.Cordell JL, Pulford KA, Bigerna B, Roncador G, Banham A, Colombo E, Pelicci PG, Mason DY, Falini B. Detection of normal and chimeric nucleophosmin in human cells. Blood. 1999;93:632–642. [PubMed] [Google Scholar]
- 14.Dalenc F, Drouet J, Ader I, Delmas C, Rochaix P, Favre G, Cohen-Jonathan E, Toulas C. Increased expression of a COOH-truncated nucleophosmin resulting from alternative splicing is associated with cellular resistance to ionizing radiation in HeLa cells. Int J Cancer. 2002;100:662–668. doi: 10.1002/ijc.10558. [DOI] [PubMed] [Google Scholar]
- 15.Grisendi S, Mecucci C, Falini B, Pandolfi PP. Nucleophosmin and cancer. Nat Rev Cancer. 2006;6:493–505. doi: 10.1038/nrc1885. [DOI] [PubMed] [Google Scholar]
- 16.Wang W, Budhu A, Forgues M, Wang XW. Temporal and spatial control of nucleophosmin by the Ran-Crm1 complex in centrosome duplication. Nat Cell Biol. 2005;7:823–830. doi: 10.1038/ncb1282. [DOI] [PubMed] [Google Scholar]
- 17.Gallagher SJ, Kefford RF, Rizos H. The ARF tumour suppressor. Int J Biochem Cell Biol. 2006;38:1637–1641. doi: 10.1016/j.biocel.2006.02.008. [DOI] [PubMed] [Google Scholar]
- 18.Reboutier D, Troadec MB, Cremet JY, Fukasawa K, Prigent C. Nucleophosmin/B23 activates Aurora A at the centrosome through phosphorylation of serine 89. J Cell Biol. 2012;197:19–26. doi: 10.1083/jcb.201107134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Falini B, Mason DY. Proteins encoded by genes involved in chromosomal alterations in lymphoma and leukemia: clinical value of their detection by immunocytochemistry. Blood. 2002;99:409–426. doi: 10.1182/blood.v99.2.409. [DOI] [PubMed] [Google Scholar]
- 20.Falini B, Mecucci C, Tiacci E, Alcalay M, Rosati R, Pasqualucci L, La Starza R, Diverio D, Colombo E, Santucci A, Bigerna B, Pacini R, Pucciarini A, Liso A, Vignetti M, Fazi P, Meani N, Pettirossi V, Saglio G, Mandelli F, Lo-Coco F, Pelicci PG, Martelli MF. Cytoplasmic nucleophosmin in acute myelogenous leukemia with a normal karyotype. N Engl J Med. 2005;352:254–266. doi: 10.1056/NEJMoa041974. [DOI] [PubMed] [Google Scholar]
- 21.Falini B, Nicoletti I, Martelli MF, Mecucci C. Acute myeloid leukemia carrying cytoplasmic/mutated nucleophosmin (NPMc+ AML): biologic and clinical features. Blood. 2007;109:874–885. doi: 10.1182/blood-2006-07-012252. [DOI] [PubMed] [Google Scholar]
- 22.Grummitt CG, Townsley FM, Johnson CM, Warren AJ, Bycroft M. Structural consequences of nucleophosmin mutations in acute myeloid leukemia. J Biol Chem. 2008;283:23326–23332. doi: 10.1074/jbc.M801706200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Albiero E, Madeo D, Bolli N, Giaretta I, Bona ED, Martelli MF, Nicoletti I, Rodeghiero F, Falini B. Identification and functional characterization of a cytoplasmic nucleophosmin leukaemic mutant generated by a novel exon-11 NPM1 mutation. Leukemia. 2007;21:1099–1103. doi: 10.1038/sj.leu.2404597. [DOI] [PubMed] [Google Scholar]
- 24.Pitiot AS, Santamaria I, Garcia-Suarez O, Centeno I, Astudillo A, Rayon C, Balbin M. A new type of NPM1 gene mutation in AML leading to a C-terminal truncated protein. Leukemia. 2007;21:1564–1566. doi: 10.1038/sj.leu.2404679. [DOI] [PubMed] [Google Scholar]
- 25.den Besten W, Kuo ML, Williams RT, Sherr CJ. Myeloid leukemia-associated nucleophosmin mutants perturb p53-dependent and independent activities of the Arf tumor suppressor protein. Cell Cycle. 2005;4:1593–1598. doi: 10.4161/cc.4.11.2174. [DOI] [PubMed] [Google Scholar]
- 26.Colombo E, Martinelli P, Zamponi R, Shing DC, Bonetti P, Luzi L, Volorio S, Bernard L, Pruneri G, Alcalay M, Pelicci PG. Delocalization and destabilization of the Arf tumor suppressor by the leukemia-associated NPM mutant. Cancer Res. 2006;66:3044–3050. doi: 10.1158/0008-5472.CAN-05-2378. [DOI] [PubMed] [Google Scholar]
- 27.Grandage VL, Gale RE, Linch DC, Khwaja A. PI3-kinase/Akt is constitutively active in primary acute myeloid leukaemia cells and regulates survival and chemoresistance via NFkappaB, Mapkinase and p53 pathways. Leukemia. 2005;19:586–594. doi: 10.1038/sj.leu.2403653. [DOI] [PubMed] [Google Scholar]
- 28.Cilloni D, Messa F, Rosso V, Arruga F, Defilippi I, Carturan S, Catalano R, Pautasso M, Panuzzo C, Nicoli P, Messa E, Morotti A, Iacobucci I, Martinelli G, Bracco E, Saglio G. Increase sensitivity to chemotherapeutical agents and cytoplasmatic interaction between NPM leukemic mutant and NF-kappaB in AML carrying NPM1 mutations. Leukemia. 2008;22:1234–1240. doi: 10.1038/leu.2008.68. [DOI] [PubMed] [Google Scholar]
- 29.Chou WC, Tang JL, Lin LI, Yao M, Tsay W, Chen CY, Wu SJ, Huang CF, Chiou RJ, Tseng MH, Lin DT, Lin KH, Chen YC, Tien HF. Nucleophosmin mutations in de novo acute myeloid leukemia: the age-dependent incidences and the stability during disease evolution. Cancer Res. 2006;66:3310–3316. doi: 10.1158/0008-5472.CAN-05-4316. [DOI] [PubMed] [Google Scholar]
- 30.Dohner K, Schlenk RF, Habdank M, Scholl C, Rucker FG, Corbacioglu A, Bullinger L, Frohling S, Dohner H. Mutant nucleophosmin (NPM1) predicts favorable prognosis in younger adults with acute myeloid leukemia and normal cytogenetics: interaction with other gene mutations. Blood. 2005;106:3740–3746. doi: 10.1182/blood-2005-05-2164. [DOI] [PubMed] [Google Scholar]
- 31.Falini B, Mecucci C, Tiacci E, Alcalay M, Rosati R, Pasqualucci L, La Starza R, Diverio D, Colombo E, Santucci A, Bigerna B, Pacini R, Pucciarini A, Liso A, Vignetti M, Fazi P, Meani N, Pettirossi V, Saglio G, Mandelli F, Lo-Coco F, Pelicci PG, Martelli MF. Cytoplasmic nucleophosmin in acute myelogenous leukemia with a normal karyotype. N Engl J Med. 2005;352:254–266. doi: 10.1056/NEJMoa041974. [DOI] [PubMed] [Google Scholar]
- 32.Thiede C, Koch S, Creutzig E, Steudel C, Illmer T, Schaich M, Ehninger G. Prevalence and prognostic impact of NPM1 mutations in 1485 adult patients with acute myeloid leukemia (AML) Blood. 2006;107:4011–4020. doi: 10.1182/blood-2005-08-3167. [DOI] [PubMed] [Google Scholar]
- 33.Falini B, Gionfriddo I, Cecchetti F, Ballanti S, Pettirossi V, Martelli MP. Acute myeloid leukemia with mutated nucleophosmin (NPM1): any hope for a targeted therapy? Blood Rev. 2011;25:247–254. doi: 10.1016/j.blre.2011.06.001. [DOI] [PubMed] [Google Scholar]
- 34.Schlenk RF, Dohner K, Kneba M, Gotze K, Hartmann F, Del Valle F, Kirchen H, Koller E, Fischer JT, Bullinger L, Habdank M, Spath D, Groner S, Krebs B, Kayser S, Corbacioglu A, Anhalt A, Benner A, Frohling S, Dohner H. Gene mutations and response to treatment with alltrans retinoic acid in elderly patients with acute myeloid leukemia. Results from the AMLSG Trial AML HD98B. Haematologica. 2009;94:54–60. doi: 10.3324/haematol.13378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Balusu R, Fiskus W, Rao R, Chong DG, Nalluri S, Mudunuru U, Ma H, Chen L, Venkannagari S, Ha K, Abhyankar S, Williams C, McGuirk J, Khoury HJ, Ustun C, Bhalla KN. Targeting levels or oligomerization of nucleophosmin 1 induces differentiation and loss of survival of human AML cells with mutant NPM1. Blood. 2011;118:3096–3106. doi: 10.1182/blood-2010-09-309674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Antonson P, Xanthopoulos KG. Molecular cloning, sequence, and expression patterns of the human gene encoding CCAAT/enhancer binding protein alpha (C/EBP alpha) Biochem Biophys Res Commun. 1995;215:106–113. doi: 10.1006/bbrc.1995.2439. [DOI] [PubMed] [Google Scholar]
- 37.Rosenbauer F, Tenen DG. Transcription factors in myeloid development: balancing differentiation with transformation. Nat Rev Immunol. 2007;7:105–117. doi: 10.1038/nri2024. [DOI] [PubMed] [Google Scholar]
- 38.Bereshchenko O, Mancini E, Moore S, Bilbao D, Mansson R, Luc S, Grover A, Jacobsen SE, Bryder D, Nerlov C. Hematopoietic stem cell expansion precedes the generation of committed myeloid leukemia-initiating cells in C/EBPalpha mutant AML. Cancer cell. 2009;16:390–400. doi: 10.1016/j.ccr.2009.09.036. [DOI] [PubMed] [Google Scholar]
- 39.Pabst T, Mueller BU, Zhang P, Radomska HS, Narravula S, Schnittger S, Behre G, Hiddemann W, Tenen DG. Dominant-negative mutations of CEBPA, encoding CCAAT/enhancer binding protein-alpha (C/EBPalpha), in acute myeloid leukemia. Nat Genet. 2001;27:263–270. doi: 10.1038/85820. [DOI] [PubMed] [Google Scholar]
- 40.Pabst T, Mueller BU. Complexity of CEBPA dysregulation in human acute myeloid leukemia. Clin Cancer Res. 2009;15:5303–5307. doi: 10.1158/1078-0432.CCR-08-2941. [DOI] [PubMed] [Google Scholar]
- 41.Lin LI, Chen CY, Lin DT, Tsay W, Tang JL, Yeh YC, Shen HL, Su FH, Yao M, Huang SY, Tien HF. Characterization of CEBPA mutations in acute myeloid leukemia: most patients with CEBPA mutations have biallelic mutations and show a distinct immunophenotype of the leukemic cells. Clin Cancer Res. 2005;11:1372–1379. doi: 10.1158/1078-0432.CCR-04-1816. [DOI] [PubMed] [Google Scholar]
- 42.Leroy H, Roumier C, Huyghe P, Biggio V, Fenaux P, Preudhomme C. CEBPA point mutations in hematological malignancies. Leukemia. 2005;19:329–334. doi: 10.1038/sj.leu.2403614. [DOI] [PubMed] [Google Scholar]
- 43.Preudhomme C, Sagot C, Boissel N, Cayuela JM, Tigaud I, de Botton S, Thomas X, Raffoux E, Lamandin C, Castaigne S, Fenaux P, Dombret H. Favorable prognostic significance of CEBPA mutations in patients with de novo acute myeloid leukemia: a study from the Acute Leukemia French Association (ALFA) Blood. 2002;100:2717–2723. doi: 10.1182/blood-2002-03-0990. [DOI] [PubMed] [Google Scholar]
- 44.Bienz M, Ludwig M, Leibundgut EO, Mueller BU, Ratschiller D, Solenthaler M, Fey MF, Pabst T. Risk assessment in patients with acute myeloid leukemia and a normal karyotype. Clin Cancer Res. 2005;11:1416–1424. doi: 10.1158/1078-0432.CCR-04-1552. [DOI] [PubMed] [Google Scholar]
- 45.Green CL, Koo KK, Hills RK, Burnett AK, Linch DC, Gale RE. Prognostic significance of CEBPA mutations in a large cohort of younger adult patients with acute myeloid leukemia: impact of double CEBPA mutations and the interaction with FLT3 and NPM1 mutations. J. Clin. Oncol. 2010;28:2739–2747. doi: 10.1200/JCO.2009.26.2501. [DOI] [PubMed] [Google Scholar]
- 46.Taskesen E, Bullinger L, Corbacioglu A, Sanders MA, Erpelinck CA, Wouters BJ, van der Poel-van de Luytgaarde SC, Damm F, Krauter J, Ganser A, Schlenk RF, Lowenberg B, Delwel R, Dohner H, Valk PJ, Dohner K. Prognostic impact, concurrent genetic mutations, and gene expression features of AML with CEBPA mutations in a cohort of 1182 cytogenetically normal AML patients: further evidence for CEBPA double mutant AML as a distinctive disease entity. Blood. 2011;117:2469–2475. doi: 10.1182/blood-2010-09-307280. [DOI] [PubMed] [Google Scholar]
- 47.Dufour A, Schneider F, Hoster E, Benthaus T, Ksienzyk B, Schneider S, Kakadia PM, Sauerland MC, Berdel WE, Buchner T, Wormann B, Braess J, Subklewe M, Hiddemann W, Bohlander SK, Spiekermann K. Monoallelic CEBPA mutations in normal karyotype acute myeloid leukemia: independent favorable prognostic factor within NPM1 mutated patients. Ann Hematol. 2012;91:1051–1063. doi: 10.1007/s00277-012-1423-4. [DOI] [PubMed] [Google Scholar]
- 48.Klein RD, Marcucci G. Familial Acute Myeloid Leukemia (AML) with Mutated CEBPA. In: Pagon RA, Bird TD, Dolan CR, Stephens K, Adam MP, editors. GeneReviews. Seattle (WA): 1993. [Google Scholar]
- 49.Pabst T, Mueller BU. Transcriptional dysregulation during myeloid transformation in AML. Oncogene. 2007;26:6829–6837. doi: 10.1038/sj.onc.1210765. [DOI] [PubMed] [Google Scholar]
- 50.Chan B, Weidemaier K, Yip WT, Barbara PF, Musier-Forsyth K. Intra-tRNA distance measurements for nucleocapsid proteindependent tRNA unwinding during priming of HIV reverse transcription. Proc Natl Acad Sci U S A. 1999;96:459–464. doi: 10.1073/pnas.96.2.459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Zhang S, Mantel C, Broxmeyer HE. Flt3 signaling involves tyrosyl-phosphorylation of SHP-2 and SHIP and their association with Grb2 and Shc in Baf3/Flt3 cells. J Leukoc Biol. 1999;65:372–380. doi: 10.1002/jlb.65.3.372. [DOI] [PubMed] [Google Scholar]
- 52.Lavagna-Sevenier C, Marchetto S, Birnbaum D, Rosnet O. The CBL-related protein CBLB participates in FLT3 and interleukin-7 receptor signal transduction in pro-B cells. J Biol Chem. 1998;273:14962–14967. doi: 10.1074/jbc.273.24.14962. [DOI] [PubMed] [Google Scholar]
- 53.Adolfsson J, Mansson R, Buza-Vidas N, Hultquist A, Liuba K, Jensen CT, Bryder D, Yang L, Borge OJ, Thoren LA, Anderson K, Sitnicka E, Sasaki Y, Sigvardsson M, Jacobsen SE. Identification of Flt3+ lympho-myeloid stem cells lacking erythro-megakaryocytic potential a revised road map for adult blood lineage commitment. Cell. 2005;121:295–306. doi: 10.1016/j.cell.2005.02.013. [DOI] [PubMed] [Google Scholar]
- 54.Stirewalt DL, Radich JP. The role of FLT3 in haematopoietic malignancies. Nat Rev Cancer. 2003;3:650–665. doi: 10.1038/nrc1169. [DOI] [PubMed] [Google Scholar]
- 55.Turner AM, Lin NL, Issarachai S, Lyman SD, Broudy VC. FLT3 receptor expression on the surface of normal and malignant human hematopoietic cells. Blood. 1996;88:3383–3390. [PubMed] [Google Scholar]
- 56.Nakao M, Yokota S, Iwai T, Kaneko H, Horiike S, Kashima K, Sonoda Y, Fujimoto T, Misawa S. Internal tandem duplication of the flt3 gene found in acute myeloid leukemia. Leukemia. 1996;10:1911–1918. [PubMed] [Google Scholar]
- 57.Meshinchi S, Alonzo TA, Stirewalt DL, Zwaan M, Zimmerman M, Reinhardt D, Kaspers GJ, Heerema NA, Gerbing R, Lange BJ, Radich JP. Clinical implications of FLT3 mutations in pediatric AML. Blood. 2006;108:3654–3661. doi: 10.1182/blood-2006-03-009233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Stirewalt DL, Kopecky KJ, Meshinchi S, Engel JH, Pogosova-Agadjanyan EL, Linsley J, Slovak ML, Willman CL, Radich JP. Size of FLT3 internal tandem duplication has prognostic significance in patients with acute myeloid leukemia. Blood. 2006;107:3724–3726. doi: 10.1182/blood-2005-08-3453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kiyoi H, Ohno R, Ueda R, Saito H, Naoe T. Mechanism of constitutive activation of FLT3 with internal tandem duplication in the juxtamembrane domain. Oncogene. 2002;21:2555–2563. doi: 10.1038/sj.onc.1205332. [DOI] [PubMed] [Google Scholar]
- 60.Griffith J, Black J, Faerman C, Swenson L, Wynn M, Lu F, Lippke J, Saxena K. The structural basis for autoinhibition of FLT3 by the juxtamembrane domain. Mol Cell. 2004;13:169–178. doi: 10.1016/s1097-2765(03)00505-7. [DOI] [PubMed] [Google Scholar]
- 61.Mizuki M, Fenski R, Halfter H, Matsumura I, Schmidt R, Muller C, Gruning W, Kratz-Albers K, Serve S, Steur C, Buchner T, Kienast J, Kanakura Y, Berdel WE, Serve H. Flt3 mutations from patients with acute myeloid leukemia induce transformation of 32D cells mediated by the Ras and STAT5 pathways. Blood. 2000;96:3907–3914. [PubMed] [Google Scholar]
- 62.Yamamoto Y, Kiyoi H, Nakano Y, Suzuki R, Kodera Y, Miyawaki S, Asou N, Kuriyama K, Yagasaki F, Shimazaki C, Akiyama H, Saito K, Nishimura M, Motoji T, Shinagawa K, Takeshita A, Saito H, Ueda R, Ohno R, Naoe T. Activating mutation of D835 within the activation loop of FLT3 in human hematologic malignancies. Blood. 2001;97:2434–2439. doi: 10.1182/blood.v97.8.2434. [DOI] [PubMed] [Google Scholar]
- 63.Thiede C, Steudel C, Mohr B, Schaich M, Schakel U, Platzbecker U, Wermke M, Bornhauser M, Ritter M, Neubauer A, Ehninger G, Illmer T. Analysis of FLT3-activating mutations in 979 patients with acute myelogenous leukemia: association with FAB subtypes and identification of subgroups with poor prognosis. Blood. 2002;99:4326–4335. doi: 10.1182/blood.v99.12.4326. [DOI] [PubMed] [Google Scholar]
- 64.Spiekermann K, Bagrintseva K, Schoch C, Haferlach T, Hiddemann W, Schnittger S. A new and recurrent activating length mutation in exon 20 of the FLT3 gene in acute myeloid leukemia. Blood. 2002;100:3423–3425. doi: 10.1182/blood-2002-03-0953. [DOI] [PubMed] [Google Scholar]
- 65.Choudhary C, Schwable J, Brandts C, Tickenbrock L, Sargin B, Kindler T, Fischer T, Berdel WE, Muller-Tidow C, Serve H. AML-associated Flt3 kinase domain mutations show signal transduction differences compared with Flt3 ITD mutations. Blood. 2005;106:265–273. doi: 10.1182/blood-2004-07-2942. [DOI] [PubMed] [Google Scholar]
- 66.Grundler R, Miething C, Thiede C, Peschel C, Duyster J. FLT3-ITD and tyrosine kinase domain mutants induce 2 distinct phenotypes in a murine bone marrow transplantation model. Blood. 2005;105:4792–4799. doi: 10.1182/blood-2004-11-4430. [DOI] [PubMed] [Google Scholar]
- 67.Schnittger S, Schoch C, Dugas M, Kern W, Staib P, Wuchter C, Loffler H, Sauerland CM, Serve H, Buchner T, Haferlach T, Hiddemann W. Analysis of FLT3 length mutations in 1003 patients with acute myeloid leukemia: correlation to cytogenetics, FAB subtype, and prognosis in the AMLCG study and usefulness as a marker for the detection of minimal residual disease. Blood. 2002;100:59–66. doi: 10.1182/blood.v100.1.59. [DOI] [PubMed] [Google Scholar]
- 68.Kottaridis PD, Gale RE, Frew ME, Harrison G, Langabeer SE, Belton AA, Walker H, Wheatley K, Bowen DT, Burnett AK, Goldstone AH, Linch DC. The presence of a FLT3 internal tandem duplication in patients with acute myeloid leukemia (AML) adds important prognostic information to cytogenetic risk group and response to the first cycle of chemotherapy: analysis of 854 patients from the United King dom Medical Research Council AML 10 and 12 trials. Blood. 2001;98:1752–1759. doi: 10.1182/blood.v98.6.1752. [DOI] [PubMed] [Google Scholar]
- 69.Gale RE, Green C, Allen C, Mead AJ, Burnett AK, Hills RK, Linch DC. The impact of FLT3 internal tandem duplication mutant level, number, size, and interaction with NPM1 mutations in a large cohort of young adult patients with acute myeloid leukemia. Blood. 2008;111:2776–2784. doi: 10.1182/blood-2007-08-109090. [DOI] [PubMed] [Google Scholar]
- 70.Sawyers CL. Finding the next Gleevec: FLT3 targeted kinase inhibitor therapy for acute myeloid leukemia. Cancer Cell. 2002;1:413–415. doi: 10.1016/s1535-6108(02)00080-6. [DOI] [PubMed] [Google Scholar]
- 71.Fathi AT, Chabner BA. FLT3 inhibition as therapy in acute myeloid leukemia: a record of trials and tribulations. Oncologist. 2011;16:1162–1174. doi: 10.1634/theoncologist.2011-0084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Wiernik PH. FLT3 inhibitors for the treatment of acute myeloid leukemia. Clin Adv Hematol Oncol. 2010;8:429–436, 444. [PubMed] [Google Scholar]
- 73.Smith CC, Wang Q, Chin CS, Salerno S, Damon LE, Levis MJ, Perl AE, Travers KJ, Wang S, Hunt JP, Zarrinkar PP, Schadt EE, Kasarskis A, Kuriyan J, Shah NP. Validation of ITD mutations in FLT3 as a therapeutic target in human acute myeloid leukaemia. Nature. 2012;485:260–263. doi: 10.1038/nature11016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Zhang W, Konopleva M, Shi YX, McQueen T, Harris D, Ling X, Estrov Z, Quintas-Cardama A, Small D, Cortes J, Andreeff M. Mutant FLT3: a direct target of sorafenib in acute myelogenous leukemia. J Natl Cancer Inst. 2008;100:184–198. doi: 10.1093/jnci/djm328. [DOI] [PubMed] [Google Scholar]
- 75.Sharma M, Ravandi F, Bayraktar UD, Chiattone A, Bashir Q, Giralt S, Chen J, Qazilbash M, Kebriaei P, Konopleva M, Andreeff M, Cortes J, McCue D, Kantarjian H, Champlin RE, de Lima M. Treatment of FLT3-ITD-positive acute myeloid leukemia relapsing after allogeneic stem cell transplantation with sorafenib. Biol Blood Marrow Transplant. 2011;17:1874–1877. doi: 10.1016/j.bbmt.2011.07.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Scholl S, Spies-Weisshart B, Klink A, Muegge LO, Fricke HJ, Hochhaus A. Secondary resistance to sorafenib in two patients with acute myeloid leukemia (AML) harboring FLT3-ITD mutations. Ann Hematol. 2011;90:473–475. doi: 10.1007/s00277-010-1027-9. [DOI] [PubMed] [Google Scholar]
- 77.Pallis A, Briasoulis E, Linardou H, Papadimitriou C, Bafaloukos D, Kosmidis P, Murray S. Mechanisms of resistance to epidermal growth factor receptor tyrosine kinase inhibitors in patients with advanced non-small-cell lung cancer: clinical and molecular considerations. Curr Med Chem. 2011;18:1613–1628. doi: 10.2174/092986711795471383. [DOI] [PubMed] [Google Scholar]
- 78.Parmar A, Marz S, Rushton S, Holzwarth C, Lind K, Kayser S, Dohner K, Peschel C, Oostendorp RA, Gotze KS. Stromal niche cells protect early leukemic FLT3-ITD+ progenitor cells against first-generation FLT3 tyrosine kinase inhibitors. Cancer Res. 2011;71:4696–4706. doi: 10.1158/0008-5472.CAN-10-4136. [DOI] [PubMed] [Google Scholar]
- 79.Zauli G, Celeghini C, Melloni E, Voltan R, Ongari M, Tiribelli M, di Iasio MG, Lanza F, Secchiero P. The Sorafenib plus Nutlin-3 combination promotes synergistic cytotoxicity in acute myeloid leukemic cells irrespectively of the FLT3 and p53 status. Haematologica. 2012;97:1722–30. doi: 10.3324/haematol.2012.062083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Tzakos AG, Fokas D, Johannes C, Moussis V, Hatzimichael E, Briasoulis E. Targeting oncogenic protein-protein interactions by diversity oriented synthesis and combinatorial chemistry approaches. Molecules. 2011;16:4408–4427. doi: 10.3390/molecules16064408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Williams AB, Li L, Nguyen B, Brown P, Levis M, Small D. Fluvastatin inhibits FLT3 glycosylation in human and murine cells and prolongs survival of mice with FLT3-ITD leukemia. Blood. 2012;120:3069–3079. doi: 10.1182/blood-2012-01-403493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Hart S, Goh KC, Novotny-Diermayr V, Tan YC, Madan B, Amalini C, Ong LC, Kheng B, Cheong A, Zhou J, Chng WJ, Wood JM. Pacritinib (SB1518), a JAK2/FLT3 inhibitor for the treatment of acute myeloid leukemia. Blood Cancer J. 2011;1:e44. doi: 10.1038/bcj.2011.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Zhou J, Bi C, Janakakumara JV, Liu SC, Chng WJ, Tay KG, Poon LF, Xie Z, Palaniyandi S, Yu H, Glaser KB, Albert DH, Davidsen SK, Chen CS. Enhanced activation of STAT pathways and overexpression of survivin confer resistance to FLT3 inhibitors and could be therapeutic targets in AML. Blood. 2009;113:4052–4062. doi: 10.1182/blood-2008-05-156422. [DOI] [PubMed] [Google Scholar]
- 84.Ikezoe T, Kojima S, Furihata M, Yang J, Nishioka C, Takeuchi A, Isaka M, Koeffler HP, Yokoyama A. Expression of p-JAK2 predicts clinical outcome and is a potential molecular target of acute myelogenous leukemia. Int J Cancer. 2011;129:2512–2521. doi: 10.1002/ijc.25910. [DOI] [PubMed] [Google Scholar]
- 85.Zarrinkar PP, Gunawardane RN, Cramer MD, Gardner MF, Brigham D, Belli B, Karaman MW, Pratz KW, Pallares G, Chao Q, Sprankle KG, Patel HK, Levis M, Armstrong RC, James J, Bhagwat SS. AC220 is a uniquely potent and selective inhibitor of FLT3 for the treatment of acute myeloid leukemia (AML) Blood. 2009;114:2984–2992. doi: 10.1182/blood-2009-05-222034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Cortes JE, Perl AE, Smith CC, Kovacsovics T, Dombret H, Dohner H, Steffen B, Pigneux A, Rousselot P, Krauter J, Martinelli G, Estey EH, Burnett AK, Ho AD, Ifrah N, de Witt T, Corringham R, James J, Lilienfeld D, Leo E, Gammon G, Levis MJ. A Phase II Open-Label, Ac220 Monotherapy Efficacy Study In Patients with Refractory/Relapsed Flt3-Itd Positive Acute Myeloid Leukemia: Updated Interim Results. ASH Annual Meeting Abstracts. 2011;118:2576. [Google Scholar]
- 87.Knapper S, Burnett AK, Littlewood T, Kell WJ, Agrawal S, Chopra R, Clark R, Levis MJ, Small D. A phase 2 trial of the FLT3 inhibitor lestaurtinib (CEP701) as first-line treatment for older patients with acute myeloid leukemia not considered fit for intensive chemotherapy. Blood. 2006;108:3262–3270. doi: 10.1182/blood-2006-04-015560. [DOI] [PubMed] [Google Scholar]
- 88.Knapper S, Mills KI, Gilkes AF, Austin SJ, Walsh V, Burnett AK. The effects of lestaurtinib (CEP701) and PKC412 on primary AML blasts: the induction of cytotoxicity varies with dependence on FLT3 signaling in both FLT3-mutated and wild-type cases. Blood. 2006;108:3494–3503. doi: 10.1182/blood-2006-04-015487. [DOI] [PubMed] [Google Scholar]
- 89.Levis M, Ravandi F, Wang ES, Baer MR, Perl A, Coutre S, Erba H, Stuart RK, Baccarani M, Cripe LD, Tallman MS, Meloni G, Godley LA, Langston AA, Amadori S, Lewis ID, Nagler A, Stone R, Yee K, Advani A, Douer D, Wiktor-Jedrzejczak W, Juliusson G, Litzow MR, Petersdorf S, Sanz M, Kantarjian HM, Sato T, Tremmel L, Bensen-Kennedy DM, Small D, Smith BD. Results from a randomized trial of salvage chemotherapy followed by lestaurtinib for patients with FLT3 mutant AML in first relapse. Blood. 2011;117:3294–3301. doi: 10.1182/blood-2010-08-301796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Stone RM, Fischer T, Paquette R, Schiller G, Schiffer CA, Ehninger G, Cortes J, Kantarjian HM, Deangelo DJ, Huntsman-Labed A, Dutreix C, Del Corral A, Giles F. Phase IB study of the FLT3 kinase inhibitor midostaurin with chemotherapy in younger newly diagnosed adult patients with acute myeloid leukemia. Leukemia. 2012;26:2061–2068. doi: 10.1038/leu.2012.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Fischer T, Stone RM, Deangelo DJ, Galinsky I, Estey E, Lanza C, Fox E, Ehninger G, Feldman EJ, Schiller GJ, Klimek VM, Nimer SD, Gilliland DG, Dutreix C, Huntsman-Labed A, Virkus J, Giles FJ. Phase IIB trial of oral Midostaurin (PKC412), the FMS-like tyrosine kinase 3 receptor (FLT3) and multi-targeted kinase inhibitor, in patients with acute myeloid leukemia and high-risk myelodysplastic syndrome with either wild-type or mutated FLT3. J. Clin. Oncol. 2010;28:4339–4345. doi: 10.1200/JCO.2010.28.9678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Novotny-Diermayr V, Hart S, Goh KC, Cheong A, Ong LC, Hentze H, Pasha MK, Jayaraman R, Ethirajulu K, Wood JM. The oral HDAC inhibitor pracinostat (SB939) is efficacious and synergistic with the JAK2 inhibitor pacritinib (SB1518) in preclinical models of AML. Blood Cancer J. 2012;2:e69. doi: 10.1038/bcj.2012.14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Basavapathruni A, Jin L, Daigle SR, Majer CR, Therkelsen CA, Wigle TJ, Kuntz KW, Chesworth R, Pollock RM, Scott MP, Moyer MP, Richon VM, Copeland RA, Olhava EJ. Conformational Adaptation Drives Potent, Selective and Durable Inhibition of the Human Protein Methyltransferase DOT1L. Chem Biol Drug Des. 2012;80:971–980. doi: 10.1111/cbdd.12050. [DOI] [PubMed] [Google Scholar]
- 94.Smith CC, Perl AE, Lasater E, Zhang C, Jeschke GR, Damon LE, Carroll M, Shah NP. PLX3397 Is An Investigational Selective FLT3 Inhibitor That Retains Activity Against the Clinically-Relevant FLT3-ITD/F691L “Gatekeeper” Mutation in Vitro. ASH Annual Meeting Abstracts. 2011;118:764. [Google Scholar]
- 95.Grundy M, Seedhouse C, Shang S, Richardson J, Russell N, Pallis M. The FLT3 internal tandem duplication mutation is a secondary target of the aurora B kinase inhibitor AZD1152- HQPA in acute myelogenous leukemia cells. Mol Cancer Ther. 2010;9:661–672. doi: 10.1158/1535-7163.MCT-09-1144. [DOI] [PubMed] [Google Scholar]
- 96.Burton E, Wong B, Zhang J, West B, Bollag G, Habets G, Galanis A, Nguyen H, Arowojolu O, Rajhowa T, Levis MJ. The Novel Inhibitor PLX3397 Effectively Inhibits FLT3-Mutant AML. ASH Annual Meeting Abstracts. 2011;118:3632. [Google Scholar]
- 97.Fiskus W, Smith CC, Smith J, Wise SC, Lasater E, Damon LE, Salerno S, Fleming A, Reyes R, Ganguly S, Berger MS, Rutkoski TJ, McGuirk J, Shah N, Bhalla KN. Activity of Allosteric, Switch-Pocket, ABL/FLT3 Kinase Inhibitor DCC2036 Against Cultured and Primary AML Progenitors with FLT-ITD or FLT3 Kinase Domain Mutations. ASH Annual Meeting Abstracts. 2011;118:2611. [Google Scholar]
- 98.Marcucci G, Mrozek K, Ruppert AS, Maharry K, Kolitz JE, Moore JO, Mayer RJ, Pettenati MJ, Powell BL, Edwards CG, Sterling LJ, Vardiman JW, Schiffer CA, Carroll AJ, Larson RA, Bloomfield CD. Prognostic factors and outcome of core binding factor acute myeloid leukemia patients with t(8; 21) differ from those of patients with inv(16): a Cancer and Leukemia Group B study. J. Clin. Oncol. 2005;23:5705–5717. doi: 10.1200/JCO.2005.15.610. [DOI] [PubMed] [Google Scholar]
- 99.Paschka P, Marcucci G, Ruppert AS, Mrozek K, Chen H, Kittles RA, Vukosavljevic T, Perrotti D, Vardiman JW, Carroll AJ, Kolitz JE, Larson RA, Bloomfield CD. Adverse prognostic significance of KIT mutations in adult acute myeloid leukemia with inv(16) and t(8; 21): a Cancer and Leukemia Group B Study. J. Clin. Oncol. 2006;24:3904–3911. doi: 10.1200/JCO.2006.06.9500. [DOI] [PubMed] [Google Scholar]
- 100.Lee JW, Kim YG, Soung YH, Han KJ, Kim SY, Rhim HS, Min WS, Nam SW, Park WS, Lee JY, Yoo NJ, Lee SH. The JAK2 V617F mutation in de novo acute myelogenous leukemias. Oncogene. 2006;25:1434–1436. doi: 10.1038/sj.onc.1209163. [DOI] [PubMed] [Google Scholar]
- 101.Dohner K, Du J, Corbacioglu A, Scholl C, Schlenk RF, Dohner H. JAK2V617F mutations as cooperative genetic lesions in t(8; 21)-positive acute myeloid leukemia. Haematologica. 2006;91:1569–1570. [PubMed] [Google Scholar]
- 102.Tajima F, Kawatani T, Ishiga K, Nanba E, Kawasaki H. Serum soluble c-kit receptor and expression of c-kit protein and mRNA in acute myeloid leukemia. Eur J Haematol. 1998;60:289–296. doi: 10.1111/j.1600-0609.1998.tb01042.x. [DOI] [PubMed] [Google Scholar]
- 103.Linnekin D. Early signaling pathways activated by c-Kit in hematopoietic cells. Int J Biochem Cell Biol. 1999;31:1053–1074. doi: 10.1016/s1357-2725(99)00078-3. [DOI] [PubMed] [Google Scholar]
- 104.Nanri T, Matsuno N, Kawakita T, Suzushima H, Kawano F, Mitsuya H, Asou N. Mutations in the receptor tyrosine kinase pathway are associated with clinical outcome in patients with acute myeloblastic leukemia harboring t(8; 21) (q22; q22) Leukemia. 2005;19:1361–1366. doi: 10.1038/sj.leu.2403803. [DOI] [PubMed] [Google Scholar]
- 105.Boissel N, Leroy H, Brethon B, Philippe N, de Botton S, Auvrignon A, Raffoux E, Leblanc T, Thomas X, Hermine O, Quesnel B, Baruchel A, Leverger G, Dombret H, Preudhomme C. Incidence and prognostic impact of c-Kit, FLT3, and Ras gene mutations in core binding factor acute myeloid leukemia (CBF-AML) Leukemia. 2006;20:965–970. doi: 10.1038/sj.leu.2404188. [DOI] [PubMed] [Google Scholar]
- 106.Paschka P, Du J, Schlenk RF, Gaidzik VI, Bullinger L, Corbacioglu A, Spath D, Kayser S, Schlegelberger B, Krauter J, Ganser A, Kohne CH, Held G, von Lilienfeld-Toal M, Kirchen H, Rummel M, Gotze K, Horst HA, Ringhoffer M, Lubbert M, Wattad M, Salih HR, Kundgen A, Dohner H, Dohner K. Secondary genetic lesions in acute myeloid leukemia with inv(16) or t(16; 16): a study of the German-Austrian AML Study Group (AMLSG) Blood. 2012;121:170–7. doi: 10.1182/blood-2012-05-431486. [DOI] [PubMed] [Google Scholar]
- 107.Care RS, Valk PJ, Goodeve AC, Abu-Duhier FM, Geertsma-Kleinekoort WM, Wilson GA, Gari MA, Peake IR, Lowenberg B, Reilly JT. Incidence and prognosis of c-KIT and FLT3 mutations in core binding factor (CBF) acute myeloid leukaemias. Br J Haematol. 2003;121:775–777. doi: 10.1046/j.1365-2141.2003.04362.x. [DOI] [PubMed] [Google Scholar]
- 108.Hirota S, Isozaki K, Moriyama Y, Hashimoto K, Nishida T, Ishiguro S, Kawano K, Hanada M, Kurata A, Takeda M, Muhammad Tunio G, Matsuzawa Y, Kanakura Y, Shinomura Y, Kitamura Y. Gain-of-function mutations of c-kit in human gastrointestinal stromal tumors. Science. 1998;279:577–580. doi: 10.1126/science.279.5350.577. [DOI] [PubMed] [Google Scholar]
- 109.Heinrich MC, Blanke CD, Druker BJ, Corless CL. Inhibition of KIT tyrosine kinase activity: a novel molecular approach to the treatment of KIT-positive malignancies. J. Clin. Oncol. 2002;20:1692–1703. doi: 10.1200/JCO.2002.20.6.1692. [DOI] [PubMed] [Google Scholar]
- 110.Cortes J, Giles F, O’Brien S, Thomas D, Albitar M, Rios MB, Talpaz M, Garcia-Manero G, Faderl S, Letvak L, Salvado A, Kantarjian H. Results of imatinib mesylate therapy in patients with refractory or recurrent acute myeloid leukemia, high-risk myelodysplastic syndrome, and myeloproliferative disorders. Cancer. 2003;97:2760–2766. doi: 10.1002/cncr.11416. [DOI] [PubMed] [Google Scholar]
- 111.Kindler T, Breitenbuecher F, Marx A, Beck J, Hess G, Weinkauf B, Duyster J, Peschel C, Kirkpatrick CJ, Theobald M, Gschaidmeier H, Huber C, Fischer T. Efficacy and safety of imatinib in adult patients with c-kit-positive acute myeloid leukemia. Blood. 2004;103:3644–3654. doi: 10.1182/blood-2003-06-2071. [DOI] [PubMed] [Google Scholar]
- 112.Giles FJ, Stopeck AT, Silverman LR, Lancet JE, Cooper MA, Hannah AL, Cherrington JM, O’Farrell AM, Yuen HA, Louie SG, Hong W, Cortes JE, Verstovsek S, Albitar M, O’Brien SM, Kantarjian HM, Karp JE. SU5416, a small molecule tyrosine kinase receptor inhibitor, has biologic activity in patients with refractory acute myeloid leukemia or myelodysplastic syndromes. Blood. 2003;102:795–801. doi: 10.1182/blood-2002-10-3023. [DOI] [PubMed] [Google Scholar]
- 113.Faderl S, Bueso-Ramos C, Liu Z, Pal A, Bornmann W, Ciurea DV, Harris D, Hazan-Halevy I, Kantarjian HM, Estrov Z. Kit inhibitor APcK110 extends survival in an AML xenograft mouse model. Invest New Drugs. 2011;29:1094–1097. doi: 10.1007/s10637-010-9459-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Neubauer H, Cumano A, Muller M, Wu H, Huffstadt U, Pfeffer K. Jak2 deficiency defines an essential developmental checkpoint in definitive hematopoiesis. Cell. 1998;93:397–409. doi: 10.1016/s0092-8674(00)81168-x. [DOI] [PubMed] [Google Scholar]
- 115.Witthuhn BA, Quelle FW, Silvennoinen O, Yi T, Tang B, Miura O, Ihle JN. JAK2 associates with the erythropoietin receptor and is tyrosine phosphorylated and activated following stimulation with erythropoietin. Cell. 1993;74:227–236. doi: 10.1016/0092-8674(93)90414-l. [DOI] [PubMed] [Google Scholar]
- 116.Argetsinger LS, Campbell GS, Yang X, Witthuhn BA, Silvennoinen O, Ihle JN, Carter-Su C. Identification of JAK2 as a growth hormone receptor-associated tyrosine kinase. Cell. 1993;74:237–244. doi: 10.1016/0092-8674(93)90415-m. [DOI] [PubMed] [Google Scholar]
- 117.James C, Ugo V, Le Couedic JP, Staerk J, Delhommeau F, Lacout C, Garcon L, Raslova H, Berger R, Bennaceur-Griscelli A, Villeval JL, Constantinescu SN, Casadevall N, Vainchenker W. A unique clonal JAK2 mutation leading to constitutive signalling causes polycythaemia vera. Nature. 2005;434:1144–1148. doi: 10.1038/nature03546. [DOI] [PubMed] [Google Scholar]
- 118.Schneider F, Bohlander SK, Schneider S, Papadaki C, Kakadyia P, Dufour A, Vempati S, Unterhalt M, Feuring-Buske M, Buske C, Braess J, Wandt H, Hiddemann W, Spiekermann K. AML1-ETO meets JAK2: clinical evidence for the two hit model of leukemogenesis from a myeloproliferative syndrome progressing to acute myeloid leukemia. Leukemia. 2007;21:2199–2201. doi: 10.1038/sj.leu.2404830. [DOI] [PubMed] [Google Scholar]
- 119.Illmer T, Schaich M, Ehninger G, Thiede C. Tyrosine kinase mutations of JAK2 are rare events in AML but influence prognosis of patients with CBF-leukemias. Haematologica. 2007;92:137–138. doi: 10.3324/haematol.10489. [DOI] [PubMed] [Google Scholar]
- 120.Iwanaga E, Nanri T, Matsuno N, Kawakita T, Mitsuya H, Asou N. A JAK2-V617F activating mutation in addition to KIT and FLT3 mutations is associated with clinical outcome in patients with t(8; 21)(q22; q22) acute myeloid leukemia. Haematologica. 2009;94:433–435. doi: 10.3324/haematol.13283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Swaminathan S, Madkaikar M, Ghosh K, Vundinti BR, Kerketta L, Gupta M. Novel immunophenotypic and morphologic presentation in acute myeloid leukemia (AML) with JAK2 V617F mutation. Eur J Haematol. 2010;84:180–182. doi: 10.1111/j.1600-0609.2009.01359.x. [DOI] [PubMed] [Google Scholar]
- 122.Peeters P, Raynaud SD, Cools J, Wlodarska I, Grosgeorge J, Philip P, Monpoux F, Van Rompaey L, Baens M, Van den Berghe H, Marynen P. Fusion of TEL, the ETS-variant gene 6 (ETV6), to the receptor-associated kinase JAK2 as a result of t(9; 12) in a lymphoid and t(9; 15; 12) in a myeloid leukemia. Blood. 1997;90:2535–2540. [PubMed] [Google Scholar]
- 123.Reiter A, Walz C, Watmore A, Schoch C, Blau I, Schlegelberger B, Berger U, Telford N, Aruliah S, Yin JA, Vanstraelen D, Barker HF, Taylor PC, O’Driscoll A, Benedetti F, Rudolph C, Kolb HJ, Hochhaus A, Hehlmann R, Chase A, Cross NC. The t(8; 9)(p22; p24) is a recurrent abnormality in chronic and acute leukemia that fuses PCM1 to JAK2. Cancer Res. 2005;65:2662–2667. doi: 10.1158/0008-5472.CAN-04-4263. [DOI] [PubMed] [Google Scholar]
- 124.Lacronique V, Boureux A, Valle VD, Poirel H, Quang CT, Mauchauffe M, Berthou C, Lessard M, Berger R, Ghysdael J, Bernard OA. A TELJAK2 fusion protein with constitutive kinase activity in human leukemia. Science. 1997;278:1309–1312. doi: 10.1126/science.278.5341.1309. [DOI] [PubMed] [Google Scholar]
- 125.Ikezoe T, Kojima S, Furihata M, Yang J, Nishioka C, Takeuchi A, Isaka M, Koeffler HP, Yokoyama A. Expression of p-JAK2 predicts clinical outcome and is a potential molecular target of acute myelogenous leukemia. Int J Cancer. 2011;129:2512–2521. doi: 10.1002/ijc.25910. [DOI] [PubMed] [Google Scholar]
- 126.Kontzias A, Kotlyar A, Laurence A, Changelian P, O’Shea JJ. Jakinibs: a new class of kinase inhibitors in cancer and autoimmune disease. Curr Opin Pharmacol. 2012;12:464–470. doi: 10.1016/j.coph.2012.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Harrison C, Kiladjian JJ, Al-Ali HK, Gisslinger H, Waltzman R, Stalbovskaya V, McQuitty M, Hunter DS, Levy R, Knoops L, Cervantes F, Vannucchi AM, Barbui T, Barosi G. JAK inhibition with ruxolitinib versus best available therapy for myelofibrosis. N Engl J Med. 2012;366:787–798. doi: 10.1056/NEJMoa1110556. [DOI] [PubMed] [Google Scholar]
- 128.Bos JL. Ras oncogenes in human cancer: a review. Cancer Res. 1989;49:4682–4689. [PubMed] [Google Scholar]
- 129.Ritter M, Kim TD, Lisske P, Thiede C, Schaich M, Neubauer A. Prognostic significance of NRAS and K-RAS mutations in 232 patients with acute myeloid leukemia. Haematologica. 2004;89:1397–1399. [PubMed] [Google Scholar]
- 130.Neubauer A, Dodge RK, George SL, Davey FR, Silver RT, Schiffer CA, Mayer RJ, Ball ED, Wurster-Hill D, Bloomfield CD, et al. Prognostic importance of mutations in the ras proto-oncogenes in de novo acute myeloid leukemia. Blood. 1994;83:1603–1611. [PubMed] [Google Scholar]
- 131.Kiyoi H, Naoe T, Nakano Y, Yokota S, Minami S, Miyawaki S, Asou N, Kuriyama K, Jinnai I, Shimazaki C, Akiyama H, Saito K, Oh H, Motoji T, Omoto E, Saito H, Ohno R, Ueda R. Prognostic implication of FLT3 and N-RAS gene mutations in acute myeloid leukemia. Blood. 1999;93:3074–3080. [PubMed] [Google Scholar]
- 132.Coghlan DW, Morley AA, Matthews JP, Bishop JF. The incidence and prognostic significance of mutations in codon 13 of the N-ras gene in acute myeloid leukemia. Leukemia. 1994;8:1682–1687. [PubMed] [Google Scholar]
- 133.Neubauer A, Maharry K, Mrozek K, Thiede C, Marcucci G, Paschka P, Mayer RJ, Larson RA, Liu ET, Bloomfield CD. Patients with acute myeloid leukemia and RAS mutations benefit most from postremission high-dose cytarabine: a Cancer and Leukemia Group B study. J. Clin. Oncol. 2008;26:4603–4609. doi: 10.1200/JCO.2007.14.0418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Paschka P, Marcucci G, Ruppert AS, Mrozek K, Chen H, Kittles RA, Vukosavljevic T, Perrotti D, Vardiman JW, Carroll AJ, Kolitz JE, Larson RA, Bloomfield CD. Adverse prognostic significance of KIT mutations in adult acute myeloid leukemia with inv(16) and t(8; 21): a Cancer and Leukemia Group B Study. J. Clin. Oncol. 2006;24:3904–3911. doi: 10.1200/JCO.2006.06.9500. [DOI] [PubMed] [Google Scholar]
- 135.Alvarez RH, Kantarjian H, Garcia-Manero G, Estrov Z, Ravandi-Kashani F, Verstovsek S, Giles F, O’Brien S, Asa Koller C, Faderl S, Thomas D, Wright JJ, Cortes J. Farnesyl Transferase Inhibitor (Tipifarnib, Zarnestra; Z) in Combination with Standard Chemotherapy with Idarubicin (Ida) and Cytarabine (ara-C) for Patients (pts) with Newly Diagnosed Acute Myeloid Leukemia (AML) or High-Risk Myelodysplastic Syndrome (MDS) ASH Annual Meeting Abstracts. 2006;108:1999. [Google Scholar]
- 136.Erba HP, Kopecky KJ, Kirschbaum MH, Tallman MS, Larson RA, Willman CL, Slovak ML, Gundacker HM, Appelbaum FR. Phase II Studies of Different Schedules and Doses of the Farnesyl Transferase Inhibitor Tipifarnib (R115777, Zarnestra, NSC-702818) for Patients of Age 70 or Older with Previously Untreated Acute Myeloid Leukemia (AML): A North American Intergroup Study (S0432) ASH Annual Meeting Abstracts. 2007;110:440. [Google Scholar]
- 137.Burnett AK, Russell NH, Culligan D, Cavanagh J, Kell J, Wheatley K, Virchis A, Hills RK, Milligan D. The addition of the farnesyl transferase inhibitor, tipifarnib, to low dose cytarabine does not improve outcome for older patients with AML. Br J Haematol. 2012;158:519–522. doi: 10.1111/j.1365-2141.2012.09165.x. [DOI] [PubMed] [Google Scholar]
- 138.Widemann BC, Arceci RJ, Jayaprakash N, Fox E, Zannikos P, Goodspeed W, Goodwin A, Wright JJ, Blaney SM, Adamson PC, Balis FM. Phase 1 trial and pharmacokinetic study of the farnesyl transferase inhibitor tipifarnib in children and adolescents with refractory leukemias: a report from the Children’s Oncology Group. Pediatr Blood Cancer. 2011;56:226–233. doi: 10.1002/pbc.22775. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Ravoet C, Mineur P, Robin V, Debusscher L, Bosly A, Andre M, El Housni H, Soree A, Bron D, Martiat P. Farnesyl transferase inhibitor (lonafarnib) in patients with myelodysplastic syndrome or secondary acute myeloid leukaemia: a phase II study. Ann Hematol. 2008;87:881–885. doi: 10.1007/s00277-008-0536-2. [DOI] [PubMed] [Google Scholar]
- 140.Haselbeck RJ, McAlister-Henn L. Function and expression of yeast mitochondrial NADand NADP-specific isocitrate dehydrogenases. J Biol Chem. 1993;268:12116–12122. [PubMed] [Google Scholar]
- 141.Plaut GW, Cook M, Aogaichi T. The subcellular location of isozymes of NADP-isocitrate dehydrogenase in tissues from pig, ox and rat. Biochim Biophys Acta. 1983;760:300–308. doi: 10.1016/0304-4165(83)90177-0. [DOI] [PubMed] [Google Scholar]
- 142.Narahara K, Kimura S, Kikkawa K, Takahashi Y, Wakita Y, Kasai R, Nagai S, Nishibayashi Y, Kimoto H. Probable assignment of soluble isocitrate dehydrogenase (IDH1) to 2q33.3. Hum Genet. 1985;71:37–40. doi: 10.1007/BF00295665. [DOI] [PubMed] [Google Scholar]
- 143.Geisbrecht BV, Gould SJ. The human PICD gene encodes a cytoplasmic and peroxisomal NADP(+)-dependent isocitrate dehydrogenase. J Biol Chem. 1999;274:30527–30533. doi: 10.1074/jbc.274.43.30527. [DOI] [PubMed] [Google Scholar]
- 144.Parsons DW, Jones S, Zhang X, Lin JC, Leary RJ, Angenendt P, Mankoo P, Carter H, Siu IM, Gallia GL, Olivi A, McLendon R, Rasheed BA, Keir S, Nikolskaya T, Nikolsky Y, Busam DA, Tekleab H, Diaz LA Jr, Hartigan J, Smith DR, Strausberg RL, Marie SK, Shinjo SM, Yan H, Riggins GJ, Bigner DD, Karchin R, Papadopoulos N, Parmigiani G, Vogelstein B, Velculescu VE, Kinzler KW. An integrated genomic analysis of human glioblastoma multiforme. Science. 2008;321:1807–1812. doi: 10.1126/science.1164382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Bleeker FE, Lamba S, Leenstra S, Troost D, Hulsebos T, Vandertop WP, Frattini M, Molinari F, Knowles M, Cerrato A, Rodolfo M, Scarpa A, Felicioni L, Buttitta F, Malatesta S, Marchetti A, Bardelli A. IDH1 mutations at residue p. R132 (IDH1(R132)) occur frequently in highgrade gliomas but not in other solid tumors. Hum mutat. 2009;30:7–11. doi: 10.1002/humu.20937. [DOI] [PubMed] [Google Scholar]
- 146.Yan H, Parsons DW, Jin G, McLendon R, Rasheed BA, Yuan W, Kos I, Batinic-Haberle I, Jones S, Riggins GJ, Friedman H, Friedman A, Reardon D, Herndon J, Kinzler KW, Velculescu VE, Vogelstein B, Bigner DD. IDH1 and IDH2 mutations in gliomas. N Engl J Med. 2009;360:765–773. doi: 10.1056/NEJMoa0808710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Mardis ER, Ding L, Dooling DJ, Larson DE, McLellan MD, Chen K, Koboldt DC, Fulton RS, Delehaunty KD, McGrath SD, Fulton LA, Locke DP, Magrini VJ, Abbott RM, Vickery TL, Reed JS, Robinson JS, Wylie T, Smith SM, Carmichael L, Eldred JM, Harris CC, Walker J, Peck JB, Du F, Dukes AF, Sanderson GE, Brummett AM, Clark E, McMichael JF, Meyer RJ, Schindler JK, Pohl CS, Wallis JW, Shi X, Lin L, Schmidt H, Tang Y, Haipek C, Wiechert ME, Ivy JV, Kalicki J, Elliott G, Ries RE, Payton JE, Westervelt P, Tomasson MH, Watson MA, Baty J, Heath S, Shannon WD, Nagarajan R, Link DC, Walter MJ, Graubert TA, DiPersio JF, Wilson RK, Ley TJ. Recurring mutations found by sequencing an acute myeloid leukemia genome. N Engl J Med. 2009;361:1058–1066. doi: 10.1056/NEJMoa0903840. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Chou WC, Hou HA, Chen CY, Tang JL, Yao M, Tsay W, Ko BS, Wu SJ, Huang SY, Hsu SC, Chen YC, Huang YN, Chang YC, Lee FY, Liu MC, Liu CW, Tseng MH, Huang CF, Tien HF. Distinct clinical and biologic characteristics in adult acute myeloid leukemia bearing the isocitrate dehydrogenase 1 mutation. Blood. 2010;115:2749–2754. doi: 10.1182/blood-2009-11-253070. [DOI] [PubMed] [Google Scholar]
- 149.Rakheja D, Konoplev S, Medeiros LJ, Chen W. IDH mutations in acute myeloid leukemia. Hum Pathol. 2012;43:1541–1551. doi: 10.1016/j.humpath.2012.05.003. [DOI] [PubMed] [Google Scholar]
- 150.Chotirat S, Thongnoppakhun W, Promsuwicha O, Boonthimat C, Auewarakul CU. Molecular alterations of isocitrate dehydrogenase 1 and 2 (IDH1 and IDH2) metabolic genes and additional genetic mutations in newly diagnosed acute myeloid leukemia patients. J Hematol Oncol. 2012;5:5. doi: 10.1186/1756-8722-5-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Duncan CG, Barwick BG, Jin G, Rago C, Kapoor-Vazirani P, Powell DR, Chi JT, Bigner DD, Vertino PM, Yan H. A heterozygous IDH1R132H/WT mutation induces genomewide alterations in DNA methylation. Genome Res. 2012;22:2339–55. doi: 10.1101/gr.132738.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Turcan S, Rohle D, Goenka A, Walsh LA, Fang F, Yilmaz E, Campos C, Fabius AW, Lu C, Ward PS, Thompson CB, Kaufman A, Guryanova O, Levine R, Heguy A, Viale A, Morris LG, Huse JT, Mellinghoff IK, Chan TA. IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature. 2012;483:479–483. doi: 10.1038/nature10866. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Gross S, Cairns RA, Minden MD, Driggers EM, Bittinger MA, Jang HG, Sasaki M, Jin S, Schenkein DP, Su SM, Dang L, Fantin VR, Mak TW. Cancer-associated metabolite 2-hydroxyglutarate accumulates in acute myelogenous leukemia with isocitrate dehydrogenase 1 and 2 mutations. J Exp Med. 2010;207:339–344. doi: 10.1084/jem.20092506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, Fantin VR, Jang HG, Jin S, Keenan MC, Marks KM, Prins RM, Ward PS, Yen KE, Liau LM, Rabinowitz JD, Cantley LC, Thompson CB, Vander Heiden MG, Su SM. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature. 2010;465:966. doi: 10.1038/nature09132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, Fantin VR, Jang HG, Jin S, Keenan MC, Marks KM, Prins RM, Ward PS, Yen KE, Liau LM, Rabinowitz JD, Cantley LC, Thompson CB, Vander Heiden MG, Su SM. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature. 2009;462:739–744. doi: 10.1038/nature08617. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Zhao S, Lin Y, Xu W, Jiang W, Zha Z, Wang P, Yu W, Li Z, Gong L, Peng Y, Ding J, Lei Q, Guan KL, Xiong Y. Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce HIF-1alpha. Science. 2009;324:261–265. doi: 10.1126/science.1170944. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Gross S, Cairns RA, Minden MD, Driggers EM, Bittinger MA, Jang HG, Sasaki M, Jin S, Schenkein DP, Su SM, Dang L, Fantin VR, Mak TW. Cancer-associated metabolite 2-hydroxyglutarate accumulates in acute myelogenous leukemia with isocitrate dehydrogenase 1 and 2 mutations. J Exp Med. 2010;207:339–344. doi: 10.1084/jem.20092506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Benson DN Jr, Bakan CE, Mishra A, Hofmeister CC, Efebera Y, Becknell B, Baiocchi RA, Zhang J, Yu J, Smith MK, Greenfield CN, Porcu P, Devine SM, Rotem-Yehudar R, Lozanski G, Byrd JC, Caligiuri MA. The PD-1/PD-L1 axis modulates the natural killer cell versus multiple myeloma effect: a therapeutic target for CT-011, a novel monoclonal anti-PD-1 antibody. Blood. 2010;116:2286–2294. doi: 10.1182/blood-2010-02-271874. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Abbas S, Lugthart S, Kavelaars FG, Schelen A, Koenders JE, Zeilemaker A, van Putten WJ, Rijneveld AW, Lowenberg B, Valk PJ. Acquired mutations in the genes encoding IDH1 and IDH2 both are recurrent aberrations in acute myeloid leukemia: prevalence and prognostic value. Blood. 2010;116:2122–2126. doi: 10.1182/blood-2009-11-250878. [DOI] [PubMed] [Google Scholar]
- 160.Green CL, Evans CM, Hills RK, Burnett AK, Linch DC, Gale RE. The prognostic significance of IDH1 mutations in younger adult patients with acute myeloid leukemia is dependent on FLT3-ITD status. Blood. 2010;116:2779–2782. doi: 10.1182/blood-2010-02-270926. [DOI] [PubMed] [Google Scholar]
- 161.Green CL, Evans CM, Zhao L, Hills RK, Burnett AK, Linch DC, Gale RE. The prognostic significance of IDH2 mutations in AML depends on the location of the mutation. Blood. 2011;118:409–412. doi: 10.1182/blood-2010-12-322479. [DOI] [PubMed] [Google Scholar]
- 162.Yen KE, Bittinger MA, Su SM, Fantin VR. Cancer-associated IDH mutations: biomarker and therapeutic opportunities. Oncogene. 2010;29:6409–6417. doi: 10.1038/onc.2010.444. [DOI] [PubMed] [Google Scholar]
- 163.Akalin A, Garrett-Bakelman FE, Kormaksson M, Busuttil J, Zhang L, Khrebtukova I, Milne TA, Huang Y, Biswas D, Hess JL, Allis CD, Roeder RG, Valk PJ, Lowenberg B, Delwel R, Fernandez HF, Paietta E, Tallman MS, Schroth GP, Mason CE, Melnick A, Figueroa ME. Base-pair resolution DNA methylation sequencing reveals profoundly divergent epigenetic landscapes in acute myeloid leukemia. PLoS Genet. 2012;8:e1002781. doi: 10.1371/journal.pgen.1002781. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Milne TA, Martin ME, Brock HW, Slany RK, Hess JL. Leukemogenic MLL fusion proteins bind across a broad region of the Hox a9 locus, promoting transcription and multiple histone modifications. Cancer Res. 2005;65:11367–11374. doi: 10.1158/0008-5472.CAN-05-1041. [DOI] [PubMed] [Google Scholar]
- 165.Scharf S, Zech J, Bursen A, Schraets D, Oliver PL, Kliem S, Pfitzner E, Gillert E, Dingermann T, Marschalek R. Transcription linked to recombination: a gene-internal promoter coincides with the recombination hot spot II of the human MLL gene. Oncogene. 2007;26:1361–1371. doi: 10.1038/sj.onc.1209948. [DOI] [PubMed] [Google Scholar]
- 166.Krivtsov AV, Armstrong SA. MLL translocations, histone modifications and leukaemia stem-cell development. Nat Rev Cancer. 2007;7:823–833. doi: 10.1038/nrc2253. [DOI] [PubMed] [Google Scholar]
- 167.Ayton PM, Cleary ML. Molecular mechanisms of leukemogenesis mediated by MLL fusion proteins. Oncogene. 2001;20:5695–5707. doi: 10.1038/sj.onc.1204639. [DOI] [PubMed] [Google Scholar]
- 168.Dohner K, Tobis K, Ulrich R, Frohling S, Benner A, Schlenk RF, Dohner H. Prognostic significance of partial tandem duplications of the MLL gene in adult patients 16 to 60 years old with acute myeloid leukemia and normal cytogenetics: a study of the Acute Myeloid Leukemia Study Group Ulm. J. Clin. Oncol. 2002;20:3254–3261. doi: 10.1200/JCO.2002.09.088. [DOI] [PubMed] [Google Scholar]
- 169.Caligiuri MA, Schichman SA, Strout MP, Mrozek K, Baer MR, Frankel SR, Barcos M, Herzig GP, Croce CM, Bloomfield CD. Molecular rearrangement of the ALL-1 gene in acute myeloid leukemia without cytogenetic evidence of 11q23 chromosomal translocations. Cancer Res. 1994;54:370–373. [PubMed] [Google Scholar]
- 170.Steudel C, Wermke M, Schaich M, Schakel U, Illmer T, Ehninger G, Thiede C. Comparative analysis of MLL partial tandem duplication and FLT3 internal tandem duplication mutations in 956 adult patients with acute myeloid leukemia. Genes Chromosomes Cancer. 2003;37:237–251. doi: 10.1002/gcc.10219. [DOI] [PubMed] [Google Scholar]
- 171.Rege-Cambrin G, Giugliano E, Michaux L, Stul M, Scaravaglio P, Serra A, Saglio G, Hagemeijer A. Trisomy 11 in myeloid malignancies is associated with internal tandem duplication of both MLL and FLT3 genes. Haematologica. 2005;90:262–264. [PubMed] [Google Scholar]
- 172.Libura M, Asnafi V, Tu A, Delabesse E, Tigaud I, Cymbalista F, Bennaceur-Griscelli A, Villarese P, Solbu G, Hagemeijer A, Beldjord K, Hermine O, Macintyre E. FLT3 and MLL intragenic abnormalities in AML reflect a common category of genotoxic stress. Blood. 2003;102:2198–2204. doi: 10.1182/blood-2003-01-0162. [DOI] [PubMed] [Google Scholar]
- 173.Yao Y, Chen P, Diao J, Cheng G, Deng L, Anglin JL, Prasad BV, Song Y. Selective inhibitors of histone methyltransferase DOT1L: design, synthesis, and crystallographic studies. J Am Chem Soc. 2011;133:16746–16749. doi: 10.1021/ja206312b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Daigle SR, Olhava EJ, Therkelsen CA, Majer CR, Sneeringer CJ, Song J, Johnston LD, Scott MP, Smith JJ, Xiao Y, Jin L, Kuntz KW, Chesworth R, Moyer MP, Bernt KM, Tseng JC, Kung AL, Armstrong SA, Copeland RA, Richon VM, Pollock RM. Selective killing of mixed lineage leukemia cells by a potent small-molecule DOT1L inhibitor. Cancer Cell. 2011;20:53–65. doi: 10.1016/j.ccr.2011.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Pollock R, Daigle SR, Therkelsen CA, Basavapathruni A, Jin L, Allain CJ, Klaus CR, Raimondi A, Scott MP, Chesworth R, Moyer MP, Copeland RA, Richon VM, Olhava EJ. Preclinical Characterization of a Potent, Selective Inhibitor of the Protein Methyltransferase DOT1L for Use in the Treatment of MLL-Rearranged Leukemia. ASH Annual Meeting Abstracts. 2012 Poster 2379. [Google Scholar]
- 176.Varambally S, Dhanasekaran SM, Zhou M, Barrette TR, Kumar-Sinha C, Sanda MG, Ghosh D, Pienta KJ, Sewalt RG, Otte AP, Rubin MA, Chinnaiyan AM. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature. 2002;419:624–629. doi: 10.1038/nature01075. [DOI] [PubMed] [Google Scholar]
- 177.Benetatos L, Voulgaris E, Vartholomatos G, Hatzimichael E. Non-coding RNAs and EZH2 interactions in cancer: Long and short tales from the transcriptome. Int J Cancer. 2012 doi: 10.1002/ijc.27859. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
- 178.Ernst T, Chase AJ, Score J, Hidalgo-Curtis CE, Bryant C, Jones AV, Waghorn K, Zoi K, Ross FM, Reiter A, Hochhaus A, Drexler HG, Duncombe A, Cervantes F, Oscier D, Boultwood J, Grand FH, Cross NC. Inactivating mutations of the histone methyltransferase gene EZH2 in myeloid disorders. Nat Genet. 2010;42:722–726. doi: 10.1038/ng.621. [DOI] [PubMed] [Google Scholar]
- 179.Nikoloski G, Langemeijer SM, Kuiper RP, Knops R, Massop M, Tonnissen ER, van der Heijden A, Scheele TN, Vandenberghe P, de Witte T, van der Reijden BA, Jansen JH. Somatic mutations of the histone methyltransferase gene EZH2 in myelodysplastic syndromes. Nat Genet. 2010;42:665–667. doi: 10.1038/ng.620. [DOI] [PubMed] [Google Scholar]
- 180.Zhang J, Ding L, Holmfeldt L, Wu G, Heatley SL, Payne-Turner D, Easton J, Chen X, Wang J, Rusch M, Lu C, Chen SC, Wei L, Collins-Underwood JR, Ma J, Roberts KG, Pounds SB, Ulyanov A, Becksfort J, Gupta P, Huether R, Kriwacki RW, Parker M, McGoldrick DJ, Zhao D, Alford D, Espy S, Bobba KC, Song G, Pei D, Cheng C, Roberts S, Barbato MI, Campana D, Coustan-Smith E, Shurtleff SA, Raimondi SC, Kleppe M, Cools J, Shimano KA, Hermiston ML, Doulatov S, Eppert K, Laurenti E, Notta F, Dick JE, Basso G, Hunger SP, Loh ML, Devidas M, Wood B, Winter S, Dunsmore KP, Fulton RS, Fulton LL, Hong X, Harris CC, Dooling DJ, Ochoa K, Johnson KJ, Obenauer JC, Evans WE, Pui CH, Naeve CW, Ley TJ, Mardis ER, Wilson RK, Downing JR, Mullighan CG. The genetic basis of early T-cell precursor acute lymphoblastic leukaemia. Nature. 2012;481:157–163. doi: 10.1038/nature10725. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Makishima H, Jankowska AM, Tiu RV, Szpurka H, Sugimoto Y, Hu Z, Saunthararajah Y, Guinta K, Keddache MA, Putnam P, Sekeres MA, Moliterno AR, List AF, McDevitt MA, Maciejewski JP. Novel homo- and hemizygous mutations in EZH2 in myeloid malignancies. Leukemia. 2010;24:1799–1804. doi: 10.1038/leu.2010.167. [DOI] [PubMed] [Google Scholar]
- 182.Ernst T, Pflug A, Rinke J, Ernst J, Bierbach U, Beck JF, Hochhaus A, Gruhn B. A somatic EZH2 mutation in childhood acute myeloid leukemia. Leukemia. 2012;26:1701–1703. doi: 10.1038/leu.2012.16. [DOI] [PubMed] [Google Scholar]
- 183.Dolnik A, Engelmann JC, Scharfenberger-Schmeer M, Mauch J, Kelkenberg-Schade S, Haldemann B, Fries T, Kronke J, Kuhn MW, Paschka P, Kayser S, Wolf S, Gaidzik VI, Schlenk RF, Rucker FG, Dohner H, Lottaz C, Dohner K, Bullinger L. Commonly altered genomic regions in acute myeloid leukemia are enriched for somatic mutations involved in chromatin remodeling and splicing. Blood. 2012;120:e83–92. doi: 10.1182/blood-2011-12-401471. [DOI] [PubMed] [Google Scholar]
- 184.Vire E, Brenner C, Deplus R, Blanchon L, Fraga M, Didelot C, Morey L, Van Eynde A, Bernard D, Vanderwinden JM, Bollen M, Esteller M, Di Croce L, de Launoit Y, Fuks F. The Polycomb group protein EZH2 directly controls DNA methylation. Nature. 2006;439:871–874. doi: 10.1038/nature04431. [DOI] [PubMed] [Google Scholar]
- 185.Tan J, Yang X, Zhuang L, Jiang X, Chen W, Lee PL, Karuturi RK, Tan PB, Liu ET, Yu Q. Pharmacologic disruption of Polycomb-repressive complex 2-mediated gene repression selectively induces apoptosis in cancer cells. Genes Dev. 2007;21:1050–1063. doi: 10.1101/gad.1524107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Zhou J, Bi C, Cheong LL, Mahara S, Liu SC, Tay KG, Koh TL, Yu Q, Chng WJ. The histone methyltransferase inhibitor, DZNep, up-regulates TXNIP, increases ROS production, and targets leukemia cells in AML. Blood. 2011;118:2830–2839. doi: 10.1182/blood-2010-07-294827. [DOI] [PubMed] [Google Scholar]
- 187.Fiskus W, Wang Y, Sreekumar A, Buckley KM, Shi H, Jillella A, Ustun C, Rao R, Fernandez P, Chen J, Balusu R, Koul S, Atadja P, Marquez VE, Bhalla KN. Combined epigenetic therapy with the histone methyltransferase EZH2 inhibitor 3-deazaneplanocin A and the histone deacetylase inhibitor panobinostat against human AML cells. Blood. 2009;114:2733–2743. doi: 10.1182/blood-2009-03-213496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Weissmann S, Alpermann T, Grossmann V, Kowarsch A, Nadarajah N, Eder C, Dicker F, Fasan A, Haferlach C, Haferlach T, Kern W, Schnittger S, Kohlmann A. Landscape of TET2 mutations in acute myeloid leukemia. Leukemia. 2012;26:934–942. doi: 10.1038/leu.2011.326. [DOI] [PubMed] [Google Scholar]
- 189.Pratcorona M, Abbas S, Sanders MA, Koenders JE, Kavelaars FG, Erpelinck-Verschueren CA, Zeilemakers A, Lowenberg B, Valk PJ. Acquired mutations in ASXL1 in acute myeloid leukemia: prevalence and prognostic value. Haematologica. 2012;97:388–392. doi: 10.3324/haematol.2011.051532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Grossmann V, Tiacci E, Holmes AB, Kohlmann A, Martelli MP, Kern W, Spanhol-Rosseto A, Klein HU, Dugas M, Schindela S, Trifonov V, Schnittger S, Haferlach C, Bassan R, Wells VA, Spinelli O, Chan J, Rossi R, Baldoni S, De Carolis L, Goetze K, Serve H, Peceny R, Kreuzer KA, Oruzio D, Specchia G, Di Raimondo F, Fabbiano F, Sborgia M, Liso A, Farinelli L, Rambaldi A, Pasqualucci L, Rabadan R, Haferlach T, Falini B. Whole-exome sequencing identifies somatic mutations of BCOR in acute myeloid leukemia with normal karyotype. Blood. 2011;118:6153–6163. doi: 10.1182/blood-2011-07-365320. [DOI] [PubMed] [Google Scholar]
- 191.Dai W. Polo-like kinases, an introduction. Oncogene. 2005;24:214–216. doi: 10.1038/sj.onc.1208270. [DOI] [PubMed] [Google Scholar]
- 192.Archambault V, Carmena M. Polo-like kinase-activating kinases: Aurora A, Aurora B and what else? Cell Cycle. 2012;11:1490–1495. doi: 10.4161/cc.19724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Lens SM, Voest EE, Medema RH. Shared and separate functions of polo-like kinases and aurora kinases in cancer. Nat Rev Cancer. 2010;10:825–841. doi: 10.1038/nrc2964. [DOI] [PubMed] [Google Scholar]
- 194.Tsou MF, Wang WJ, George KA, Uryu K, Stearns T, Jallepalli PV. Polo kinase and separase regulate the mitotic licensing of centriole duplication in human cells. Dev Cell. 2009;17:344–354. doi: 10.1016/j.devcel.2009.07.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Trenz K, Errico A, Costanzo V. Plx1 is required for chromosomal DNA replication under stressful conditions. EMBO J. 2008;27:876–885. doi: 10.1038/emboj.2008.29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Tamura Y, Simizu S, Muroi M, Takagi S, Kawatani M, Watanabe N, Osada H. Pololike kinase 1 phosphorylates and regulates Bcl-x(L) during pironetin-induced apoptosis. Oncogene. 2009;28:107–116. doi: 10.1038/onc.2008.368. [DOI] [PubMed] [Google Scholar]
- 197.Takai N, Hamanaka R, Yoshimatsu J, Miyakawa I. Polo-like kinases (Plks) and cancer. Oncogene. 2005;24:287–291. doi: 10.1038/sj.onc.1208272. [DOI] [PubMed] [Google Scholar]
- 198.Renner AG, Dos Santos C, Recher C, Bailly C, Creancier L, Kruczynski A, Payrastre B, Manenti S. Polo-like kinase 1 is overexpressed in acute myeloid leukemia and its inhibition preferentially targets the proliferation of leukemic cells. Blood. 2009;114:659–662. doi: 10.1182/blood-2008-12-195867. [DOI] [PubMed] [Google Scholar]
- 199.Liu X, Erikson RL. Polo-like kinase (Plk)1 depletion induces apoptosis in cancer cells. Proc Natl Acad Sci U S A. 2003;100:5789–5794. doi: 10.1073/pnas.1031523100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Hikichi Y, Honda K, Hikami K, Miyashita H, Kaieda I, Murai S, Uchiyama N, Hasegawa M, Kawamoto T, Sato T, Ichikawa T, Cao S, Nie Z, Zhang L, Yang J, Kuida K, Kupperman E. TAK-960, a novel, orally available, selective inhibitor of polo-like kinase 1, shows broad-spectrum preclinical antitumor activity in multiple dosing regimens. Mol Cancer Ther. 2012;11:700–709. doi: 10.1158/1535-7163.MCT-11-0762. [DOI] [PubMed] [Google Scholar]
- 201.Didier C, Cavelier C, Quaranta M, Demur C, Ducommun B. Evaluation of Polo-like Kinase 1 inhibition on the G2/M checkpoint in Acute Myelocytic Leukaemia. Eur J Pharmacol. 2008;591:102–105. doi: 10.1016/j.ejphar.2008.06.062. [DOI] [PubMed] [Google Scholar]
- 202.Shi JQ, Lasky K, Shinde V, Stringer B, Qian MG, Liao D, Liu R, Driscoll D, Nestor MT, Amidon BS, Rao Y, Duffey MO, Manfredi MG, Vos TJ, N DA, Hyer ML. MLN0905, a small-molecule plk1 inhibitor, induces antitumor responses in human models of diffuse large B-cell lymphoma. Mol Cancer Ther. 2012;11:2045–2053. doi: 10.1158/1535-7163.MCT-11-1036. [DOI] [PubMed] [Google Scholar]
- 203.Valsasina B, Beria I, Alli C, Alzani R, Avanzi N, Ballinari D, Cappella P, Caruso M, Casolaro A, Ciavolella A, Cucchi U, De Ponti A, Felder E, Fiorentini F, Galvani A, Gianellini LM, Giorgini ML, Isacchi A, Lansen J, Pesenti E, Rizzi S, Rocchetti M, Sola F, Moll J. NMS-P937, an orally available, specific small-molecule pololike kinase 1 inhibitor with antitumor activity in solid and hematologic malignancies. Mol Cancer Ther. 2012;11:1006–1016. doi: 10.1158/1535-7163.MCT-11-0765. [DOI] [PubMed] [Google Scholar]
- 204.Garuti L, Roberti M, Bottegoni G. Polo-like kinases inhibitors. Curr Med Chem. 2012;19:3937–3948. doi: 10.2174/092986712802002455. [DOI] [PubMed] [Google Scholar]
- 205.Tsykunova G, Reikvam H, Ahmed AB, Nepstad I, Gjertsen BT, Bruserud Ø. Targeting of polo-like kinases and their cross talk with Aurora kinases--possible therapeutic strategies in human acute myeloid leukemia? Expert Opin Investig Drugs. 2012;21:587–603. doi: 10.1517/13543784.2012.668525. [DOI] [PubMed] [Google Scholar]
- 206.Ikezoe T, Yang J, Nishioka C, Takezaki Y, Tasaka T, Togitani K, Koeffler HP, Yokoyama A. A novel treatment strategy targeting polo-like kinase 1 in hematological malignancies. Leukemia. 2009;23:1564–1576. doi: 10.1038/leu.2009.94. [DOI] [PubMed] [Google Scholar]
- 207.Muller-Tidow C, Bug G, Schlenk R, Lubbert M, Kramer A, Krauter J, Nachbaur D, Valent P, Taube T, Munzert G, Lee K-H, Dohner H. Phase I/II Study of BI 2536, An Intravenous Polo-Like Kinase-1 (Plk-1) Inhibitor, in Elderly Patients with Relapsed or Refractory Acute Myeloid Leukemia (AML): First Results of a Multi-Center Trial. ASH Annual Meeting Abstracts. 2008;112:2973. [Google Scholar]
- 208.Bug G, Muller-Tidow C, Schlenk RF, Kramer A, Lubbert M, Krug U, Voss F, Taube T, Fritsch H, Garin-Chesa P, Ottmann OG, Dohner H. Phase I/II Study of Volasertib (BI 6727), An Intravenous Polo-Like Kinase (Plk) Inhibitor, in Patients with Acute Myeloid Leukemia (AML): Updated Results of the Dose Finding Phase I Part for Volasertib in Combination with Low-Dose Cytarabine (LD-Ara-C) and As Monotherapy in Relapsed/Refractory AML. ASH Annual Meeting Abstracts. 2011;118:1549. [Google Scholar]
- 209.Bug G, Schlenk RF, Muller-Tidow C, Lubbert M, Kramer A, Fleischer F, Taube T, Ottmann OG, Doehner H. Phase I/II Study of BI 6727 (volasertib), An Intravenous Polo-Like Kinase-1 (Plk1) Inhibitor, In Patients with Acute Myeloid Leukemia (AML): Results of the Dose Finding for BI 6727 In Combination with Low-Dose Cytarabine. ASH Annual Meeting Abstracts. 2010;116:3316. [Google Scholar]
- 210.Simon R, Maitournam A. Evaluating the efficiency of targeted designs for randomized clinical trials. Clin Cancer Res. 2004;10:6759–6763. doi: 10.1158/1078-0432.CCR-04-0496. [DOI] [PubMed] [Google Scholar]
- 211.Knapper S, White P, Levis MJ, Hills RK, Russell NH, Burnett A. The Efficacy of the FLT3 Inhibitor Lestaurtinib in AML Depends on Adequate Plasma Inhibitory Activity (PIA), and Is Unaffected by Rising FLT Ligand Levels: An Update of the NCRI AML15 & 17 Trials. ASH Annual Meeting Abstracts. 2011;118:421. [Google Scholar]
- 212.Watt TC, Cooper T. Sorafenib as treatment for relapsed or refractory pediatric acute myelogenous leukemia. Pediatr Blood Cancer. 2012;59:756–757. doi: 10.1002/pbc.23394. [DOI] [PubMed] [Google Scholar]
- 213.Man CH, Fung TK, Ho C, Han HH, Chow HC, Ma AC, Choi WW, Lok S, Cheung AM, Eaves C, Kwong YL, Leung AY. Sorafenib treatment of FLT3-ITD(+) acute myeloid leukemia: favorable initial outcome and mechanisms of subsequent nonresponsiveness associated with the emergence of a D835 mutation. Blood. 2012;119:5133–5143. doi: 10.1182/blood-2011-06-363960. [DOI] [PubMed] [Google Scholar]
- 214.Macdonald DA, Assouline SE, Brandwein J, Kamel-Reid S, Eisenhauer EA, Couban S, Caplan S, Foo A, Walsh W, Leber B. A Phase I/II Study of Sorafenib (Bay 43-9006) in Combination with Low Dose Cytarabine (LDAC) in Elderly Patients with AML or High-Risk MDS from The NCIC Clinical Trials Group: Trial IND. 186. Leuk Lymphoma. 2012 doi: 10.3109/10428194.2012.737917. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
- 215.Borthakur G, Kantarjian H, Ravandi F, Zhang W, Konopleva M, Wright JJ, Faderl S, Verstovsek S, Mathews S, Andreeff M, Cortes JE. Phase I study of sorafenib in patients with refractory or relapsed acute leukemias. Haematologica. 2011;96:62–68. doi: 10.3324/haematol.2010.030452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Rudolph D, Steegmaier M, Hoffmann M, Grauert M, Baum A, Quant J, Haslinger C, Garin-Chesa P, Adolf GR. BI 6727, a Polo-like kinase inhibitor with improved pharmacokinetic profile and broad antitumor activity. Clin Cancer Res. 2009;15:3094–3102. doi: 10.1158/1078-0432.CCR-08-2445. [DOI] [PubMed] [Google Scholar]
- 217.Pemmaraju N, Kantarjian HM, O’():Brien S, Kadia TM, Cortes JE, Borthakur G, Faderl S, Garcia-Manero G, Estrov Z, Ravandi F, Quintas-Cardama A, Jabbour EJ, Koller CA, Dellasala SE, Pierce SA, M Burton E, Verstovsek S. Results of A Phase I Study of Ruxolitinib in Patients (pts) with Relapsed/Refractory Acute Leukemia. ASH Annual Meeting Abstracts. 2012 Abstract 3617. [Google Scholar]
- 218.Shabbir M, Stuart R. Lestaurtinib, a multitargeted tyrosine kinase inhibitor: from bench to bedside. Expert Opin Investig Drugs. 2010;19:427–436. doi: 10.1517/13543781003598862. [DOI] [PubMed] [Google Scholar]