GET THE APP

The Current Molecular Site of the Myeloproliferative Neoplasms -
Journal of Leukemia

Journal of Leukemia
Open Access

ISSN: 2329-6917

+44 1300 500008

Review Article - (2013) Volume 1, Issue 1

The Current Molecular Site of the Myeloproliferative Neoplasms - TET-á-tête with the Jackpot but no LNK so far to Resolve Complexity

Oliver Bock*
Institute of Psychiatry, Social Psychiatry and Psychotherapy, Hannover Medical School, Germany
*Corresponding Author: Oliver Bock, Department of Psychiatry, Social Psychiatry and Psychotherapy, Hannover Medical School, Carl-Neuberg-Strasse 1, 30625 Hannover, Germany, Tel: 0049-511-532-6748, Fax: 0049-511-532-6648 Email:

Abstract

The myeloproliferative neoplasms essential thrombocythemia (ET), polycythemia vera (PV) and primary myelofibrosis (PMF) are clonal haematological stem cell disorders with different clinical courses and prognosis. The predominant feature in these entities is the overproduction of at least one cellular lineage in the bone marrow. Peripheral blood cells are terminally differentiated and functional but may affect e.g. blood viscosity with a high hematocrit in PV leading to the risk for thrombotic events or bleeding. PMF, preceded by a hypercellular pre-fibrotic stage, has the highest risk to develop manifest bone marrow fibrosis with consecutive bone marrow failure. The risk for a blast crisis and transformation into secondary acute leukemia is also particularly high in PMF and PV but rather uncommon in ET. Even though distinguishable from each other by clinical and histopathological criteria ET, PV and PMF may show overlaps, i.e. in early stages. Moreover, reactive hypercellular states in the bone marrow may mimic a myeloproliferative neoplasm (MPN) making definite molecular markers very useful. ET, PV and PMF currently share a number of molecular defects which are detectable by accurate technologies and, in a diagnostic sense, subsequently allow a clear-cut discriminationfrom reactive states. Some of these molecular defects are frequent in the stage of secondary acute leukemia; others were proposed to be potential predictors for a progression of the disease course. At best, the protagonist of molecular defects in Ph- MPN, the Janus Kinase 2 with its V617F mutation, became a therapeutic target by using small molecule inhibitors. This review focuses on the hitherto discovered molecular defects in ET, PV and PMF in chronic phase and disease progression, highlights some of the current and upcoming therapies and proposes a disease model.

Keywords: Myeloproliferative neoplasms; Molecular pathobiology; Molecularly targeted therapies

Introduction

The groundbreaking discovery of the somatic point mutation in the Janus kinase 2, JAK2 (V617F), in the so-called classical Philadelphia chromosome negative (Ph-) chronic myeloproliferative disorders (MPD) as well as a molecular defect in the thrombopoietin receptor MPL (W515L/K) justified a revision of the WHO classification published in 2008 [1]. The WHO classification replaced the term Ph- MPD by “myeloproliferative neoplasm” (MPN), a term that was introduced to emphasize the neoplastic nature of these disease entities. The revised classification also acknowledged the presence of molecular markers such as JAK2 (V617F) and proposed new standards for the diagnostic algorithm, i.e. introduction of molecular diagnostics. The formerly called chronic idiopathic myelofibrosis (CIMF) was renamed to “primary myelofibrosis” to allow a better differentiation from secondary myelofibrosis demonstrable in other MPN or myeloid malignancies.

The current WHO classification gathers the four so-called classical MPN, the Philadelphia-chromosome positive (Ph+) t(9;22) (q34;q11), BCR-ABL positive chronic myeloid leukemia with the Ph- MPN essential thrombocythemia (ET), polycythemia vera (PV) and primary myelofibrosis (PMF) along with more rare entities such as chronic neutrophilic leukemia (CNL), systemic mastocytosis (SM), hypereosinophilic syndrome, the chronic eosinophilic leukemia not other specified as well as MPN unclassifiable [1].

Before the discovery of the gain-of-function mutation V617F in JAK2 in a considerable number of patients with Ph- MPN was published almost 8 years ago [2-5], the knowledge on genetic defects was rather sparse over decades. Some of the cases showed demonstrable chromosomal deletions such as del(20q) or del(13q) in PMF [6] or a loss of heterozygosity (LOH) at 9p in patients with polycythemia vera (PV) [7]. The latter finding, however, was probably the most important cornerstone for the subsequent discovery of JAK2 aberrations. Several findings like the erythroid colony (EEC) formation in PV in a cytokineindependent manner [8], the demonstration of clonality in MPN [9], cytokine hypersensitivity [10], the aberrant processing and expression of growth factor receptors such as MPL (Myeloproliferative leukemia virus oncogene) in PV and ET [11,12] and studies on factors involved in stroma pathology such as transforming growth factor β-1 (TGFβ- 1) basic fibroblast growth factor (bFGF), vascular endothelial growth factor (VEGF) and aberrant proplatelet formation [13-17] contributed to the still leaky understanding of the nature of Ph- MPN. Moreover, some studies introduced new findings as diagnostic markers such as reduced MPL expression in certain MPD subtypes or the Polycythaemia vera rubra receptor-1 (PRV-1) mRNA expression in granulocytes from peripheral blood [18,19].

The hitherto discovered molecular defects in Ph- MPN and associated myeloid disorders can functionally be grouped in those 1) associated with constitutive activation of cytokine signalling driven by aberrant factors such as JAK2, MPL, LNK, CBL, NRAS, NF1, 2) those affecting epigenetics and regulation of transcription driven by alterations in TET2, ASXL1, EZH2 and other PRC2 members, DNMTA3A 3) those affecting the splicing machinery like SRSF2 and SF3B1, 4) those associated with disease progression and transformation into secondary acute leukemia like IKZF1, RUNX1, TP53, IDH1/2, DNMTA3A.

However, the full spectrum of intracellular alterations initiated by these molecular defects remains to be investigated. For example a protein with a new structure and folding due to a change in the amino acid sequence such as JAK2 (V617F) may not only be constitutively activated but may provide aberrant interactions with other factors.

In the following section, the currently known molecular defects will be listed and discussed in an alphabetical order and not based on the cellular effects proposed elsewhere. Table 1 summarizes this section and allows a quick overview. At the end, the current state of therapeutic options through molecular targeting and a disease model for Ph- MPN will be introduced.

Gene symbol Gene Cytogenetic
location
Molecular defect* Frequency in Ph-
 MPN (%)*
ASXL1 Additional sex comb like 1 gene 20q11.21 Truncating mutations mainly in
exon 12
~36% PMF, post-PV-MF, post-ET-MF
≤ 6% chronic PV/ET
CBL Casitas B-lineage lymphoma protooncogene 11q23.3 Missense mutations (exon 8 and 9), nonsense and splicing alterations (also intron regions) ~6% PMF
Rare in PV, ET
CUX1(CUTL1) Cut-like 1 (CCAAT displacement protein) 7q22.1 Deletion, aUPD <1% PMF, PV, ET
~5% in post-MPN sAML
DNMT3A DNA Methyltransferase 3A 2p23.3 Frameshift mutations, non-sense mutations, deletions ~10% PMF
~10% post-MPN sAML
~7% PV, ~3% ET
IDH1 Isocitrate dehydrogenase 1 2q33.3    
Point mutations
~4% PMF, ~2% PV, ~1% ET
>20% in post-MPN sAML
IDH2 Isocitrate dehydrogenase 2 15q26.1
IKZF1 Ikaros family zinc finger 1 7p12.2 Deletion <1% in chronic MPN, ~20% in post-MPN sAML
JAK2 Janus kinase 2 9p24.1 aUPD, point mutation in exon 14 (V617F) >95% PV, ~50% PMF,
~50% ET
~50% in post-MPN sAML
JAK2 Janus kinase 2 9p24.1 Mutations in exon 12 (K539L and others) ~3% PV, rare in PMF and ET
LNK(SH2B3) Lymphocyte adaptor protein
(SH2B adaptor protein 3)
12q24.12 Mutation, Loss-of-function ~5% PMF and ET
~13% in post-MPN sAML
Must be considered to be detectable in JAK2 (wild-type) erythrocytosis
MPL Myeloproliferative leukemia virus oncogene 1p34.2 Point mutations in the juxtamembrane domain (W515K/L/A) & transmembrane domain (S505N) ~10% PMF
~3-5% ET
NF1 Neurofibromin 1 17q11.2 Deletion ~1% PMF
NRAS Neuroblastoma RAS viral oncogene homol. 1p13.2 Gain-of-function Up to ~13% in post-MPN sAML
PRC2 Polycomb repressive complex 2 (without RBBP4/-7, PHF1/-19, MTF2)
EZH2 Enhancer of Zeste, drosophila, homolog 2 7q36.1 LOH, deletions ~13% PMF
~3% PV
None in ET ?!
SUZ12 Suppressor of Zeste 12, drosophila, homolog of 17q11.2 Mutations, deletions ~1% MPN
(and MDS)
EED Embryonic ectoderm development protein, mouse, homolog of 11q14.2 Mutations, deletions ~1% MPN
(and MDS)
JARID2 Jumonji, AT-rich interactive domain 2 6p22.3 Deletions ~2% MPN
(and MDS)
AEBP2 Adipocyte enhancer binding protein 2 12p12.3 Deletions Rare in chronic MPN
RUNX1 Runt-related transcription factor 1 21q22.12 Deletion, non-sense alterations <1% in chronic MPN
27-37% in sAML
SF3B1 Splicing factor 3B, subunit 1 2q33.1 Mutations in exon 12 to 15 (K700E) ~6% PMF
SRSF2 Splicing factor, serine/arginine -rich 17q25.1 Point mutation or insertion/deletion
(P95H/L/R others)
~17% PMF
~19% sAML from chronic MPN
SOCS1
SOCS2
SOCS3
Suppressor of cytokine signalling 16p13.13
12q
17q25.3
Altered methylation status SOCS1: ~15%
SOCS2: ~28%
SOCS3: ~28%
in chronic MPN
TET2 TET oncogene family, member 2 4q24 Point/frameshift mutations (mainly exon 3 and 11), rare splicing alterations and deletions, missense mutations ~17% PMF
~16% PV
~5% ET
TP53 Tumor protein p53 17p13.1 Loss of function Rare in chronic MPN
~27% in sAML

*=For References see Text section.

Table 1: Compilation of genetic lesions in Ph- MPN and secondary AML.

Genetic Lesions – Compilation in Table 1

Additional sex comb like 1 gene (ASXL1)

ASXL1 is the first of three members of the ASXL family (ASXL1-3). ASXL1 is involved in the regulation of retinoic acid receptor mediated transcription and important for repression of HOX genes. Its function in mammals and especially in hematopoiesis is yet not fully understood. Besides a mild decrease in progenitor cells, mice models knocked out for ASXL1 showed no dramatic effects on hematopoiesis [20].

Somatic mutations involving the ASXL1 gene mainly affect exon 12 and 13 and are demonstrable in chronic MPN and at higher frequency in those showing disease progression. Up to 36% of PMF cases and post-PV/-ET showed mutations in the ASXL1 gene whereas chronic PV and ET cases rarely showed this molecular defect (<6%) [21,22]. It was shown that ASXL1 mutations in MPN can occur before appearance of other defects such as mutated JAK2 and TET2 [23]. ASXL1 mutations are also demonstrable in subtypes of the myelodyplastic syndrome (MDS), post-MDS AML, the chronic myelomonocytic leukemia (CMML) and also de novo AML [24,25].

Casitas B-lineage lymphoma proto-oncogene (CBL)

Two studies were first in demonstration of the frequent occurrence of molecular defects in the CBL gene apart from already known fusion genes [26,27]. Acquired uniparental disomy (aUPD) generally leads to loss-of-heterozygosity and can induce either the loss-of-function of tumor suppressors or gain-of-function of proto-oncogenes. Interestingly, aUPD could be demonstrated in the gene of tumor suppressor CBL which induces a gain-of-function leading to inhibition of the wild-type CBL and prolonged action of tyrosine kinases.

The proto-oncogene CBL is a negative regulator of several receptor tyrosine kinase signaling pathways and an adaptor protein in tyrosinephosphorylation dependent signalling [28].

Ubiquitination of receptor protein-tyrosine kinases (PTKs) terminates signaling by marking active receptors for degradation. CBL is an adaptor protein for receptor PTKs. It positively regulates receptor PTK ubiquitination. Ubiquitin-protein ligases, also known as E3s, are the components of ubiquitination pathways that recognize target substrates and promote their ligation to ubiquitin. A previous study determined that the CBL protein acts as an E3 that can recognize tyrosine-phosphorylated substrates, such as the activated plateletderived growth factor receptor [29]. It was concluded that CBL is an ubiquitin-protein ligase and thus provide a distinct mechanism for substrate targeting in the ubiquitin system.

CBL mutations were often associated with aUPD at 11q and shows involvement of CBL exon 8 and 9 but also of introns. PMF cases were affected by up to 6%, PV and ET were rarely affected by CBL mutations. Notably, an underlying molecular defect in CBL seems to correlate with a shorter overall survival and shorter progression-free survival in patients with (n=19) or without (n=87) CBL mutation [26]. However, all in all CBL mutations are rare genetic events in chronic MPN [22].

Cut-like 1 (CUTL1/CUX1)

CUX1 was initially reported to be a repressor of developmentally regulated genes because of the ability to bind and modify certain histones. In a human acute leukemia cell line (HL-60) a repressor function of the gp91-phox gene could be demonstrated in vitro [30]. Later on, it was proposed that CUX1 might exhibit general tumor suppressor functions. For instance, CUX1 showed LOH and reduced mRNA levels in tumors of the uterus [31]. Meanwhile, CUX1 transcription factors were demonstrated to regulate a number of genes and microRNA [32]. Data on genetic aberrations in the accelerated phase or transformation to secondary acute leukemia in MPN revealed that CUX1 on chromosome 7q was affected and involved in disease progression [33,34]. A deletion on chromosome 7q was demonstrable in 5, 17% of patients with disease progression (3/58) but only in 0,62% of patients in chronic phase (2/321) [33].

DNA Methyltransferase 3A (DNMT3A)

DNA methyltransferases transfer methyl groups to the cytosine at CpG islands. Whereas DNMT3 (A+B) share similar functions DNMT1 is believed to be the keeper of the methylation status [35].

DNMT3A mutations were first described in de novo AML in 62 out of 281 patients (~22%) [36]. Mutations comprise frameshift, nonsense, splice-site as well as deletions. DNMT3A mutations were demonstrable more often in the intermediate-risk cytogenetic profile but did not occur in the group of patients with a favourable cytogenetic profile. Of note, patients with DNMT3A mutations had a significantly shorter median overall survival [36].

In MPN, mutation rates are more frequent in PMF and cases with disease progression (~10%) than in PV (~7%) and ET (3%) [37,38]. DNMT3A mutations often showed co-incidence with other defects such as in ASXL1, IDH1/2, JAK2 and TET2. Because the definite function of mutated DNMT3A is yet unclear and also the mix of defects might be predictable for the molecular fate further studies will elucidate if DNMT3A acts as an oncogene or tumor-suppressor in MPN.

Enhancer of Zeste, homolog 2 (EZH2)/Polycomb repressive complex 2 (PRC2)

EZH2 together with EZH1 is part of the polycomb repressor complex 2 (PRC2). PRC2 itself is involved in many cellular processes such as proliferation, differentiation, senescence and cell fate [39]. PRC2 methylates certain histones thereby regulating gene expression. Apart from EZH1 and EZH2 PRC2 is composed of other factors: SUZ12, RBBP4, RBBP7, EED, AEBP2, JARID2, PHF1, PHF19, MTF2 [40].

EZH2 as part of PRC2 is a methyltransferase with high activity in contrast to EZH1 which has a rather low activity [41]. In myeloid disorders EZH2 mutations apparently lead to inactivation or aberrant function of the protein. EZH2 mutations can be found in ~3% of PV and ~13% of PMF but not in ET cases [42]. Mutations of EZH2 might be associated with a poor clinical course especially in PMF where EZH2 mutations together with co-existing mutations in ASXL1, JAK2 (V617F), CBL or TET2 define high-risk patients according to the International Prognostic Score System (IPSS) [43].

EZH2 together with SUZ12 and EED builds the core complex of PRC2 to enable full function. These and the other PRC2 factors EZH1, SUZ12, EED and JARID2 were investigated for aberrations as well. Apart from EZH2, one study found that genetic aberrations in other PRC2 members are not common after screening a large number of myeloid malignancies [44]. Another study found that SUZ12 and EED are affected in ~1% of cases and JARID2 in ~2% of cases with MPN or MDS [45]. JARID2 and AEBP2 like the other PRC2 factors appear to be affected later in the disease course of MPN and MDS and might be associated with transformation in secondary AML [46].

Ikaros Family Zinc Finger 1 (IKZF1)

This gene encodes for a transcription factor important for the sufficient development of the lymphoid lineage. IKFZ1 is involved in gene regulation through chromatin remodelling [47]. Animal models deficient in IKZF1 action showed defects of B cells, T cells and Natural killer cells or, in case of the heterozygosity for the dominant-negative allele, an acute lymphoblastic leukemia (ALL). Of note, the dominantnegative isoform Ik-6 was shown to be substantially involved in the development of B-cell malignancies [48]. Ph+ BCR-ABL positive ALL frequently showed deletions of the IKZF1 gene [49]. Secondary AML evolved from MPN showed an IKZF1 deletion in up to 20% while chronic MPN rarely showed this defect [50]. Deletions of IKZF1 in this study were proposed to appear later in the disease course and which is related to severe disease progression.

Isocitrate Dehydrogenase 1 and 2 (IDH1/2)

IDH1 and IDH2 are catalytic enzymes involved in conversion of isocitrate to alpha-ketoglutarate by concomitant production of NADPH [51].

Initially, Parsons et al. describe the occurrence of somatic point mutations in IDH1 (R132H) and IDH2 (R172K) as exclusive for glioblastoma [52]. Later on, by using parallel DNA sequencing a large study found these molecular defects also in de novo AML [53] and in chronic MPN evolving to leukemic transformation [54]. The occurrence of mutant IDH is highest in secondary AML developed in chronic MPN (~21%) and rare in PMF (~4%), PV (~2%) and ET (~1%). In our study from 263 MPN and 43 MDS samples none showed the expected R132H mutation in IDH1 or the R172K in IDH2, respectively [55]. These negative findings were supported also by others [22]. However, in de novo AML we found up to 7% of cases showing the IDH1 (R132H).

Mutant IDH was shown to be a gain-of-function defect which abnormally uses alpha-ketoglutarate instead of isocitrate as a substrate. As a consequence, alpha-ketoglutarate will be converted to 2-hydroxyglutarate which accumulates in the affected cell and also in the sera where it could serves as a clinical biomarker [56]. A knock-in mice model more precisely revealed that IDH1 (R132H) increased the number of hematopoietic progenitors, provoked massive splenomegaly and, in the end, was responsible for a severe anaemia [57]. In addition, the myeloid lineage in this model showed aberrant hypermethylation of histones thereby paralleling the findings from human AML samples [53]. Interestingly, a leukemogenic consequence of the above mentioned conversion from alpha-ketoglutarate to 2-hydroxyglutarate through mutant IDH is a decreased TET2 activity. TET2 normally acts as a 2-oxoglutarate/Iron (Fe2+)-dependent dioxygenase which converts 5-methylcytosine to 5-hydroxymethylcytosine. TET2 therefore is involved in the demethylation process of DNA and its inhibition could substantially contribute to the leukemogenic potential of an affected cell [58,59].

Janus Kinase 2/JAK2 (V617F)

Janus kinases belong to the non-receptor tyrosine kinases with JAK1, JAK2, JAK3 and TYK2 being the currently known family members. JAKs like other family of kinases play important roles in signal transduction pathways that govern cellular proliferation, differentiation, survival and apoptosis [60]. Many hematopoietic growth factors and cytokines such as erythropoietin (EPO), thrombopoietin (TPO), G-CSF, GM-CSF, interferon-γ, interleukins-2/-3/-5/-6/-11/-12/-13 but also the growth hormone and insulin recruit JAK2 after binding to their receptors [60]. Some of these factors such as TPO recruit different non-receptor tyrosine kinases which activates JAK2 but also TYK2.

The relatively large proteins of the JAK family (120-140 kDa) possess 7 homology domains (JH1-7). JH1 is the catalytically active domain with tyrosine kinase activity whereas the adjacent neighbour domain JH2 is a pseudokinase with auto-inhibitory function [61].

Following the intriguing finding that in patients with PV ~30% showed aUPD of chromosome 9p [7], 4 groups independently published the discovery of a gain-of-function mutation in JAK2 [2-5]. Among MPN, PV showed the highest JAK2 (V617F) frequency (~95%) followed by PMF (~50%) and ET (~50%) [62-67].

The molecular architecture is a hotspot in exon 14 where the wildtype G is exchanged by a mutant T with replacement of the amino acid valine by phenylalanine at position 617 (V617F) The bulky amino acid phenylalanine changes the conformation of the protein thereby no longer allowing the interaction of the autoinhibitory JH2 pseudokinase domain with the catalytic domain JH1. This disables control of its catalytic activity and leads to a constitutively activated kinase. Subsequently, downstream targets of this kinase will become activated in an affected cell [68]. The effects of mutated JAK2 (V617F) showed a considerable diversity with respect to cellular lineages and MPN subtypes affected by this mutation. In vitro the JAK2 (V617F) clone is autonomous and highly proliferative even in the absence of growth factors. This is demonstrable by higher numbers of endogenous erythroid colonies (EEC) in all subtypes harbouring the JAK2 (V617F). Interestingly, cellular lineages are additionally hypersensitive to growth factors such as EPO, G-CSF or interleukins. In mouse models the mutation is powerful enough to mediate a PV-like phenotype with development of myelofibrosis [4,69,70]. The first mice models showed strain-specific differences. Whereas both the Balb/C and the C57Bl/6 mice showed an increase in hemoglobin and hematocrit, the number of leukocytes as well as the degree of myelofibrosis was strikingly higher in the Balb/C mice [69]. It became clear that the dosage of the mutant JAK2 gene was responsible for the phenotypic degree of the developing MPN. A heterozygous JAK2 (V617F) provokes a more chronic PV-like phenotype whereas homozygous expression led to a more severe course with the development of myelofibrosis [71]. This was in contrast to more previous in vivo findings showing no stringent correlation of the JAK2 status with the clinical course and degree of fibrosis [72,73]. However, JAK2 (V617F) - mutated ET patients show PV-like features, e.g. elevated hemoglobin and hematocrit levels and PV with a homozygous mutation have been suggested to have a higher risk for development of myelofibrosis [63-65,74].

An intriguing study unmasked the cell cycle regulator CDC25A as one cellular target of mutated JAK2 (V617) with high impact on the proliferative capacity of affected cells in vitro and in vivo [75]. CDC25A was shown to be regulated by STAT5 on the translational level and its inhibition in JAK2 (V617F) affected cells markedly reduced proliferation and extent of clonality. Therefore, CDC25A generally represents a potential target for new therapeutic strategies.

Though the JAK2 (V617F) was a breakthrough in the field, it became clear that this molecular defect is not the definite initiator in MPN pathogenesis. This was demonstrated by X-chromosome inactivation patterns (XCIP) in female ET patients harbouring the JAK2 (V617F) mutation [76-78]. As shown by XCIP, the proportion of cells with JAK2 (V617F) was smaller than the total number of clonal cells in a given patient [76]. In addition to classical XCIP approaches, one study elegantly investigated the portion of cells affected by the JAK2 (V617F) mutation in comparison to the total number of clonal cells in a given patient by using the deletion 20q (del20q) as an autosomal, XCIP-independent clonality marker [77]. A clear discrepancy between a higher proportion of cells carrying the del20q and a lower percentage of cells carrying the JAK2 (V617F) could be demonstrated. These data indicate that the hematopoietic stem cell (HSC) must have acquired at least one molecular defect which precedes the occurence of JAK2 (V617F). Subsequent discoveries of newer molecular defects indeed showed that ASXL1 belong to the events which can occur before JAK2 (V617F) was acquired [23].

A very important prerequisite for the acquisition of the JAK2 (V617F) mutation seems to be a certain combination of singlenucleotide polymorphisms (SNPs) on one allele of the JAK2 gene which confers susceptibility for the V617F mutation [79,80]. Patients with PV, ET and myelofibrosis (n=311) all had the so-called JAK2 haplotype 46/1 and in most of the cases the V617F mutation arose on this allele [79]. This genetic constitution as shown by a specific SNP profile probably justifies the term “predisposition” for development of a Ph- MPN (Figure 1) [81-83].

leukemia-progenitor-cells

Figure 1: Proposed steps in the development of Ph- MPN: 1. A genetically normal hematopoietic stem cell (HSC) and/or early progenitor cells might be predisposed for a genetic lesion which is currently unknown. Predisposition as shown for the JAK2 haplotype 46/1 may exist for other genes but are yet undiscovered 2. Cells affected by a genetic defect may undergo apoptosis. An affected cell that survives is in continuous competition with the larger pool of normal HSC. A small pre-malignant clone is clinically unapparent. 3. The latency until clonal hematopoiesis in Ph- MPN become clinically apparent is unknown and may span decades. During this time single cells may acquire additional molecular hits such as ASXL1, TET2, JAK2 and/or epigenetic aberrations which lastly drive the clone to clinical appearence. 4. The composition of molecular defects (“hit signature”) and the occurrence in progenitors of particular cellular lineages definitely initiate the disease and form the type of Ph- MPN. 5. Occurrence of additional genetic lesions associated with sAML is rarely demonstrable before disease progressiontakes place. Most of the genes are not specific for sAML evolved from Ph- MPN. Secondary AML may or may not harbour the JAK2 (V617F) mutation; up to 50% of sAML evolved in a chronic Ph- MPN with a pre-existent JAK2 (V617F) show the JAK2 (wild-type) in the leukemic cells [81]. 6. The bone marrow stroma including cell populations forming the connective tissue (e.g. fibroblasts, adipocytes, endothelial cells, osteoblasts, osteoklasts) and the so-called vascular niche and the endosteal niche apparently play a relevant role in the process not only of physiological stem cell homeostasis but also in clonal evolution of malignant stem cells and progression of disease, e.g. myelofibrosis, neoangiogenesis, and aberrant stem cell trafficking in Ph- MPN. Endothelial cells in JAK2 (V617F) Ph- MPN may belong to the malignant clone [82,83].

Janus Kinase 2/ JAK2 (K539L and other Aberrations in Exon 12)

In the diagnostic setting patients may show high haemoglobin levels, high hematocrit and also low serum EPO levels but do not harbour the JAK2 (V617F) mutation. These patients must be considered to be affected by a rare but important defect in exon 12 of JAK2. These molecular aberrations are diverse and may be point mutations leading to a K539L substitution or comprise insertions/deletions leading to various amino acid substitutions [84]. The bone marrow in patients affected by exon 12 aberrations shows a prominent erythropoiesis, but otherwise no true peculiarities, because also the megakaryocyte population appear normal in size, number and distribution. However, another study found the bone marrow with JAK2 exon 12 mutations more peculiar with also the megakaryocyte population being affected [85]. Interestingly, with regard to JAK2 exon 12 aberrations a large retrospective study found two features which were noteworthy to stress: 1. Gender did matter because more women than men were affected when compared to idiopathic erythrocytosis showing the JAK2 wild-type status, 2. Patients presented at younger age when compared to JAK2 (V617F), [86]. However, compared to JAK2 (V617F) the frequency of JAK2 exon 12 aberrations is low, molecular diagnostics is laborious, and molecular testing should be restricted to individuals which are JAK2 (V617F) negative and whose clinical parameters strongly suggest a clonal erythropoiesis.

LNK (SH2B adaptor protein 3 - SHLB3)

LNK has a central role in the control of JAK2 activity and involves receptors which recruit JAK2 like EPO-R and MPL [87]. Mice deficient in LNK showed an affected pool of hematopoietic stem cells (HSC), i.e. an increase of cells in numbers, but also extended self-renew capacity and quiescence [88]. LNK deficient cells showed an increased sensitivity to several cytokines and activated downstream pathways involving JAK2 [89]. LNK knock-out mice developed a disease with MPN-like features such as thrombocytosis, bone marrow fibrosis and extramedullary hematopoiesis. JAK2 (V617F) mutated animal models with a concomitant LNK deficiency showed a faster progress to a fullblown MPN-like disease [90]. As shown for both, JAK2 (V617F) and MPL (W515L), LNK is capable of hindering their aberrant constitutive signalling. Interestingly, in JAK2 (V617F) mutated MPN patients LNK expression was increased and correlated with a higher burden of the mutant JAK2 (V617F) allele [91]. This work also showed that LNK is an important negative controller of signals mediated through the TPO/ MPL axis. Up-regulation of LNK in JAK2 (V617F) patients can be interpreted as a counteracting mechanism.

In chronic MPN, LNK mutations affect only ~5% of ET and PMF patients, but in post-MPN AML LNK mutations were detectable in up to 13% of cases [92]. Accordingly, LNK aberrations appear to be late genetic events in the disease course of MPN. Interestingly, one study found LNK mutations in two patients with isolated erythrocytosis and wild-type JAK2 status and expanded the hitherto sparse knowledge on so-called idiopathic erythrocytosis [93].

Myeloproliferative Leukemia Virus Oncogene (MPL)

Thrombopoietin (TPO) and its receptor MPL is indispensable for hematopoietic stem cell survival, differentiation of hematopoietic progenitor cells, and is the key regulatory element of megakaryopoiesis and platelet formation [94]. The MPL receptor is expressed by hematopoietic tissues, hematopoietic stem cells, erythroid progenitors, megakaryocytes, and platelets [95].

Early studies in MPN showed that reduced megakaryocyte MPL expression is due to impaired post-translational modifications [11,96,97]. Demonstration of reduced platelet MPL expression and weakly labelled megakaryocytes by immunohistochemistry were found to be of diagnostic value in MPN [12,18,96,98,99]. Several studies demonstrated that megakaryocytes and the erythroid lineage co-express MPL and the erythropoietin receptor (EPO-R) as well as transcription factors like NF-E1 and -E2 [100-102]. Because of diverse structural homologies between MPL, EPO-R and their ligands, it became evident that the megakaryocytic and the erythroid lineage arise from a common progenitor cell [102]. Therefore, abnormal expression and function of MPL or EPO-R could induce proliferation of both the megakaryocytic and erythroid lineage [103,104].

In 2006, the discovery of another gain-of-function mutation in MPN was published [105]. Whereas mutation screening of the granulocyte-colony stimulating factor receptor (G-CSF-R) and the EPO-R in this cited work showed no molecular defect, a point mutation in the juxtamembrane domain of the thrombopoietin receptor MPL (exon 10) was discovered [105]. In MPL (W515L), like with the JAK2 (V617F) mutation, a single nucleotide is changed (G to T), leading to a substitution of the amino acid tryptophane by leucin. Both substitutions apparently change the folding of the protein or drive new interactions of the mutant protein with others. In the end, the function is substantially perturbed. Cell lines transfected with MPL (W515L) showed cytokine-independent growth and hypersensitivity for TPO. The MPL (W515L) mice model showed exaggerated number of platelets, a massive hepatosplenomegaly and splenic infarctions. This study and subsequently performed analysis in a large cohort of 1182 patients showed a MPL (W515L) frequency of ~5% in patients with PMF and ET [106]. It is of note that the latter study revealed another mutation MPL (W515K) as well as the occurrence of MPL mutations together with JAK2 (V617F) in individual patients. Correlation of the molecular status with clinical presentation has revealed no specific phenotype and analysis of larger trials was anticipated in order to find peculiarities of patients carrying the MPL (W515L/K) mutation. Beer et al. [107] investigated the occurrence of MPL mutations in 776 samples from ET patients [107]. The overall frequency of MPL exon 10 mutations was 8.5% in patients with unmutated JAK2. One patient with concomitant JAK2 (V617F) and MPL mutation could be identified. Patients with the W515K mutation had have a higher mutant allele burden than those with W515L. Compared with JAK2 (V617F) MPL mutated ET patients showed lower haemoglobin levels, higher EPO levels, higher platelets, an endogenous megakaryocyte colony growth but no EEC and an overall less bone marrow cellularity. However, MPL mutations lacked prognostic significance with regard to haemorrhage, thrombosis, development of myelofibrosis and survival.

The MPL (W515L/K) mutation was detectable in CD34-positive cells from patients with PMF which in vitro gave rise to multiple hematopoietic lineages carrying the mutation [108]. MPL (W515L/K) is like JAK2 (V617F) an event demonstrable in HSC and early progenitors.

Other MPL mutations such as the MPL (S505N) were found initially in familial MPN and later on also in sporadic MPN [109,110]. In a study of 221 patients with JAK2 (wild-type) ET and PMF, 33 (15%) carried a MPL mutation [110]. Only 1 ET patient carried both the MPL mutation and the JAK2 (V617F) mutation. Interestingly, 3 of the patients showed MPL double mutations on the same chromosome.

Neurofibromin (NF1)

The NF1 gene encodes for neurofibromin, a cytoplasmic protein that is predominantly expressed in neurons, Schwann cells, oligodendrocytes, and leukocytes. NF1 has the capacity to regulate several intracellular processes, e.g. the RAS-cyclic AMP pathway, the ERK/MAP serine/threonine kinase cascade, adenylyl cyclase, and cytoskeletal assembly [111]. Apparently 5 neurofibromin isoforms are expressed in different tissues of which isoform 10a possess a transmembrane domain. This isoform is widely expressed among human tissues.

NF1 has been suggested earlier to be a tumor suppressor because of its role in negative regulation of the Ras pathway and in the development of certain types of neurofibromatosis [112,113]. Later on, aberrations of NF1 in other cancers such as glioblastoma [114] and certain neurotropic melanoma types [115] were described.

Deletions of NF1 in chronic MPN are rare events with only 2 affected MPN patients in a study group of 151 patients [116]. Interestingly, a recent study found frequent codeletions of NF1 with SUZ12 as a member of the polycomb repressor complex 2 (PRC2) [46].

Neuroblastoma RAS (NRAS)

Mutations in RAS proteins (NRAS, KRAS, HRAS) were widely studied in subtypes of cancer such as colorectal cancer or lung cancer and mainly affect codon 61 in NRAS and codons 12/13 in KRAS, respectively. Mutations in RAS genes in the end disrupt GTPase activity with subsequent constitutive activity of pathways such as MAPK. Meanwhile RAS mutations predict the response of targeted therapies against growth factor receptors such as EGFR [117]. Among other genes such as CBL, RUNX1, TET2, NRAS was shown to be present in post-MPN AML and therefore is a genetic defect associated with disease progression [118].

Runt-related transcription factor 1 (RUNX1)

Alpha- and beta subunits of Runx transcription factors are encoded by RUNX1-3 genes and the CBFB gene, respectively. The alpha subunit binds directly to DNA whereas the beta subunit increases the binding affinity of the alpha-subunit without direct binding to DNA.

RUNX1 (alternatively named AML1) is critically involved in the proper development of all hematopoietic lineages but also is involved in development of certain neural compartments and bone formation [119]. Moreover, RUNX1/AML1 was shown to be involved in the t(8;21) translocation of the M2 subtype of AML according to the French- American-British-Classification (FAB) [120].

RUNX1 was suggested to be one of the genes associated with disease progression because mutations are frequently detectable in secondary AML (post-MPN) in 27%-37% but rarely in chronic MPN [118,121-123]. In a more descriptive study, we previously discussed that aberrant microRNA expression might be involved in dysregulated expression of RUNX1 in chronic MPN [124].

Splicing Factor 3B, Subunit 1 (SF3B1)

Splicing factor, serine/arginine-rich, 2 (SRSF2)

Genetic aberrations of factors involved in RNA splicing were previously reported for SF3B1 mainly in the MDS subtype refractory anaemia with ring sideroblasts (RARS). In addition to SF3B1 a landmark study found other affected factors involved in RNA splicing such as SRSF2, U2AF1, ZRSR2 and SF3A1 [125]. The effects of abnormal RNA splicing due to these aberrations compromised normal hematopoiesis and revealed a relevant insight into MDS pathobiology.

SF3B1 and SRSF2 mutations were detected in ~6% and ~17% of patients with PMF, respectively [126,127]. SRSF2 mutations in PMF patients were independently associated with poor survival, the concomitant affection by IDH1/2 mutations, and leukemia-free survival [127]. In another study post-MPN AML were investigated for SRSF2 mutations and showed a mutation rate of ~19% compared to secondary AML evolved from MDS (~5%) and de novo AML (~6%) [128]. SRSF2 mutations were associated with poor overall survival in univariate and multivariate analyses. It was concluded that SRSF2 is a relevant contributor to leukemic transformation in MPN.

Suppressor of Cytokine Signalling (SOCS1,-2,-3)

The suppressors of cytokine signalling (SOCS) are critically involved in the regulation of cellular proliferation, survival, and apoptosis via cytokine-induced JAK/STAT signalling. The suppressor of cytokine signalling -1 (SOCS-1) interacts with various components of cytokine and hematopoietic growth factor signalling including those activated by interleukin-6 or steel factor (c-kit) [129]. It has been demonstrated that SOCS-1 is a specific inhibitor of Janus kinase 2 (JAK2) signalling through binding of a conserved regulatory tyrosine in the activation loop of the catalytic domain JH1 of JAK2 [130]. Furthermore, at least 2 domains, the pre-SH2 domain or kinase inhibitory region (KIR), and the SH2 box, are responsible for effective binding and inhibition of signals transmitted by JAK2 [131].

In the accelerated phase of Ph+ BCR-ABL positive CML an overexpression of SOCS-2 was shown and it was suggested that the normally intact negative feedback loop should be overcome by SOCS-2 up-regulation [132]. We previously found an overexpression of SOCS1 mRNA in bone marrow cells of Ph- MPN [133]. This study involved an overall of 89 MPN cases and except for PMF showed an overexpression independent of the JAK2 status (V617F vs. wild-type). Another study showed hypermethylation of the SOCS1 promoter region in 6/39 MPN patients (3 with and 3 without JAK2-V617F) [134] which suggested decreased mRNA expression in these 15% of cases. Another work performed studies on the methylation status of CpG islands of SOCS1- SOCS3 in MPN and correlated successfully hypermethylation of SOCS1 or SOCS3 in 23/81 MPN cases (28%) with decreased mRNA expression in a given case [135]. Hypermethylation of SOCS2 was then demonstrated by another group in cytokine-independent cell lines and 2 out of 7 MPN patients [136].

Tet Oncogene Family Member 2 (TET2)

The first larger study on TET2 defects showed considerable variations including frameshifts, nonsense or missense mutations in exon 4 and 12 of patients with MPN [137]. The overall mutation frequency in MPN was ~13% with ~5% in ET and up to 16 % and ~17% in PV and PMF, respectively. TET2 mutations occur in both JAK2 (V617F) and JAK2 (wild-type) cases with frequencies of ~17% and ~7%, respectively. Interestingly, occurrence of TET2 aberrations significantly increase with age because ~23% of affected patients were above 60 years of age compared to only 4% of TET2 positive cases in younger individuals. The presence of TET2 mutations did not correlate with prognostic factors such as overall survival, rate of leukemic transformation or risk of thrombosis in PV and PMF. However, in patients with PV another study showed that TET2 mutations may occur in JAK2 (V617F) and JAK2 (wild-type) clones in an individual patient [138]. Interestingly, different TET2 alterations appear to occur in the same individual suggesting molecular dynamics probably due to genetic instability. Of note, this study did not observe a correlation of TET2 mutations and older age.

Following their first discovery in MPN TET2 mutations were also found in other myeloid malignancies with considerable high frequencies in de novo AML (42%) [139], in systemic mastocytosis (29%) [140], in MDS with rearranged 4q24 (19%), in CMML, in secondary AML (up to 32%) and in familial MPN [139-141].

Alterations in the TET2 gene therefore appear to be a common event in myeloid malignancies including those showing dysplastic features, chronic myeloproliferation and leukemic transformation. Haploinsufficiency of TET2 is thereby sufficient enough to contribute to the phenotype of disease because in the majority of cases only one allele is affected.

Before the discovery of TET2 alterations the function of this gene was almost unclear. However, TET1 was shown to catalyze the reaction from 5-methylcytosine in the DNA to 5-hydroxymethylcytosine [142] suggesting a role for TET proteins in epigenetic regulation. Recent studies in mice revealed that TET2 is a major player in the demethylation process of DNA. Animals deficient in TET2 showed decreased levels of 5-hydroxymethylcytosine along with high levels of 5-methylcytosine in the DNA, they showed less differentiation capacity of the hematopoietic stem cell compartment but increased potential of its self-renewal capacity [143,144]. Interestingly, in the TET2 knock out models the remaining differentiation capacity mainly affected the monocytic and granulocytic lineage. Following a latency of several months, the animals died from disease phenotypes very similar to CMML, secondary AML or even MDS [144]. These data underline the function of TET2 as a tumor suppressor which normally maintain normal hematopoiesis.

Tumor Protein p53 (TP53)

Among the genes affected by molecular defects in solid cancer and other malignancies TP53 is one of the best studied gene so far [145,146]. Normal function comprises involvement in DNA repair, control of cell cycle and cell fate, i.e. apoptosis and senescence. Accordingly, a loss-offunction defect in TP53 substantially affects homeostasis of a given cell.

TP53 mutations are frequently detectable in post-MPN AML but are occasionally detectable also in chronic MPN earlier before transformation becomes clinically overt [118]. Mutation of TP53 in post-AML was shown in ~27% of cases under investigation and affected mainly both alleles [34]. In one study TP53 aberrations were associated with gains of chromosome 1q, a locus harbouring the TP53 inhibitor MDM4 [147]. Post-MPN AML patients in this study were affected by TP53 or 1q gains in ~45% of cases.

Taken together, TP53 mutations are uncommon in chronic MPN but are very frequent in post-MPN AML suggesting this molecular defect as a potential predictor of disease progression and transformation. In one study TP53 mutations were the only independent prognostic factor of poor survival [34].

Therapeutic Targets

The following section does not claim to give a complete overview of currently established and up-coming therapies in MPN. However, it aims to summarize briefly the standard therapies in MPN followed by introduction of some studies which investigated promising agents for molecularly targeted therapies.

Standard therapies for PV patients focus on reduction of the high hematocrit by phlebotomy to avoid or minimize risks for thrombotic events or haemorrhage. Patients with high-risk PV may receive aspirine to improve microvascular events as well as hydroxyurea or pegylated interferon to control disease-related symptoms and improve quality of life [148]. Low-risk ET usually undergoes a “watch and wait” therapy, intermediate-risk ET may treated with aspirin, and patients with a prominent thrombocytosis and a high risk for thrombotic events must be treated with platelet-lowering agents such as anagrelide or hydroxyurea with or without aspirin [149,150]. In PMF, therapy can be adjusted to risk scores like the dynamic international prognostic scoring system (DPSS) [151], the IPSS or the International Working Group for Myelofibrosis Research and Treatment consensus criteria. It aims to improve constitutional symptoms which arise from organomegaly of spleen and liver but also from insufficient hematopoiesis leading to cytopenia. Drugs used in PMF treatment comprise erythropoiesisstimulating agents, androgens, danazol, corticosteroids, thalidomide, lenalidomide, hydroxyurea, and cladribine [152]. In combination these drugs may be induce longer lasting improvement of the disease course, e.g. when lenalidomide and corticosteroids were combined. Patients in one study showed a reduction of splenomegaly (42%), improvement of anemia (30%), a decline of myelofibrosis as evidenced by bone marrow histopathology in 10 out of 11 patients, and a lower JAK2 (V617F) mutant allele burden in 8 patients (decline of up to 50% in 4 patients, no longer detectable mutant allele burden in 1 patient) [153]. However, none of the aforementioned therapies was powerful enough to change the overall median survival in PMF which ranges from 3 – 5 years [152]. The only curative therapeutic option in PMF is the allogeneic stem cell transplantation which by implementation of intensity reduced chemotherapy regimen showed an event-free overall survival of 67% after 5 years [154].

Soon after discovery of the predominant molecular defects of JAK2 and MPL in MPN the first molecularly targeted therapies were initiated by using low molecular mass ATP-competitive inhibitors of either JAK2 or against JAK1 and JAK2 [155]. A selection of currently studied JAK kinase inhibitors comprise ruxolitinib (Incyte in the U.S., Novartis elsewhere), Lestaurtinib (Cephalon), Pacritinib (S*Bio), TG101348 (Sanofi), CYT387 (YM BioSciences), AZD1480 (AstraZeneca) and LY2784544 (Lilly) [156]. Some of these drugs are now in clinical phase I-III studies but the only drug approved for therapy in patients with intermediate to high risk MF by the Food and Drug Administration (FDA) in the U.S. so far is ruxolitinib. Ruxolitinib is an orally administered potent inhibitor of JAK1 and JAK2 with half maximal concentrations (IC50) of 3.3 nM and 2.8 nM, respectively. TYK2 and also JAK3 could be inhibited by ruxolitinib as well but the IC50 must then 6-fold and 130-fold higher, respectively [157]. Ruxolitinib inhibited downstream signalling by JAK2 (V617F) and wild-type JAK2 with reduced phosphorylation of JAK2 effectors STAT5 and ERK1/2 in vitro. Cell lines transfected with the JAK2 (V617F) showed reduced cellular proliferation and induction of apoptosis when treated with ruxolitinib [157]. Primary cells from PV patients with a high mutant JAK2 (V617F) allele burden showed decreased formation of endogenous erythroid colonies and a decline of erythroid/myeloid progenitor cells. In a mice model with a JAK2 (V617F) induced MPN-like disease, ruxolitinib (formerly called INCB018424) reduced splenomegaly, decreased peripheral cytokine levels (IL-6, TNF-α) to baseline, and increased survival of the animals [157].

Following promising phase I/II studies which showed a notable improvement of clinical symptoms without severe toxicity [158], the double-blinded placebo-controlled (phase III) COMFORT-I study included a total of 309 patients with PMF, post-PV-MF and post-ETMF having an IPSS score of intermediate to high risk. In the verum group (initial n=155) receiving 15 or 20 mg ruxolitinib (BID=twice daily) ~42% of patients showed a reduction of ≥ 35% spleen volume after 24 weeks when compared to 0.7% in the placebo group (initial n=154) [159]. While the placebo group showed a modest increase in the JAK2 (V617F) allele burden at 24 and 48 weeks, the ruxolitinib treated patients showed a decline of the mutant allele burden by ~10.9% and 21.5%, respectively. Notable benefits from clinical discomfort and relief of symptoms such as abdominal pain, itching, night sweats, bone/ muscle pain, inactivity/fatigue could be demonstrated in many patients treated with ruxolitinib in contrast to the placebo group. Ruxolitinib was all in all well tolerated. General side effects in the verum group were ecchymosis, dizziness, and headache along with haematological side effects such as thrombocytopenia and anemia. Discontinuation from the study was a rare event. After a median follow-up of 52 and 51 weeks, respectively, 13 patients in the verum group and 24 patients in the placebo group were dead [159].

In a randomized phase III study the COMFORT-II investigators compared ruxolitinib with the best available therapy (comprises hydroxyurea, interferon, corticosteroids, EPO, androgens and others; combination allowed) in patients with PMF, post-PV-MF and post- ET-MF having an IPSS score of intermediate to high risk and a defined size of palpable splenomegaly [160]. Patients were randomized 2:1 to ruxolitinib (n=146) and the best available therapy group (n=73). Primary and secondary efficacy outcome measures were a reduction of ≥ 35% spleen volume at week 48 and 24, respectively. In the ruxolitinib group 28% (week 48) and 32% (week 24) of patients showed a ≥ 35% spleen reduction compared to 0% (week 48) and 0% (week 24) receiving the best available therapy. 80% of the ruxolitinib treated patients still showed a therapy response after 12 months. As also shown in the COMFORT-I study the ruxolitinib group had a significantly better relief from clinical symptoms compared to the best available therapy regimen. After a median time of 61.1 weeks, 8% and 4% deaths were reported in the ruxolitinib and the best available therapy group, respectively [160]. Notable cross-over to the ruxolitinib group and incomplete survival follow-up was suggested to hinder reasonable interpretation at this time point. Main adverse effects on hematopoiesis were thrombocytopenia and anaemia with more severe cases in the ruxolitinib group compared to the best available therapy group.

Ruxolitinib was also investigated in PV and ET patients refractory for hydroxyurea treatment and in PV patients to test for superiority over the best available therapy [155].

Apart from ruxolitinib other molecularly targeted therapy studies were conducted or are under way by using lestaurtinib (inhibitor of FLT3, JAK2, RET = rearranged during transfection) in patients with PMF, post-PV-MF and post-ET-MF [161], pacritinib (SB1518, a JAK2 and FLT3 inhibitor) in patients with myelofibrosis [162], TG101348 (a JAK2-, FLT3-, RET- inhibitor) in patients with PMF, post-PV-MF and post-ET-MF [163], CYT387 (an inhibitor of JAK1, JAK2, TYK2) [164], AZD1480 (a JAK1 and JAK2 inhibitor) [165,166] and LY2784544 (JAK2 inhibitor) [167].

Especially FLT3 inhibitors were also widely investigated for their potency in de novo AML whereas some others like the newer JAK3 inhibitor tasocitinib seems to be also potent in diseases with a strong inflammatory background like rheumatoid arthritis [156]. Moreover, these small molecule inhibitors seem to act as immunmodulators and may help to prevent organ rejection after transplantation [168].

Other drugs investigated for targeting aberrant molecular pathways in MPN and other myeloid malignancies are inhibitors of the mTOR pathway such as everolimus [169], the EGFR inhibitor erlotinib [170] and the proteasome inhibitor bortezomib affecting the NFκB pathway [171]. Because a breakthrough by using a single agent is pending, investigators proposed combination therapies such as the mTOR pathway inhibitors RAD001 and PP242 with AZD1480 and ruxolitinib [166].

Compounds developed for targeting especially epigenetic changes in MPN were givinostat (ITF2357), panobinostat (LBH589), vorinostat and decitabine (5-aza-2’-deoxycytidine) and might reverse changes such as histone H3 phosphorylation or aberrant methylation [156].

It is noteworthy to underline that above mentioned targeted therapies in MPN so far do not directly hit any of the molecular defects discovered up to date. The number of JAK2 inhibitors all target both mutant and wild-type JAK2 and based on the fact that JAK2 (V617F) is not the founder of the malignant clone JAK2 inhibitors will not cure the MPN. On the other hand JAK2 inhibitors decrease cytokine signalling and the pro-inflammatory response which is apparently responsible for a number of clinical complaints in MPN patients. The above reviewed effects of JAK2 inhibitors such as ruxolitinib are indeed a notable relief from clinical symptoms which cannot be underestimated. The rationale for a molecularly targeted therapy with one of the established inhibitors is therefore apparent.

Cytokine signalling and mainly secondary changes in the bone marrow environment due to inflammatory processes, hypoxia, proangiogenic signalling by VEGF with increased micro vessel density as demonstrable in PMF induce some of the clinical symptoms in MPN. Therefore, the inhibition of cytokine signalling by JAK inhibitors affect cellular homeostasis and reduce MPN-typical features. On the other side the relevant adverse effects of JAK inhibition affect hematopoiesis leading mainly to anaemia and low platelet counts. Accordingly, a continuous therapy will compromise the clinical course in other ways.

It is likely that upcoming therapies will be composed of different compounds directed against the variety of molecular defects demonstrable in a given patient. Future targeted therapy in Ph- MPN therefore could be a personalized therapy based on the individual set of affected genes.

Summary

First of all, the discoveries of molecular defects in the Ph- MPN over the recent few years were awesome. Research groups and networks composed of experts in haematology, pathology and molecular biology in the U.S., Europe and all over the world cooperated to pave the way for one of the most impressive progress in myeloid malignancies. Only 7 years after the somatic mutation in JAK2 (V617F) was detected this breakthrough not only proved to be indispensable for a more accurate diagnosis. JAK2 in Ph- MPN is meanwhile target for small molecule inhibitors and stands for a new era of molecular therapeutics.

After discovery of JAK2 (V617F) and MPL (W515L/K) in 2005 and 2006, respectively, many other molecular aberrations were found in Ph- MPN and associated myeloid malignancies to better discriminate from reactive lesions and, at least in part, allow a more distinct differential diagnosis. For example the JAK2 (K539L) mutation is probably another prototype for an entity-specific molecular defect in patients with PV-like clinical presentation but lack of JAK2 (V617F) and bone marrow features uncommon for PV. Moreover, these findings expand the knowledge on the fundamental pathobiology in chronic diseases which for many decades were more on the sidelines. Accordingly, the publication record for PMF, PV and ET climbed up enormously. However, there is probably no direct correlation between the number of molecular defects discovered in Ph- MPN so far and the extent of true understanding of the nature of the three entities in question.

We have to realize that 3 distinct diseases in terms of histopathological and clinical presentation exist even though overlaps are not uncommon.

We realize that many molecular defects except for those associated with leukemic transformation are demonstrable in all 3 diseases at the chronic stage.

It is widely agreed that JAK2 (V617F) is an important event which substantially contributes to the phenotype of a Ph- MPN. However, it is well accepted that other somatic molecular defects precede JAK2 (V617F) in a HSC to found clonal hematopoiesis. Accordingly, Ph- MPN phenotypes are the result of different genotypes, i.e. accumulating somatic molecular hits form a signature responsible for what clinically become apparent. Aberrations in ASXL1, TET2, EZH2 and probably in yet unknown genes are important examples for early hits in Ph- MPN before occurrence of JAK2 (V617F). On the other hand, a molecular hit through aberrant TET2 function can also occur later in the disease course when JAK2 (V617F) is already a driver over a long time.

Even though Ph- MPN pathobiology starts on the level of a HSC it is likely that progenitor cells of different lineages (all affected by the “master defect”) later on acquire a different signature of additional molecular hits. This also could explain the different phenotype with more demonstrable hits underlying PMF pathobiology compared to the more mild disease course in ET showing less genetic lesions. Moreover, the hit signature probably explains the different risks for transformation into secondary AML.

Accordingly, a comprehensive molecular profiling for a much better discrimination of subtypes in Ph- MPN would be ideal. An elegant study recently performed a so-called genome-wide methylome/epigenome profiling which showed aberrant promoter hypermethylation in PV and ET whereas PMF was epigenetically a distinct subgroup with aberrant hyper- and hypomethylation [172]. Methylated genes in PV and ET affected signalling pathways enriched for binding sites for transcription factors such as GATA-1. By contrast, aberrantly methylated genes in PMF were more involved in inflammatory pathways and were enriched for transcription factors other than PV and ET. Interestingly, within PMF cases those with ASXL1 mutations formed a distinct subgroup showing all in all increased methylation [172]. As expected TET2 mutated MPN cases showed decreased levels of hydroxymethylation and, much more intriguing, another distinct set of hypermethylated genes.

The latter study is another example for the importance and impact of molecular techniques which should be applied to gain a better insight into the pathobiology of the Ph- MPN. Once discovered, molecular marker or sets of marker not only allow a better understanding in an intellectual sense but preferably generate markers for a more accurate diagnosis.

We have robust and evidenced criteria for the diagnosis of a given Ph- MPN. Fortunately, this still includes histopathological evaluation of a bone marrow biopsy which in my opinion is a very important instrument for diagnosis of MPN and most other myeloid malignancies. Accordingly, the doctor should convince the patient not only about the necessity of a core biopsy for initial diagnosis but also for the agreement to take sequential biopsies during the time course. This, because many important features especially of Ph- MPN like the first slight increase of blasts or any changes in stroma pathology, e.g. a regression of myelofibrosis during therapy, is detectable only through this invasive procedure. Because of the progress in laboratory technologies over the last 10-15 years the bone marrow biopsy, when adequately preserved and gently decalcified, is almost unrestricted stable and useful for molecular analysis of DNA, mRNA and microRNA [173-175]. Moreover, the correlation of histopathology and molecular data in one biopsy sample optimize diagnosis finding.

However, the demonstration of some of the hitherto discovered molecular markers in Ph- MPN is somewhat challenging with regard to techniques which are required for optimal sample analyses. While a JAK2 (V617F) assay meanwhile is part of the standard molecular diagnostics in Ph- MPN, a comprehensive analysis for TET2 aberrations is more elaborated and time-consuming. The molecular nature of TET2 aberrations and other genes such as CBL can be more complex (Table 1) and will probably not be unmasked by a single laboratory assay. Accordingly, the introduction of appropriate assays which are still robust even when performed in different laboratories is of utmost necessity. For example, the limited accuracy of standard SANGER sequencing for detection of some of the currently used molecular markers in Ph- MPN and other myeloid malignancies is an important issue which has to be considered [176]. For both the diagnostic setting and basic clinical research the most sensitive and the best workable technique must be established.

Where do we currently stand with the Ph- MPN?

The number of molecular defects detectable in Ph- MPN continuously increases. All the markers lack specificity because they are demonstrable also in a variety of other myeloid malignancies. Yet no marker stands definitely for a beginning transformation into secondary leukemia, yet no definite marker may indicate a beginning myelofibrosis and last not least, standardized laboratory assays for many markers are pending. In other words, the complexity of Ph- MPN became more complex and not easier to look through.

Nevertheless, we now have a variety of markers which can be studied in more detail not only to describe their occurrence in a given entity but to reveal more relevant insights into pathobiology. Technical progress should allow a more comprehensive determination of molecular defects present in a given patient and in the end might lead to an individualized molecularly targeted therapy.

Notice

The term “chronic MPN” as used throughout the manuscript comprises disease stages without hints of disease progression and leukemic transformation.

If not otherwise cited this review was inspired by a previously published book chapter contributed by this author and colleagues [177].

The author apologize that many excellent experimental data and clinical research as well as comprehensive reviews were not cited due to the large number of publications in the field and due to space restriction.

References

  1. Tefferi A, Vardiman JW (2008) Classification and diagnosis of myeloproliferative neoplasms: the 2008 World Health Organization criteria and point-of-care diagnostic algorithms. Leukemia 22: 14-22.
  2. Kralovics R, Passamonti F, Buser AS, Teo SS, Tiedt R, et al. (2005) A gain-of-function mutation of JAK2 in myeloproliferative disorders. N Engl J Med 352: 1779-1790.
  3. Baxter EJ, Scott LM, Campbell PJ, East C, Fourouclas N, et al. (2005) Acquired mutation of the tyrosine kinase JAK2 in human myeloproliferative disorders. Lancet 365: 1054-1061.
  4. James C, Ugo V, Le Couedic JP, Staerk J, Delhommeau F, et al. (2005). A unique clonal JAK2 mutation leading to constitutive signalling causes polycythaemia vera. Nature 434: 1144-1148.
  5. Levine RL, Wadleigh M, Cools J, Ebert BL, Wernig G, et al. (2005). Activating mutation in the tyrosine kinase JAK2 in polycythemia vera, essential thrombocythemia, and myeloid metaplasia with myelofibrosis. Cancer Cell 7: 387-397.
  6. Reilly JT, Wilson G, Barnett D, Watmore A, Potter A (1994) Karyotypic and ras gene mutational analysis in idiopathic myelofibrosis. Br J Haematol 88: 575-581.
  7. Kralovics R, Guan Y, Prchal JT (2002) Acquired uniparental disomy of chromosome 9p is a frequent stem cell defect in polycythemia vera. Exp Hematol 30: 229-236.
  8. Prchal JF, Axelrad AA (1974) Letter: Bone-marrow responses in polycythemia vera. N Engl J Med 290: 1382.
  9. Adamson JW, Fialkow PJ, Murphy S, Prchal JF, Steinmann L (1976) Polycythemia vera: stem-cell and probable clonal origin of the disease. N Engl J Med 295: 913-916.
  10. Prchal JF, Adamson JW, Murphy S, Steinmann L, Fialkow PJ (1978) Polycythemia vera. The in vitro response of normal and abnormal stem cell lines to erythropoietin. J Clin Invest 61: 1044-1047.
  11. Moliterno AR, Hankins WD, Spivak JL (1998) Impaired expression of the thrombopoietin receptor by platelets from patients with polycythemia vera. N Engl J Med 338: 572-580.
  12. Teofili L, Pierconti F, Di Febo A, Maggiano N, Vianelli N, et al. (2002) The expression pattern of c-mpl in megakaryocytes correlates with thrombotic risk in essential thrombocythemia. Blood 100: 714-717.
  13. Martyré MC, Magdelenat H, Bryckaert MC, Laine-Bidron C, Calvo F (1991) Increased intraplatelet levels of platelet-derived growth factor and transforming growth factor-beta in patients with myelofibrosis with myeloid metaplasia. Br J Haematol 77: 80-86.
  14. Bock O, Schlué J, Lehmann U, von Wasielewski R, Länger F, et al. (2002) Megakaryocytes from chronic myeloproliferative disorders show enhanced nuclear bFGF expression. Blood 100: 2274-2275.
  15. Steurer M, Zoller H, Augustin F, Fong D, Heiss S, et al. (2007) Increased angiogenesis in chronic idiopathic myelofibrosis: vascular endothelial growth factor as a prominent angiogenic factor. Hum Pathol 38: 1057-1064.
  16. Gianelli U, Vener C, Raviele PR, Savi F, Somalvico F, et al. (2007) VEGF expression correlates with microvessel density in Philadelphia chromosome-negative chronic myeloproliferative disorders. Am J Clin Pathol 128: 966-973.
  17. Muth M, Büsche G, Bock O, Hussein K, Kreipe H (2010) Aberrant proplatelet formation in chronic myeloproliferative neoplasms. Leuk Res 34: 1424-1429.
  18. Mesa RA, Hanson CA, Li CY, Yoon SY, Rajkumar SV, et al. (2002) Diagnostic and prognostic value of bone marrow angiogenesis and megakaryocyte c-Mpl expression in essential thrombocythemia. Blood 99: 4131-4137.
  19. Temerinac S, Klippel S, Strunck E, Röder S, Lübbert M, et al. (2000) Cloning of PRV-1, a novel member of the uPAR receptor superfamily, which is overexpressed in polycythemia rubra vera. Blood 95: 2569-2576.
  20. Fisher CL, Pineault N, Brookes C, Helgason CD, Ohta H, et al. (2010) Loss-of-function Additional sex combs like 1 mutations disrupt hematopoiesis but do not cause severe myelodysplasia or leukemia. Blood 115: 38-46.
  21. Carbuccia N, Murati A, Trouplin V, Brecqueville M, Adélaïde J, et al. (2009) Mutations of ASXL1 gene in myeloproliferative neoplasms. Leukemia 23: 2183-2186.
  22. Brecqueville M, Rey J, Bertucci F, Coppin E, Finetti P, et al. (2012) Mutation analysis of ASXL1, CBL, DNMT3A, IDH1, IDH2, JAK2, MPL, NF1, SF3B1, SUZ12, and TET2 in myeloproliferative neoplasms. Genes Chromosomes Cancer 51: 743-755.
  23. Abdel-Wahab O, Manshouri T, Patel J, Harris K, Yao J, et al. (2010) Genetic analysis of transforming events that convert chronic myeloproliferative neoplasms to leukemias. Cancer Res 70: 447-452.
  24. Gelsi-Boyer V, Trouplin V, Adélaïde J, Bonansea J, Cervera N, et al. (2009) Mutations of polycomb-associated gene ASXL1 in myelodysplastic syndromes and chronic myelomonocytic leukaemia. Br J Haematol 145: 788-800.
  25. Metzeler KH, Becker H, Maharry K, Radmacher MD, Kohlschmidt J, et al. (2011) ASXL1 mutations identify a high-risk subgroup of older patients with primary cytogenetically normal AML within the ELN Favorable genetic category. Blood 118: 6920-6929.
  26. Grand FH, Hidalgo-Curtis CE, Ernst T, Zoi K, Zoi C, et al. (2009) Frequent CBL mutations associated with 11q acquired uniparental disomy in myeloproliferative neoplasms. Blood 113: 6182-6192.
  27. Sanada M, Suzuki T, Shih LY, Otsu M, Kato M, et al. (2009) Gain-of-function of mutated C-CBL tumour suppressor in myeloid neoplasms. Nature 460: 904-908.
  28. Smit L, Borst J (1997) The Cbl family of signal transduction molecules. Crit Rev Oncog 8: 359-379.
  29. Joazeiro CA, Wing SS, Huang H, Leverson JD, Hunter T, et al. (1999) The tyrosine kinase negative regulator c-Cbl as a RING-type, E2-dependent ubiquitin-protein ligase. Science 286: 309-312.
  30. Lievens PM, Donady JJ, Tufarelli C, Neufeld EJ (1995) Repressor activity of CCAAT displacement protein in HL-60 myeloid leukemia cells. J Biol Chem 270: 12745-12750.
  31. Zeng WR, Scherer SW, Koutsilieris M, Huizenga JJ, Filteau F, et al. (1997) Loss of heterozygosity and reduced expression of the CUTL1 gene in uterine leiomyomas. Oncogene 14: 2355-2365.
  32. Hulea L, Nepveu A (2012) CUX1 transcription factors: from biochemical activities and cell-based assays to mouse models and human diseases. Gene 497: 18-26.
  33. Klampfl T, Harutyunyan A, Berg T, Gisslinger B, Schalling M, et al. (2011) Genome integrity of myeloproliferative neoplasms in chronic phase and during disease progression. Blood 118: 167-176.
  34. Milosevic JD, Puda A, Malcovati L, Berg T, Hofbauer M, et al. (2012) Clinical significance of genetic aberrations in secondary acute myeloid leukemia. Am J Hematol 87: 1010-1016.
  35. Chédin F (2011) The DNMT3 family of mammalian de novo DNA methyltransferases. Prog Mol Biol Transl Sci 101: 255-285.
  36. Ley TJ, Ding L, Walter MJ, McLellan MD, Lamprecht T, et al. (2010) DNMT3A mutations in acute myeloid leukemia. N Engl J Med 363: 2424-2433.
  37. Abdel-Wahab O, Pardanani A, Rampal R, Lasho TL, Levine RL, et al. (2011) DNMT3A mutational analysis in primary myelofibrosis, chronic myelomonocytic leukemia and advanced phases of myeloproliferative neoplasms. Leukemia 25: 1219-1220.
  38. Stegelmann F, Bullinger L, Schlenk RF, Paschka P, Griesshammer M, et al. (2011) DNMT3A mutations in myeloproliferative neoplasms. Leukemia 25: 1217-1219.
  39. Margueron R, Reinberg D (2011) The Polycomb complex PRC2 and its mark in life. Nature 469: 343-349.
  40. Sauvageau M, Sauvageau G (2010) Polycomb group proteins: multi-faceted regulators of somatic stem cells and cancer. Cell Stem Cell 7: 299-313.
  41. Margueron R, Li G, Sarma K, Blais A, Zavadil J, et al. (2008) Ezh1 and Ezh2 maintain repressive chromatin through different mechanisms. Mol Cell 32: 503-518.
  42. Ernst T, Chase AJ, Score J, Hidalgo-Curtis CE, Bryant C, et al. (2010) Inactivating mutations of the histone methyltransferase gene EZH2 in myeloid disorders. Nat Genet 42: 722-726.
  43. Guglielmelli P, Biamonte F, Score J, Hidalgo-Curtis C, Cervantes F, et al. (2011) EZH2 mutational status predicts poor survival in myelofibrosis. Blood 118: 5227-5234.
  44. Kroeze LI, Nikoloski G, da Silva-Coelho P, van Hoogen P, Stevens-Linders E, et al. (2012) Genetic defects in PRC2 components other than EZH2 are not common in myeloid malignancies. Blood 119: 1318-1319.
  45. Score J, Hidalgo-Curtis C, Jones AV, Winkelmann N, Skinner A, et al. (2012) Inactivation of polycomb repressive complex 2 components in myeloproliferative and myelodysplastic/myeloproliferative neoplasms. Blood 119: 1208-1213.
  46. Puda A, Milosevic JD, Berg T, Klampfl T, Harutyunyan AS, et al. (2012) Frequent deletions of JARID2 in leukemic transformation of chronic myeloid malignancies. Am J Hematol 87: 245-250.
  47. O'Neill DW, Schoetz SS, Lopez RA, Castle M, Rabinowitz L, et al. (2000) An ikaros-containing chromatin-remodeling complex in adult-type erythroid cells. Mol Cell Biol 20: 7572-7582.
  48. Nakase K, Ishimaru F, Avitahl N, Dansako H, Matsuo K, et al. (2000) Dominant negative isoform of the Ikaros gene in patients with adult B-cell acute lymphoblastic leukemia. Cancer Res 60: 4062-4065.
  49. Mullighan CG, Miller CB, Radtke I, Phillips LA, Dalton J, et al. (2008) BCR-ABL1 lymphoblastic leukaemia is characterized by the deletion of Ikaros. Nature 453: 110-114.
  50. Jäger R, Gisslinger H, Passamonti F, Rumi E, Berg T, et al. (2010) Deletions of the transcription factor Ikaros in myeloproliferative neoplasms. Leukemia 24: 1290-1298.
  51. Borodovsky A, Seltzer MJ, Riggins GJ (2012) Altered cancer cell metabolism in gliomas with mutant IDH1 or IDH2. Curr Opin Oncol 24: 83-89.
  52. Parsons DW, Jones S, Zhang X, Lin JC, Leary RJ, et al. (2008) An integrated genomic analysis of human glioblastoma multiforme. Science 321: 1807-1812.
  53. Mardis ER, Ding L, Dooling DJ, Larson DE, McLellan MD, et al. (2009) Recurring mutations found by sequencing an acute myeloid leukemia genome. N Engl J Med 361: 1058-1066.
  54. Green A, Beer P (2010) Somatic mutations of IDH1 and IDH2 in the leukemic transformation of myeloproliferative neoplasms. N Engl J Med 362: 369-370.
  55. Hussein K, Engelhardt BM, Kreipe H, Bock O (2010) IDH mutation analysis is not suitable for the routine molecular diagnostic algorithm in myeloproliferative and myelodysplastic neoplasms. Blood 116: 5073-5074.
  56. Gross S, Cairns RA, Minden MD, Driggers EM, Bittinger MA, et al. (2010) Cancer-associated metabolite 2-hydroxyglutarate accumulates in acute myelogenous leukemia with isocitrate dehydrogenase 1 and 2 mutations. J Exp Med 207: 339-344.
  57. Sasaki M, Knobbe CB, Munger JC, Lind EF, Brenner D, et al. (2012) IDH1(R132H) mutation increases murine haematopoietic progenitors and alters epigenetics. Nature 488: 656-659.
  58. Figueroa ME, Abdel-Wahab O, Lu C, Ward PS, Patel J, et al. (2010) Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 18: 553-567.
  59. Xu W, Yang H, Liu Y, Yang Y, Wang P, et al. (2011) Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell 19: 17-30.
  60. Rane SG, Reddy EP (2002) JAKs, STATs and Src kinases in hematopoiesis. Oncogene 21: 3334-3358.
  61. Ghoreschi K, Laurence A, O'Shea JJ (2009) Janus kinases in immune cell signaling. Immunol Rev 228: 273-287.
  62. Lippert E, Boissinot M, Kralovics R, Girodon F, Dobo I, et al. (2006) The JAK2-V617F mutation is frequently present at diagnosis in patients with essential thrombocythemia and polycythemia vera. Blood 108: 1865-1867.
  63. Tefferi A, Lasho TL, Schwager SM, Strand JS, Elliott M, et al. (2006) The clinical phenotype of wild-type, heterozygous, and homozygous JAK2V617F in polycythemia vera. Cancer 106: 631-635.
  64. Antonioli E, Guglielmelli P, Pancrazzi A, Bogani C, Verrucci M, et al. (2005) Clinical implications of the JAK2 V617F mutation in essential thrombocythemia. Leukemia 19: 1847-1849.
  65. Wolanskyj AP, Lasho TL, Schwager SM, McClure RF, Wadleigh M, et al. (2005) JAK2 mutation in essential thrombocythaemia: clinical associations and long-term prognostic relevance. Br J Haematol 131: 208-213.
  66. Horn T, Kremer M, Dechow T, Pfeifer WM, Geist B, et al. (2006) Detection of the activating JAK2 V617F mutation in paraffin-embedded trephine bone marrow biopsies of patients with chronic myeloproliferative diseases. J Mol Diagn 8: 299-304.
  67. Bock O, Büsche G, Koop C, Schröter S, Buhr T, et al. (2006) Detection of the single hotspot mutation in the JH2 pseudokinase domain of Janus kinase 2 in bone marrow trephine biopsies derived from chronic myeloproliferative disorders. J Mol Diagn 8: 170-177.
  68. Oku S, Takenaka K, Kuriyama T, Shide K, Kumano T, et al. (2010) JAK2 V617F uses distinct signalling pathways to induce cell proliferation and neutrophil activation. Br J Haematol 150: 334-344.
  69. Wernig G, Mercher T, Okabe R, Levine RL, Lee BH, et al. (2006) Expression of Jak2V617F causes a polycythemia vera-like disease with associated myelofibrosis in a murine bone marrow transplant model. Blood 107: 4274-4281.
  70. Lacout C, Pisani DF, Tulliez M, Gachelin FM, Vainchenker W, et al. (2006) JAK2V617F expression in murine hematopoietic cells leads to MPD mimicking human PV with secondary myelofibrosis. Blood 108: 1652-1660.
  71. Akada H, Yan D, Zou H, Fiering S, Hutchison RE, et al. (2010) Conditional expression of heterozygous or homozygous Jak2V617F from its endogenous promoter induces a polycythemia vera-like disease. Blood 115: 3589-3597.
  72. Bock O, Neuse J, Hussein K, Brakensiek K, Buesche G, et al. (2006) Aberrant collagenase expression in chronic idiopathic myelofibrosis is related to the stage of disease but not to the JAK2 mutation status. Am J Pathol 169: 471-481.
  73. Hussein K, Brakensiek K, Buesche G, Buhr T, Wiese B, et al. (2007) Different involvement of the megakaryocytic lineage by the JAK2 V617F mutation in Polycythemia vera, essential thrombocythemia and chronic idiopathic myelofibrosis. Ann Hematol 86: 245-253.
  74. Campbell PJ, Scott LM, Buck G, Wheatley K, East CL, et al. (2005) Definition of subtypes of essential thrombocythaemia and relation to polycythaemia vera based on JAK2 V617F mutation status: a prospective study. Lancet 366: 1945-1953.
  75. Gautier EF, Picard M, Laurent C, Marty C, Villeval JL, et al. (2012) The cell cycle regulator CDC25A is a target for JAK2V617F oncogene. Blood 119: 1190-1199.
  76. Kiladjian JJ, Elkassar N, Cassinat B, Hetet G, Giraudier S, et al. (2006) Essential thrombocythemias without V617F JAK2 mutation are clonal hematopoietic stem cell disorders. Leukemia 20: 1181-1183.
  77. Kralovics R, Teo SS, Li S, Theocharides A, Buser AS, et al. (2006) Acquisition of the V617F mutation of JAK2 is a late genetic event in a subset of patients with myeloproliferative disorders. Blood 108: 1377-1380.
  78. Campbell PJ, Baxter EJ, Beer PA, Scott LM, Bench AJ, et al. (2006) Mutation of JAK2 in the myeloproliferative disorders: timing, clonality studies, cytogenetic associations, and role in leukemic transformation. Blood 108: 3548-3555.
  79. Jones AV, Chase A, Silver RT, Oscier D, Zoi K, et al. (2009) JAK2 haplotype is a major risk factor for the development of myeloproliferative neoplasms. Nat Genet 41: 446-449.
  80. Olcaydu D, Harutyunyan A, Jäger R, Berg T, Gisslinger B, et al. (2009) A common JAK2 haplotype confers susceptibility to myeloproliferative neoplasms. Nat Genet 41: 450-454.
  81. Theocharides A, Boissinot M, Girodon F, Garand R, Teo SS, et al. (2007) Leukemic blasts in transformed JAK2-V617F-positive myeloproliferative disorders are frequently negative for the JAK2-V617F mutation. Blood 110: 375-379.
  82. Oppliger Leibundgut E, Horn MP, Brunold C, Pfanner-Meyer B, Marti D, et al. (2006) Hematopoietic and endothelial progenitor cell trafficking in patients with myeloproliferative diseases. Haematologica 91: 1465-1472.
  83. Lataillade JJ, Pierre-Louis O, Hasselbalch HC, Uzan G, Jasmin C, et al. (2008) Does primary myelofibrosis involve a defective stem cell niche? From concept to evidence. Blood 112: 3026-3035.
  84. Scott LM, Tong W, Levine RL, Scott MA, Beer PA, et al. (2007) JAK2 exon 12 mutations in polycythemia vera and idiopathic erythrocytosis. N Engl J Med 356: 459-468.
  85. Lakey MA, Pardanani A, Hoyer JD, Nguyen PL, Lasho TL, et al. (2010) Bone marrow morphologic features in polycythemia vera with JAK2 exon 12 mutations. Am J Clin Pathol 133: 942-948.
  86. Schnittger S, Bacher U, Haferlach C, Geer T, Müller P, et al. (2009) Detection of JAK2 exon 12 mutations in 15 patients with JAK2V617F negative polycythemia vera. Haematologica 94: 414-418.
  87. Tong W, Zhang J, Lodish HF (2005) Lnk inhibits erythropoiesis and Epo-dependent JAK2 activation and downstream signaling pathways. Blood 105: 4604-4612.
  88. Bersenev A, Wu C, Balcerek J, Tong W (2008) Lnk controls mouse hematopoietic stem cell self-renewal and quiescence through direct interactions with JAK2. J Clin Invest 118: 2832-2844.
  89. Velazquez L, Cheng AM, Fleming HE, Furlonger C, Vesely S, et al. (2002) Cytokine signaling and hematopoietic homeostasis are disrupted in Lnk-deficient mice. J Exp Med 195: 1599-1611.
  90. Bersenev A, Wu C, Balcerek J, Jing J, Kundu M, et al. (2010) Lnk constrains myeloproliferative diseases in mice. J Clin Invest 120: 2058-2069.
  91. Baran-Marszak F, Magdoud H, Desterke C, Alvarado A, Roger C, et al. (2010) Expression level and differential JAK2-V617F-binding of the adaptor protein Lnk regulates JAK2-mediated signals in myeloproliferative neoplasms. Blood 116: 5961-5971.
  92. Pardanani A, Lasho T, Finke C, Oh ST, Gotlib J, et al. (2010) LNK mutation studies in blast-phase myeloproliferative neoplasms, and in chronic-phase disease with TET2, IDH, JAK2 or MPL mutations. Leukemia 24: 1713-1718.
  93. Lasho TL, Pardanani A, Tefferi A (2010) LNK mutations in JAK2 mutation-negative erythrocytosis. N Engl J Med 363: 1189-1190.
  94. Kaushansky K (2002) Mpl and the hematopoietic stem cell. Leukemia 16: 738-739.
  95. Kaushansky K (1998) The role of the MPL receptor in myeloproliferative disorders. Leukemia 12: S47-50.
  96. Horikawa Y, Matsumura I, Hashimoto K, Shiraga M, Kosugi S, et al. (1997) Markedly reduced expression of platelet c-mpl receptor in essential thrombocythemia. Blood 90: 4031-4038.
  97. Moliterno AR, Spivak JL (1999) Posttranslational processing of the thrombopoietin receptor is impaired in polycythemia vera. Blood 94: 2555-2561.
  98. Tefferi A, Yoon SY, Li CY (2000) Immunohistochemical staining for megakaryocyte c-mpl may complement morphologic distinction between polycythemia vera and secondary erythrocytosis. Blood 96: 771-772.
  99. Yoon SY, Li CY, Tefferi A (2000) Megakaryocyte c-Mpl expression in chronic myeloproliferative disorders and the myelodysplastic syndrome: immunoperoxidase staining patterns and clinical correlates. Eur J Haematol 65: 170-174.
  100. Martin DI, Zon LI, Mutter G, Orkin SH (1990) Expression of an erythroid transcription factor in megakaryocytic and mast cell lineages. Nature 344: 444-447.
  101. Vigon I, Mornon JP, Cocault L, Mitjavila MT, Tambourin P, et al. (1992) Molecular cloning and characterization of MPL, the human homolog of the v-mpl oncogene: identification of a member of the hematopoietic growth factor receptor superfamily. Proc Natl Acad Sci U S A 89: 5640-5644.
  102. McDonald TP, Sullivan PS (1993) Megakaryocytic and erythrocytic cell lines share a common precursor cell. Exp Hematol 21: 1316-1320.
  103. Kieran MW, Perkins AC, Orkin SH, Zon LI (1996) Thrombopoietin rescues in vitro erythroid colony formation from mouse embryos lacking the erythropoietin receptor. Proc Natl Acad Sci U S A 93: 9126-9131.
  104. Yan XQ, Lacey DL, Saris C, Mu S, Hill D, et al. (1999) Ectopic overexpression of c-mpl by retroviral-mediated gene transfer suppressed megakaryopoiesis but enhanced erythropoiesis in mice. Exp Hematol 27: 1409-1417.
  105. Pikman Y, Lee BH, Mercher T, McDowell E, Ebert BL, et al. (2006) MPLW515L is a novel somatic activating mutation in myelofibrosis with myeloid metaplasia. PLoS Med 3: e270.
  106. Pardanani AD, Levine RL, Lasho T, Pikman Y, Mesa RA, et al. (2006) MPL515 mutations in myeloproliferative and other myeloid disorders: a study of 1182 patients. Blood 108: 3472-3476.
  107. Beer PA, Campbell PJ, Scott LM, Bench AJ, Erber WN, et al. (2008) MPL mutations in myeloproliferative disorders: analysis of the PT-1 cohort. Blood 112: 141-149.
  108. Chaligné R, James C, Tonetti C, Besancenot R, Le Couédic JP, et al. (2007) Evidence for MPL W515L/K mutations in hematopoietic stem cells in primitive myelofibrosis. Blood 110: 3735-3743.
  109. Chaligné R, Tonetti C, Besancenot R, Roy L, Marty C, et al. (2008) New mutations of MPL in primitive myelofibrosis: only the MPL W515 mutations promote a G1/S-phase transition. Leukemia 22: 1557-1566.
  110. Pietra D, Brisci A, Rumi E, Boggi S, Elena C, et al. (2011) Deep sequencing reveals double mutations in cis of MPL exon 10 in myeloproliferative neoplasms. Haematologica 96: 607-611.
  111. Trovó-Marqui AB, Tajara EH (2006) Neurofibromin: a general outlook. Clin Genet 70: 1-13.
  112. DeClue JE, Papageorge AG, Fletcher JA, Diehl SR, Ratner N, et al. (1992) Abnormal regulation of mammalian p21ras contributes to malignant tumor growth in von Recklinghausen (type 1) neurofibromatosis. Cell 69: 265-273.
  113. Basu TN, Gutmann DH, Fletcher JA, Glover TW, Collins FS, et al. (1992) Aberrant regulation of ras proteins in malignant tumour cells from type 1 neurofibromatosis patients. Nature 356: 713-715.
  114. Cancer Genome Atlas Research Network (2008) Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455: 1061-1068.
  115. Gutzmer R, Herbst RA, Mommert S, Kiehl P, Matiaske F, et al. (2000) Allelic loss at the neurofibromatosis type 1 (NF1) gene locus is frequent in desmoplastic neurotropic melanoma. Hum Genet 107: 357-361.
  116. Stegelmann F, Bullinger L, Griesshammer M, Holzmann K, Habdank M, et al. (2010) High-resolution single-nucleotide polymorphism array-profiling in myeloproliferative neoplasms identifies novel genomic aberrations. Haematologica 95: 666-669.
  117. McCubrey JA, Steelman LS, Chappell WH, Abrams SL, Wong EW, et al. (2007) Roles of the Raf/MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Biochim Biophys Acta 1773: 1263-1284.
  118. Beer PA, Delhommeau F, LeCouédic JP, Dawson MA, Chen E, et al. (2010) Two routes to leukemic transformation after a JAK2 mutation-positive myeloproliferative neoplasm. Blood 115: 2891-2900.
  119. Cohen MM Jr (2009) Perspectives on RUNX genes: an update. Am J Med Genet A 149A: 2629-2646.
  120. Miyoshi H, Shimizu K, Kozu T, Maseki N, Kaneko Y, et al. (1991) t(8;21) breakpoints on chromosome 21 in acute myeloid leukemia are clustered within a limited region of a single gene, AML1. Proc Natl Acad Sci U S A 88: 10431-10434.
  121. Steensma DP, Gibbons RJ, Mesa RA, Tefferi A, Higgs DR (2005) Somatic point mutations in RUNX1/CBFA2/AML1 are common in high-risk myelodysplastic syndrome, but not in myelofibrosis with myeloid metaplasia. Eur J Haematol 74: 47-53.
  122. Ding Y, Harada Y, Imagawa J, Kimura A, Harada H (2009) AML1/RUNX1 point mutation possibly promotes leukemic transformation in myeloproliferative neoplasms. Blood 114: 5201-5205.
  123. Kreipe HH (2011) Precursors of acute leukemia: myelodysplastic syndromes and myeloproliferative neoplasms. Pathologe 32: 271-276.
  124. Hussein K, Dralle W, Theophile K, Kreipe H, Bock O (2009) Megakaryocytic expression of miRNA 10a, 17-5p, 20a and 126 in Philadelphia chromosome-negative myeloproliferative neoplasm. Ann Hematol 88: 325-332.
  125. Yoshida K, Sanada M, Shiraishi Y, Nowak D, Nagata Y, et al. (2011) Frequent pathway mutations of splicing machinery in myelodysplasia. Nature 478: 64-69.
  126. Lasho TL, Finke CM, Hanson CA, Jimma T, Knudson RA, et al. (2012) SF3B1 mutations in primary myelofibrosis: clinical, histopathology and genetic correlates among 155 patients. Leukemia 26: 1135-1137.
  127. Lasho TL, Jimma T, Finke CM, Patnaik M, Hanson CA, et al. (2012) SRSF2 mutations in primary myelofibrosis: significant clustering with IDH mutations and independent association with inferior overall and leukemia-free survival. Blood 120: 4168-4171.
  128. Zhang SJ, Rampal R, Manshouri T, Patel J, Mensah N, et al. (2012) Genetic analysis of patients with leukemic transformation of myeloproliferative neoplasms shows recurrent SRSF2 mutations that are associated with adverse outcome. Blood 119: 4480-4485.
  129. Ilangumaran S, Rottapel R (2003) Regulation of cytokine receptor signaling by SOCS1. Immunol Rev 192: 196-211.
  130. Yasukawa H, Misawa H, Sakamoto H, Masuhara M, Sasaki A, et al. (1999) The JAK-binding protein JAB inhibits Janus tyrosine kinase activity through binding in the activation loop. EMBO J 18: 1309-1320.
  131. Giordanetto F, Kroemer RT (2003) A three-dimensional model of Suppressor Of Cytokine Signalling 1 (SOCS-1). Protein Eng 16: 115-124.
  132. Schultheis B, Carapeti-Marootian M, Hochhaus A, Weisser A, Goldman JM, et al. (2002) Overexpression of SOCS-2 in advanced stages of chronic myeloid leukemia: possible inadequacy of a negative feedback mechanism. Blood 99: 1766-1775.
  133. Bock O, Hussein K, Brakensiek K, Buhr T, Schlué J, et al. (2007) The suppressor of cytokine signalling-1 (SOCS-1) gene is overexpressed in Philadelphia chromosome negative chronic myeloproliferative disorders. Leuk Res 31: 799-803.
  134. Jost E, do ON, Dahl E, Maintz CE, Jousten P, et al. (2007) Epigenetic alterations complement mutation of JAK2 tyrosine kinase in patients with BCR/ABL-negative myeloproliferative disorders. Leukemia 21: 505-510.
  135. Teofili L, Martini M, Cenci T, Guidi F, Torti L, et al. (2008) Epigenetic alteration of SOCS family members is a possible pathogenetic mechanism in JAK2 wild type myeloproliferative diseases. Int J Cancer 123: 1586-1592.
  136. Quentmeier H, Geffers R, Jost E, Macleod RA, Nagel S, et al. (2008) SOCS2: inhibitor of JAK2V617F-mediated signal transduction. Leukemia 22: 2169-2175.
  137. Tefferi A, Pardanani A, Lim KH, Abdel-Wahab O, Lasho TL, et al. (2009) TET2 mutations and their clinical correlates in polycythemia vera, essential thrombocythemia and myelofibrosis. Leukemia 23: 905-911.
  138. Delhommeau F, Dupont S, Della Valle V, James C, Trannoy S, et al. (2009) Mutation in TET2 in myeloid cancers. N Engl J Med 360: 2289-2301.
  139. Abdel-Wahab O, Mullally A, Hedvat C, Garcia-Manero G, Patel J, et al. (2009) Genetic characterization of TET1, TET2, and TET3 alterations in myeloid malignancies. Blood 114: 144-147.
  140. Tefferi A, Levine RL, Lim KH, Abdel-Wahab O, Lasho TL, et al. (2009) Frequent TET2 mutations in systemic mastocytosis: clinical, KITD816V and FIP1L1-PDGFRA correlates. Leukemia 23: 900-904.
  141. Saint-Martin C, Leroy G, Delhommeau F, Panelatti G, Dupont S, et al. (2009) Analysis of the ten-eleven translocation 2 (TET2) gene in familial myeloproliferative neoplasms. Blood 114: 1628-1632.
  142. Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H, et al. (2009) Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324: 930-935.
  143. Moran-Crusio K, Reavie L, Shih A, Abdel-Wahab O, Ndiaye-Lobry D, et al. (2011) Tet2 loss leads to increased hematopoietic stem cell self-renewal and myeloid transformation. Cancer Cell 20: 11-24.
  144. Li Z, Cai X, Cai CL, Wang J, Zhang W, et al. (2011) Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies. Blood 118: 4509-4518.
  145. Vousden KH, Lane DP (2007) p53 in health and disease. Nat Rev Mol Cell Biol 8: 275-283.
  146. Bourdon JC (2007) p53 and its isoforms in cancer. Br J Cancer 97: 277-282.
  147. Harutyunyan A, Klampfl T, Cazzola M, Kralovics R (2011) p53 lesions in leukemic transformation. N Engl J Med 364: 488-490.
  148. Cortelazzo S, Finazzi G, Ruggeri M, Vestri O, Galli M, et al. (1995) Hydroxyurea for patients with essential thrombocythemia and a high risk of thrombosis. N Engl J Med 332: 1132-1136.
  149. Harrison CN, Campbell PJ, Buck G, Wheatley K, East CL, et al. (2005) Hydroxyurea compared with anagrelide in high-risk essential thrombocythemia. N Engl J Med 353: 33-45.
  150. Tefferi A (2013) Primary myelofibrosis: 2013 update on diagnosis, risk-stratification, and management. Am J Hematol 88: 141-150.
  151. Quintás-Cardama A, Kantarjian HM, Manshouri T, Thomas D, Cortes J, et al. (2009) Lenalidomide plus prednisone results in durable clinical, histopathologic, and molecular responses in patients with myelofibrosis. J Clin Oncol 27: 4760-4766.
  152. Kröger N, Holler E, Kobbe G, Bornhäuser M, Schwerdtfeger R, et al. (2009) Allogeneic stem cell transplantation after reduced-intensity conditioning in patients with myelofibrosis: a prospective, multicenter study of the Chronic Leukemia Working Party of the European Group for Blood and Marrow Transplantation. Blood 114: 5264-5270.
  153. Mascarenhas J, Mughal TI, Verstovsek S (2012) Biology and clinical management of myeloproliferative neoplasms and development of the JAK inhibitor ruxolitinib. Curr Med Chem 19: 4399-4413.
  154. Quintás-Cardama A, Kantarjian H, Cortes J, Verstovsek S (2011) Janus kinase inhibitors for the treatment of myeloproliferative neoplasias and beyond. Nat Rev Drug Discov 10: 127-140.
  155. Quintás-Cardama A, Vaddi K, Liu P, Manshouri T, Li J, et al. (2010) Preclinical characterization of the selective JAK1/2 inhibitor INCB018424: therapeutic implications for the treatment of myeloproliferative neoplasms. Blood 115: 3109-3117.
  156. Verstovsek S, Kantarjian H, Mesa RA, Pardanani AD, Cortes-Franco J, et al. (2010) Safety and efficacy of INCB018424, a JAK1 and JAK2 inhibitor, in myelofibrosis. N Engl J Med 363: 1117-1127.
  157. Verstovsek S, Mesa RA, Gotlib J, Levy RS, Gupta V, et al. (2012) A double-blind, placebo-controlled trial of ruxolitinib for myelofibrosis. N Engl J Med 366: 799-807.
  158. Harrison C, Kiladjian JJ, Al-Ali HK, Gisslinger H, Waltzman R, et al. (2012) JAK inhibition with ruxolitinib versus best available therapy for myelofibrosis. N Engl J Med 366: 787-798.
  159. Santos FP, Kantarjian HM, Jain N, Manshouri T, Thomas DA, et al. (2010) Phase 2 study of CEP-701, an orally available JAK2 inhibitor, in patients with primary or post-polycythemia vera/essential thrombocythemia myelofibrosis. Blood 115: 1131-1136.
  160. William AD, Lee AC, Blanchard S, Poulsen A, Teo EL, et al. (2011) Discovery of the macrocycle 11-(2-pyrrolidin-1-yl-ethoxy)-14,19-dioxa-5,7,26-triaza-tetracyclo[19.3.1.1(2,6).1(8,12)]heptacosa-1(25),2(26),3,5,8,10,12(27),16,21,23-decaene (SB1518), a potent Janus kinase 2/fms-like tyrosine kinase-3 (JAK2/FLT3) inhibitor for the treatment of myelofibrosis and lymphoma. J Med Chem 54: 4638-4658.
  161. Pardanani A, Gotlib JR, Jamieson C, Cortes JE, Talpaz M, et al. (2011) Safety and efficacy of TG101348, a selective JAK2 inhibitor, in myelofibrosis. J Clin Oncol 29: 789-796.
  162. Tyner JW, Bumm TG, Deininger J, Wood L, Aichberger KJ, et al. (2010) CYT387, a novel JAK2 inhibitor, induces hematologic responses and normalizes inflammatory cytokines in murine myeloproliferative neoplasms. Blood 115: 5232-5240.
  163. Ioannidis S, Lamb ML, Wang T, Almeida L, Block MH, et al. (2011) Discovery of 5-chloro-N2-[(1S)-1-(5-fluoropyrimidin-2-yl)ethyl]-N4-(5-methyl-1H-pyrazol-3-yl)pyrimidine-2,4-diamine (AZD1480) as a novel inhibitor of the Jak/Stat pathway. J Med Chem 54: 262-276.
  164. Bogani C, Bartalucci N, Martinelli S, Tozzi L, Guglielmelli P, et al. (2013) mTOR inhibitors alone and in combination with JAK2 inhibitors effectively inhibit cells of myeloproliferative neoplasms. PLoS One 8: e54826.
  165. Passamonti F, Maffioli M, Caramazza D (2012) New generation small-molecule inhibitors in myeloproliferative neoplasms. Curr Opin Hematol 19: 117-123.
  166. Quaedackers ME, Mol W, Korevaar SS, van Gurp EA, van Ijcken WF, et al. (2009) Monitoring of the immunomodulatory effect of CP-690,550 by analysis of the JAK/STAT pathway in kidney transplant patients. Transplantation 88: 1002-1009.
  167. Guglielmelli P, Barosi G, Rambaldi A, Marchioli R, Masciulli A, et al. (2011) Safety and efficacy of everolimus, a mTOR inhibitor, as single agent in a phase 1/2 study in patients with myelofibrosis. Blood 118: 2069-2076.
  168. Li Z, Xu M, Xing S, Ho WT, Ishii T, et al. (2007) Erlotinib effectively inhibits JAK2V617F activity and polycythemia vera cell growth. J Biol Chem 282: 3428-3432.
  169. Mesa RA, Verstovsek S, Rivera C, Pardanani A, Hussein K, et al. (2008) Bortezomib therapy in myelofibrosis: a phase II clinical trial. Leukemia 22: 1636-1638.
  170. Nischal S, Bhattacharyya S, Christopeit M, Yu Y, Zhou L, et al. (2013) Methylome profiling reveals distinct alterations in phenotypic and mutational subgroups of myeloproliferative neoplasms. Cancer Res 73: 1076-1085.
  171. Lewis F, Maughan NJ, Smith V, Hillan K, Quirke P (2001) Unlocking the archive--gene expression in paraffin-embedded tissue. J Pathol 195: 66-71.
  172. Lehmann U, Kreipe H (2001) Real-time PCR analysis of DNA and RNA extracted from formalin-fixed and paraffin-embedded biopsies. Methods 25: 409-418.
  173. Dijkstra JR, Mekenkamp LJ, Teerenstra S, De Krijger I, Nagtegaal ID (2012) MicroRNA expression in formalin-fixed paraffin embedded tissue using real time quantitative PCR: the strengths and pitfalls. J Cell Mol Med 16: 683-690.
  174. Kohlmann A, Klein HU, Weissmann S, Bresolin S, Chaplin T, et al. (2011) The Interlaboratory RObustness of Next-generation sequencing (IRON) study: a deep sequencing investigation of TET2, CBL and KRAS mutations by an international consortium involving 10 laboratories. Leukemia 25: 1840-1848.
  175. Porwit A, McCullough J, Erber WN (2011) Blood and Bone Marrow Pathology. Churchill Livingstone, Edinburgh, London, New York, Oxford, Philadelphia, St. Louis, Sydney, Toronto.
Citation: Bock O (2013) The Current Molecular Site of the Myeloproliferative Neoplasms - TET-á-tête with the Jackpot but no LNK so far to Resolve Complexity. J Leuk (Los Angel) 1:108.

Copyright: © 2013 Bock O. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top