Loss of Ezh2 Promotes the Development of Mutant-RUNX1 Induced MDS

Blood ◽  
2012 ◽  
Vol 120 (21) ◽  
pp. 402-402
Author(s):  
Goro Sashida ◽  
Satomi Tanaka ◽  
Makiko Mochizuki-Kashio ◽  
Atsunori Saraya ◽  
Tomoya Muto ◽  
...  

Abstract Abstract 402 Polycomb group proteins are transcriptional repressors that epigenetically regulate transcription via histone modifications. There are two major polycomb-complexes, the Polycomb Repressive Complexes 1 and 2 (PRC1, PRC2). PRC2 contains SUZ12, EED, and EZH1/EZH2, and catalyzes the trimethylation of histone H3 at lysine 27 (H3K27me3), silencing target-genes. We have shown that the self-renewal of Ezh2-deficient HSCs is not compromised and H3K27me3 marks are not completely depleted in the absence of Ezh2, possibly as a result of Ezh1 complementation. EZH2 is generally thought to act as an oncogene in lymphoma and solid tumors by silencing tumor suppressor genes. Recently however, loss-of-function mutations of EZH2 have been found in myeloid malignancies such as AML, MDS and MPN, suggesting that EZH2 also functions as a tumor suppressor, although it remains unclear how EZH2 prevents the transformation of myeloid malignancies. RUNX1 is a critical transcription factor in the regulation of the self-renewal and differentiation of HSCs. RUNX1 mutations are frequently found in MDS, AML following MDS (MDS/AML) and de novo AML patients. One of the most frequent mutations, RUNX1S291fs, lacks the transactivation domain in C-terminus, but retains the RUNT DNA biding domain, resulting in a dominant negative phenotype. RUNX1S291fs-transduced bone marrow cells have been shown to generate MDS/AML in vivo. Given that RUNX1 and EZH2 mutations coexist in MDS and AML patients as reported recently, we generated a novel mouse model of MDS utilizing RUNX1S291fs retrovirus and Ezh2 conditional knockout mice in order to understand how EZH2 loss contributes to the pathogenesis of MDS upon genetic mutation of RUNX1. We first harvested CD34-Lin-Sca1+c-Kit+(LSK) HSCs from tamoxifen-inducible Cre-ERT;Ezh2wild/wild (EW) and Cre-ERT;Ezh2flox/flox (EF) mice (CD45.2) and transduced these cells with RUNX1S291fs retrovirus or an empty vector, which contains IRES-GFP. Then, we transplanted RUNX1S291fs-transduced Cre-ERT;Ezh2wild/wild (S291EW) or Cre-ERT;Ezh2flox/flox (S291EF) HSCs into lethally irradiated recipient mice (CD45.1) together with life saving dose 1×105 CD45.1 bone marrow cells. At 6 weeks post transplantation, we deleted Ezh2 via administration of tamoxifen, and observed disease progression until 12 months post transplantation. The empty vector transduced control mice with or without Ezh2 (EW and EF) did not develop myeloid malignancies. Two out of 16 S291EW mice died due to MDS progression, while 12 out of 16 and 1 out of 17 S291EF mice developed MDS and MDS/AML, respectively. S291EF mice showed significantly shorter median survival than S291EW mice (314 days versus undefined, p=0.037). In the peripheral blood, we observed significantly lower CD45.2+GFP+ chimerism in S291EF mice; however S291EF mice eventually showed macrocytic anemia and variable white blood cell counts accompanied with dysplastic features of MDS. Despite low CD45.2+GFP+ chimerism in peripheral blood, S291EF mice showed a higher chimerism of CD45.2+GFP+ cells in the bone marrow and had a significantly increased number of LSK and CD34-LSK cells compared to EW, EF, and S291EW mice, indicating that Ezh2 loss promoted HSCs/progenitors expansion, but impaired myeloid differentiation in the presence of RUNX1S291fs. We also saw enhanced apoptosis of CD71+Ter119+ erythroblasts in S291EF MDS mice, which may account for the anemia we observed. Since S291EF MDS bone marrow cells were transplantable in secondary experiments, we performed limiting-dilution assays to evaluate the frequency of MDS initiating cells and found that the frequency of MDS initiating cells was much higher in S291EF pre-MDS Lin-Mac1-Kit+ cells compared to S291EW pre-MDS Lin-Mac1-Kit+ cells. To understand this molecular mechanism, we performed gene expression analysis during MDS progression. S291EF MDS LSKs showed 1979 and 1875 dysregulated (>5-fold) genes, compared to EW LSK and S291EF pre-MDS LSK, respectively. We are now working to understand how these dysregulated genes are involved in the development of RUNX1S291fs-induced MDS after deletion of Ezh2. In summary, we have successfully recapitulated the clinical feature of MDS in mice reconstituted with Ezh2 null HSCs expressing a RUNX1 mutant, and demonstrated that Ezh2 functions as a tumor suppressor in this context. Disclosures: No relevant conflicts of interest to declare.

Blood ◽  
2008 ◽  
Vol 112 (11) ◽  
pp. 1406-1406
Author(s):  
Matthew J Boyer ◽  
Feng Xu ◽  
Hui Yu ◽  
Tao Cheng

Abstract DNA methylation is an epigenetic means of gene regulation and is carried out by a family of methyltransferases of which DNMT1 acts to maintain methylation marks following DNA replication and DNMT3a and DNMT3b methylate DNA de novo. DNMT3b has been shown to be essential for mammalian development and necessary for differentiation of germline and neural progenitor cells. Mutations of DNMT3b in humans lead to a rare autosomal recessive disorder characterized by immunodeficiency, centromeric instability, and facial abnormalities. We have shown by real-time, RT-PCR that DNMT3b mRNA is uniquely over-expressed by approximately 30-fold in immunophenotypically-defined longterm repopulating hematopoietic stem cells (HSCs) that are CD34−lineage−c-kit+Sca-1+ as compared to progenitor and differentiated cell types within the bone marrow and with respect to the other members of the DNMT family, namely DNMT1 and DNMT3a. To determine DNMT3b’s function in HSCs competitive bone marrow transplantation was undertaken. Isolated lineage− enriched bone marrow cells were transduced with a retroviral backbone based on the Murine Stem Cell Virus (MSCV) carrying either GFP and a short, hairpin RNA (shRNA) targeting DNMT3b or GFP alone. Following transduction 1×105 GFP+ cells along with 1×105 competitor cells were transplanted into 9.5 Gray irradiated congenic recipients. Two months following transplantation mice receiving bone marrow cells transduced with DNMT3b shRNA showed a significantly lower engraftment of donor cells as a percentage of total competitor cell engraftment in the peripheral blood as compared to those receiving cells transduced with GFP alone (24.8 vs 3.7, p<0.05) which persisted at 3 months (22.8 vs 1.5, p<0.05). Similarly, within the donor derviced cells in the peripheral blood there was a lower percentage of myeloid (CD11b+) cells at 2 and 3 months in the recipients of DNMT3b shRNA transduced cells as compared to controls. However there was no observed difference in the percentage of peripheral B (CD45R+) or T (CD3+) cells within the donor-derived cells. To determine the mechanism behind the observed engraftment defect with DNMT3b knockdown we cultured GFP+ transduced bone marrow cells in vitro with minimal cytokine support. As a control for our targeting methodology we also transduced bone marrow cells from mice harboring two floxed DNMT3b alleles with a MSCV carrying Cre recombinase and GFP. While lineage− bone marrow cells transduced with GFP alone increased 10-fold in number over two weeks of culture, cells in which DNMT3b was down regulated by shRNA or Cre-mediated recombination only doubled. Culture of lineage− bone marrow cells in methylcellulose medium by the colony-forming cell (CFC) assay revealed increases in the granulocytic and total number of colonies with DNMT3b knockdown or Cre-mediated recombination of DNMT3b similar to the increased myeloid engraftment of DNMT3b shRNA transduced cells observed 1 month following competitive bone marrow transplantation. However when 5,000 of these cells from the first CFC assay were sub-cultured there was a significant loss of colony forming ability within all lineages when DNMT3b was targeted by shRNA or Cre-mediated recombination. Taken together with the decreased engraftment of DNMT3b shRNA cells following competitive bone marrow transplantation, the observed limited proliferation in liquid culture and loss of colony forming ability during serial CFC assays is suggestive of a self-renewal defect of HSCs in the absence of DNMT3b, that was previously only reported in the absence of both DNMT3a and DNMT3b. Further elucidation of this proposed self-renewal defect is being undertaken and results of ongoing studies including long-term culture initiating cell (LTC-IC) assays and identification of genomic sites of DNA methylation within different hematopoietic subsets will also be presented.


Blood ◽  
2012 ◽  
Vol 120 (21) ◽  
pp. 553-553
Author(s):  
Cara L Lunn ◽  
Justin Tibbitts ◽  
James N Ley ◽  
Jin Shao ◽  
Timothy Graubert ◽  
...  

Abstract Abstract 553 Myelodysplastic syndromes (MDS) are stem cell disorders characterized by ineffective hematopoiesis with increased levels of hematopoietic cell apoptosis. Recent discoveries by our group and others suggest that perturbations in pre-mRNA splicing may play a role in MDS pathogenesis. Indeed, more than half of all MDS patients have a mutation in one of eight splicing factors. U2AF1 (U2AF35), a gene encoding a splicing factor involved in intronic 3'-splice site recognition, is mutated in 8.7% of MDS patients. The consequence of the highly recurrent serine to phenylalanine mutation at position 34 (S34F) of U2AF1 in hematopoiesis is unknown. Therefore, to examine the effects of mutant U2AF1(S34F) on hematopoiesis, we utilized the MSCV-IRES-GFP retroviral system to introduce mutant U2AF1(S34F) or wild type U2AF1(WT), or an empty vector control, into mouse bone marrow cells for in vitro and in vivo studies. Expression of U2AF1(S34F) results in reduced expansion of transduced bone marrow cells (marked by GFP) compared to both U2AF1(WT) or empty vector-transduced cells grown in suspension culture (2 vs 4-fold change, respectively; p<0.001, n=3). Additionally, U2AF1(S34F)-transduced cells have increased levels of apoptosis (Annexin V+/7AAD+) in culture compared to U2AF1(WT) (p=0.03) and empty vector-transduced cells (p=0.02) (n=3). We also examined the effects of the U2AF1(S34F) mutation in vivo using bone marrow transplantation. The percentage of GFP+ cells in the peripheral blood of recipient mice transplanted with MSCV-transduced bone marrow was significantly reduced at 6 months post-transplant with U2AF1(S34F) expression (average=4%) compared to U2AF1(WT) (average=44%) and empty vector (average=65%) (p<0.02, n= 6–9 mice each). Transduction efficiencies were similar within experiments. There was no consistent alteration in lineage distribution of GFP+ cells in the peripheral blood of these mice. To overcome some of the limitations of retroviral models, we created a single-copy, doxycycline-inducible U2AF1(S34F) transgenic mouse to model the effect of U2AF1(S34F) expression on hematopoiesis. As a control for U2AF1 protein overexpression, we created an additional single-copy, doxycycline-inducible U2AF1(WT) transgenic mouse with transgene integration into the same locus as the U2AF1(S34F) mouse. Induction of U2AF1(S34F) transgene expression in bone marrow cells in culture with doxycycline treatment (200 ng/ml for 5 days) resulted in reduced cell numbers when compared to uninduced U2AF1(S34F) transgenic cells (ratio of growth of induced/uninduced cells = 0.38), while cell proliferation was not altered for U2AF1(WT) transgenic cells (ratio of growth of induced/uninduced cells = 1.13) (p<0.001, n=3). In addition, doxycycline-induced U2AF1(S34F) expression results in increased apoptosis (Annexin V+) compared to uninduced U2AF1(S34F) transgenic cells (21% vs 11%, p=0.01) and induced U2AF1(WT) transgenic cells in culture (21% vs 9.3%, p=0.008) (n=4). To examine the effects of mutant U2AF1(S34F) induction in vivo, we transplanted mutant U2AF1(S34F) or U2AF1(WT) transgenic bone marrow cells into congenic wild type recipient mice and induced transgene expression 6 weeks post-transplant using 2 mg/ml doxycycline in the drinking water for 5 days. Induction of U2AF1(S34F) expression in vivo results in reduced number of WBCs in the peripheral blood of recipient mice compared to mice with uninduced U2AF1(S34F) transgenic bone marrow (3.4k vs 5.6k, p=0.01, n=3). In addition, recipient mice with induced U2AF1(S34F) bone marrow had reduced number of bone marrow cells per femur when compared to uninduced U2AF1(S34F) recipient mice (3.9M vs 13.1M, p=0.04) and induced U2AF1(WT) recipient mice (3.9M vs 12.4M, p=0.03) (n=3). The number of neutrophils in peripheral blood (p<0.001), bone marrow (p=0.04), and spleen (p=0.04) of induced U2AF1(S34F) recipient mice were all significantly lower compared to uninduced U2AF1(S34F) mice (n=3). The total numbers of c-Kit+/lineage-/Sca+ hematopoietic progenitor cells were not affected in induced U2AF1(S34F) recipient mice compared to uninduced U2AF1(S34F) (p=0.75) or induced U2AF1(WT) recipient mice (p=0.46, n=3) after 5 days of treatment. Collectively, these results suggest that the U2AF1(S34F) mutation may contribute to abnormal hematopoiesis in vivo. Longer periods of doxycycline-induction in vivo are ongoing and will be presented. Disclosures: No relevant conflicts of interest to declare.


Blood ◽  
2004 ◽  
Vol 104 (11) ◽  
pp. 2788-2788
Author(s):  
Lilia Stepanova ◽  
Brian P. Sorrentino

Abstract Homeobox (Hox) transcription factors are important regulators of hematopoietic cell proliferation and differentiation. Of them, HoxB4 is of particular interest because overexpression promotes rapid expansion of mouse hematopoietic stem cells (HSCs) without causing neoplastic transformation. Despite the effects of HoxB4 overexpression on HSCs, mice that are homozygous for HoxB4 gene deletion have only subtle defects in HSCs and progenitor cells. We hypothesized that other paralogs of HoxB4 may also be capable of inducing HSC expansion could thereby compensate for loss of HoxB4 function. To test this hypothesis, we have studied the effects of retroviral overexpression of a HoxC4 gene in murine progenitors and HSCs. The murine HoxC4 cDNA was cloned and inserted into an MSCV vector that co-expresses an IRES-YFP reporter gene. We transduced murine bone marrow cells with a MSCV-HoxC4-YFP vector and compared the secondary replating efficiency of myeloid colonies (CFU-Cs) to that seen using either a MSCV-HoxB4-GFP or an MSCV-GFP vector. This assay tests for progenitor cell self-renewal which is increased using HoxB4-expressing vectors. Cells transduced with the MSCV-HoxC4-YFP vector formed 20–40 times more secondary CFU-Cs than with cells transduced with the MSCV-GFP control vector. This increase in CFU-C replating efficiency was equivalent to that seen with the MSCV-HoxB4-IRES-GFP vector. To test the in vivo effects of the MSCV-HoxC4-YFP vector on self-renewal of HSCs, we transplanted lethally irradiated mice with a mixture of cells; 20% transduced with the MSCV-HoxC4-YFP vector and 80 % mock-transduced. Peripheral blood analysis of the transplanted recipients up to 28 weeks post-transplantation showed that the percentage of cells transduced with the MSCV-HoxC4-YFP vector was 70–85% in both lymphoid and myeloid cells in the peripheral blood. A similar degree of chimerism was noted in concurrent controls using the MSCV-HoxB4-GFP vector. In contrast, the percentages of peripheral blood cells transduced with the MSCV-GFP vector was only 15–25%, paralleling the input ratios of transplanted cells. Secondary transplantation experiments showed stable levels of chimerism in both HoxC4 and HoxB4 groups, indicating that the expansion seen with the MSCV-HoxC4-YFP vector occurred at the HSC level. These results indicate that retroviral-mediated expression of HoxC4, like HoxB4, can cause significant expansion of HSCs in vivo. Because several other Hox genes can cause hematopoietic abnormalities and leukemia when expressed from a retroviral vector, we transplanted lethally irradiated mice with 4x106 cells that were transduced with the MSCV-HoxC4-YFP vector and monitored the animals for survival and complete blood counts. Now, at 33 weeks post transplantation, no tumor formation was observed in mice expressing either the HoxB4 or the HoxC4 vector, and peripheral blood counts have remained normal. Our results show that retroviral overexpression of HoxC4 can induce a significant expansion of the HSCs in vivo, and suggest that expression of HoxC4 may compensate for the loss of HoxB4 in knockout mice. We are currently analyzing the effects of HoxA4 and HoxD4 to determine if they share the same functional characteristics, and are also determining whether HoxB4 and HoxC4 are modulating the same downstream genes using microarray analysis of transduced murine bone marrow cells.


Blood ◽  
2008 ◽  
Vol 112 (11) ◽  
pp. 932-932 ◽  
Author(s):  
John S. Welch ◽  
Wenlin Yuan ◽  
Timothy James Ley

Abstract Acute promyelocytic leukemia (APL) is characterized by the t(15;17) translocation that leads to expression of a fusion protein, PML-RARα, and haploinsufficiency for both RARα and PML. We have generated a novel murine model of APL using homologous recombination to place a pathogenic human PML-RARα cDNA into the murine PML locus (mPML-PRflox). Expression of PML-RARα is initially prevented by stop codons within the loxP-flanked PGK-neo cassette. This new model recapitulates key elements of human APL lacking in other models; it provides PML locus appropriate regulation of PML-RARα expression, haploinsufficiency of PML, and somatic acquisition of the fusion protein. We have exposed mPML-PRflox mice to a conditionally-active Cre transgene (ER-T2- Cre, which activates Cre only during Tamoxifen treatment), and a compartmentally restricted Cre transgene (Lysozyme M-Cre, which has low activity in early myeloid precursors and increasing activity with myeloid differentiation). Bone marrow and spleen cells doubly heterozygous (DH) for mPML-PRflox and an activated Cre allele express PML-RARα mRNA, display neutrophil POD disruption typical of PML-RARα activity, and exhibit a shift in myelopoiesis toward CFU-G formation. We found that expression of PML-RARα following transient ER-T2-Cre activation leads to myeloid self-renewal ex vivo and clonal expansion in vivo. Bone marrow cells from DH mPML-PRflox/ER-T2- Cre mice exposed to Tamoxifen could be serially replated in methylcellulose. With successive replating, the proportion of cells carrying a floxed PML-RARα allele increased from 30% in bone marrow cells to 95% following the third replating; DH mPML-PRflox/ ER-T2-Cre bone marrow cells that were not exposed to Tamoxifen, and bone marrow cells from wild type mice, could not be serially replated. Cells bearing a floxed PML-RARα allele also expanded in vivo. In DH mPML-PRflox/ER-T2-Cre mice, a single dose of Tamoxifen (4 mg) resulted in 5% of peripheral blood cells carrying a floxed PML-RARα allele on day 7, but this population expanded progressively to 40% on day 80 without further Tamoxifen exposure (n=5). Five doses of Tamoxifen (4 mg) lead to 20% peripheral blood cells caring a floxed PML-RARα allele on day 18, and this increased progressively to 80% on day 98 (n=4). Peripheral blood of DH mPML-PRflox/ER-T2-Cre mice unexposed to Tamoxifen carried undetectable or trace numbers of cells with a floxed PML-RARα allele on day 80 (n=3). In contrast, myeloid progenitors from DH mPMLPRflox/ Lysozyme M-Cre mice did not display self-renewal ex vivo or an expansion of floxed cells with successive methylcellulose replating. DH mPML-PRflox/Lysozyme M-Cre mice did develop promyelocytic leukemia with long latency (14 months) and low penetrance (7%, n=3), which could be transplanted to secondary recipients. These tumors displayed a high percentage of floxed PML-RARα alleles in peripheral blood, spleen and bone marrow cells (range 50 – 95%) and possessed 4.5 fold higher expression levels of PML-RARα mRNA than our previously characterized Cathepsin G PML-RARα knock-in tumors. Importantly, non-leukemic,18 month-old DH mPML-PRflox/Lysozyme M-Cre mice displayed little evidence of PML-RARα dependent clonal expansion. In these mice, peripheral blood cells and spleen cells maintained low levels of the floxed PML-RARα allele (10–20%), equivalent to 6 week-old mice (n=36). These data suggest that PML-RARα expression by the murine PML locus leads directly to a myeloid self-renewal program and clonal expansion. Since Lysozyme M-Cre is expressed at low levels in early myeloid progenitors, the low penetrance of leukemia and rare clonal expansion in DH mPML-PRflox/Lysozyme M-Cre animals suggests that APL leukemogenesis may require PML-RARα expression in an early myeloid progenitor compartment, rather than a late compartment.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 793-793
Author(s):  
Hamza Celik ◽  
Kramer C Ashley ◽  
Martens Andy ◽  
Elizabeth Eultgen ◽  
Cates Mallaney ◽  
...  

Abstract Despite the increasing availability of targeted therapies for myeloproliferative neoplasms (MPNs), there remains a subset of these patients that transform to secondary acute myeloid leukemia (sAML). MPN patients who develop sAML have a dismal outcome, with a median survival of six months. The mechanisms and pathways that contribute to transformation from MPN to sAML have not been well delineated. The most commonly mutated genes found in the MPN initiating clones include JAK2, MPL and CALR. Transformation to sAML however requires acquisition of additional co-operating mutations such as TET2, IDH1/2, and NRAS. Recent genome sequencing studies identified deletions of JARID2, a gene associated with the Polycomb Repressive Complex 2 (PRC2) involved in implementing global H3K27me3 in post-MPN sAML. Mutations in JARID2 are found only in the blast phase of MPNs, but not in chronic phase of the disease. This data suggests that a JARID2 deletion could be a sAML-specific transforming event by acting as a tumor suppressor in HSCs. To investigate the role of Jarid2 as a tumor suppressor, we utilized an inducible mouse model of the prototypical MPN driver mutation Jak2V617F. We have established our model system by crossing Mx1-CRE:Jarid2fl/fl (Jarid2KO) or Mx1-CRE:Jarid2fl/+ (Jarid2HET) with JAK2V617F mice to generate a Mx1-CRE:Jarid2fl/fl Jak2V617F/+ or Mx1-CRE: Jarid2fl/+Jak2V617F/+ strain. Mx1-CRE mediates both activation of Jak2V617Fand deletion of Jarid2 simultaneously in adult hematopoietic compartment upon injection of the double-stranded RNA analog polyinosinic:polycytidylic acid (pIpC). In all cases, the absence of Jarid2 in Jak2V617F/+ background accelerated MPN progression, characterized by elevated hemoglobin and hematocrit, increased red blood cells, leukocytosis, thrombocytosis, and splenomegaly compared to the control groups. Median survival of Jarid2KO-Jak2V617F/+ and Jarid2HET-Jak2V617F/+ strains also revealed a dose-dependence of Jarid2 on life expectancy with a median of 27 and 56 days post pIpC treatment, respectively (Figure 1). Together, these data suggest that loss of Jarid2 in Jak2V617F/+ background accelerates the progression of MPN. To more completely understand the tumor suppressor function of Jarid2 in MPN, we aimed to define its role in normal hematopoiesis. Jarid2 is highly expressed in myeloid-biased compared to lymphoid-biased HSCs, suggestive of a role in HSC subtype determination. Moreover, conditional knock-out studies show that each core component of PRC2 (EED, SUZ12 and EZH2) has distinct as well as overlapping functional properties in hematopoiesis. To study the function of Jarid2 in normal hematopoiesis, we crossed Jarid2fl/fl mice to the Vav-CRE strain to facilitate conditional inactivation of Jarid2 in hematopoietic cells. Vav1-CRE is expressed throughout life in definitive HSCs and all hematopoietic lineages starting at E10.5. Analysis of eight-week old Vav1-CRE:Jarid2fl/fl mice showed that complete loss of Jarid2 induced a significant compromise in hematopoiesis with a robust reduction in phenotypically-defined bone marrow HSCs, a defective B-cell generation in the bone marrow (BM), a differentiation block in T-cell development in thymus, and a significant reduction in peripheral blood counts. A competitive transplantation strategy was also employed to assess the stem cell potential of Jarid2-KO HSCs. 2.5 x 105 whole bone marrow cells isolated from adult mice were transplanted into lethally irradiated recipient mice along with 2.5x105 whole bone marrow cells from congenic wild-type mice. Peripheral blood analysis of these mice over the period of 16-weeks post-transplant has shown that the loss of Jarid2 disrupts HSC function, leading to enhanced myeloid and reduced lymphoid output. Collectively, these data illustrate that Jarid2 is required for hematopoietic hemostasis including appropriate lineage fate determination of HSCs. The loss of Jarid2 in a Jak2V617F background promotes acceleration of MPN and implicates Jarid2 as a hematopoietic tumor suppressor. Figure 1. Kaplan-Meier analysis of a cohort of Jarid2KO-Jak2V617F (n = 5) and Jarid2HET -Jak2V617F (n = 6) and littermate controls (n = 4-8 each). Figure 1. Kaplan-Meier analysis of a cohort of Jarid2KO-Jak2V617F (n = 5) and Jarid2HET -Jak2V617F (n = 6) and littermate controls (n = 4-8 each). Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2013 ◽  
Vol 122 (21) ◽  
pp. 110-110
Author(s):  
Takahisa Tomioka ◽  
Goro Sashida ◽  
Kotaro Shide ◽  
Kazuya Shimoda ◽  
Naoto Yamaguchi ◽  
...  

Abstract Polycomb group proteins are transcriptional repressors that epigenetically regulate transcription via histone modifications. There are two major polycomb-complexes, the Polycomb Repressive Complexes (PRC1 and PRC2). PRC2 contains SUZ12, EED, and EZH2 that catalyze the trimethylation of histone H3 at lysine 27 (H3K27me3) and silence target-genes expression. EZH2 is generally thought to act as an oncogene in lymphoma by silencing tumor suppressor genes through H3K27me3 modifications. However, loss-of-function mutations of EZH2 have been found in myeloid malignancies such as MDS and MPN including primary myelofibrosis (PMF). In a recent study, EZH2 mutations were independently associated with shorter survival in PMF patients, suggesting that EZH2 functions as a tumor suppressor in PMF. Although JAK2V617F mutant is found in approximately 50% of PMF patients, it remains obscure whether the presence of JAK2V617F mutant predicts survival of PMF patients, and the functional contribution of JAK2V617F to the development of PMF has not been fully delineated. JAK2 has been shown to directly phosphorylate H3Y41 (H3Y41p) and reduce HP1a binding, thereby activating expression of target genes. However, it is unknown how JAK2V617F epigenetically alter expression of target genes in the development of PMF. Given that JAK2V617F mutation is significantly associated with EZH2 mutations in PMF patients, in order to understand how EZH2 mutations contribute to the pathogenesis of JAK2V617F-positive PMF, we generated a novel mouse model of PMF utilizing H2K-JAK2V617F transgenic mice and Ezh2 conditional knockout mice. We first harvested 5x106 bone marrow cells from tamoxifen-inducible Cre-ERT;Ezh2wild/wild (WT), Cre-ERT;Ezh2flox/flox (Ezh2 cKO), JAK2V617F TG/Cre-ERT;Ezh2wild/wild (JAK2 TG) and JAK2V617F TG/Cre-ERT;Ezh2flox/flox (JAK2 TG/Ezh2 cKO) mice, and then transplanted into lethally irradiated recipient mice. At 4 weeks post transplantation, we deleted Ezh2 via administration of tamoxifen, and observed disease progression until 9 months post transplantation. WT and Ezh2 cKO mice did not develop myeloid malignancies. While all 11 JAK2 TG mice died due to PMF-like disease after a long latency as previously reported, 10 out of 10 JAK2 TG/Ezh2 cKO mice immediately developed PMF and died by approximately 50 days post-deletion of Ezh2. JAK2 TG/Ezh2 cKO mice showed a significantly shorter median survival than did JAK2 TG mice (36.5 days versus 245 days, p<0.01). In the peripheral blood, moribund JAK2 TG/Ezh2 cKO mice showed increased mature neutrophils, severe anemia, and thrombocytopenia, compared to WT or JAK2 TG mice at 2 months post transplantation. At the time of sacrifice, JAK2 TG/Ezh2 cKO mice showed a significant hypoplastic bone marrow without an increased myeloblast cells, but also had a marked splenomegaly due to infiltration of myeloid cells compared to JAK2 TG mice. In addition, JAK2 TG/Ezh2 cKO mice showed a severe myelofibrosis in both bone marrow and spleen, indicating that Ezh2 loss obviously promotes JAK2 V617F-driven PMF in vivo. To understand a molecular mechanism how Ezh2 functions as a tumor suppressor for PMF, we performed gene expression analysis in Lin-Sca1+c-Kit+ (LSK) cells. While Ezh2 cKO LSKs and JAK2 TG LSKs showed up-regulation (>2-fold) of 1044 and 861 genes, respectively, JAK2 TG/Ezh2 cKO LSKs showed up-regulation (>2-fold) of more genes (1306), compared to WT LSKs. As expected, H3Y41p and H3K27me3 target genes were significantly upregulated in JAK2 TG/Ezh2 cKO LSKs, whereas H3K27me3 targets were significantly repressed in JAK2 TG LSKs, consistent with the tumor suppressor role of Ezh2 in PMF. We are now working to understand how dysregulated genes are involved in the progression of JAK2V617F-induced PMF after deletion of Ezh2. In conclusion, we have successfully established the progressive PMF in mice reconstituted with Ezh2 null cells expressing JAK2V617F mutant, and demonstrated that Ezh2 functions as a tumor suppressor in this context. This model can be utilized for innovating new therapies for PMF. Disclosures: No relevant conflicts of interest to declare.


Blood ◽  
2004 ◽  
Vol 104 (11) ◽  
pp. 3548-3548
Author(s):  
Kentaro Kinjo ◽  
Deepa B. Shankar ◽  
Jerry Cheng ◽  
Samuel Esparza ◽  
Noah Federman ◽  
...  

Abstract CREB or cAMP responsive element binding protein is a 43-kDa-basic/leucine zipper (bZip) transcription factor that regulates gene expression through the activation of cAMP-dependent or -independent signal transduction pathways. CREB promotes growth and survival in a variety of cell types and is overexpressed in the bone marrow of greater than 60% of AML patients. To understand the role of CREB in myelopoiesis, we characterized the effects of CREB overexpression in transgenic mice. We created mice in which CREB expression was targeted to the myeloid lineage using the hMRP8 promoter. CREB transgenic mice showed evidence of monocytosis, compared to age-matched littermate controls. We performed colony assays with methylcellulose containing SCF, IL-6, and IL-3. Bone marrow cells from CREB transgenic mice formed robust colonies earlier and had increased numbers of colony forming units (CFU-GM) when compared to control mice. Cytospin analysis of these cells showed the presence of more immature myeloid cells compared to controls. At day 12, cells from colonies were 50% c-Kit positive, 83% Gr-1 positive, and 67% Mac-1 positive by FACS analysis. To assess self-renewal of progenitors from CREB transgenic mice, serial replating experiments were performed. Bone marrow cells from transgenic mice were highly successful in repopulating the methylcellulose containing SCF, IL-6, and IL-3, in contrast to the control cells, which were unable to grow after serial replating. Following tertiary replating of the CREB transgenic mouse bone marrow, we observed that the colonies (96+3.5) appeared more homogeneous with immature cells that were >99% c-Kit positive and <1% GR-1, Mac-1 positive. These results suggest that persistent expression of CREB leads to a blast-like phenotype in the absence of differentiation. To determine whether increased CREB expression confers growth factor-independence, we cultured bone marrow cells in methylcellulose that did not contain cytokines. We observed a 10-fold increase in the numbers of cells from CREB transgenic mice (two different founder lines) compared to normal bone marrow. When cultured in methylcellulose containing M-CSF, the bone marrow cells from CREB transgenic mice formed larger and significantly greater numbers of colonies. However, these cells did not grow in the presence of G-CSF or EPO alone. To determine if the myeloproliferative (monocytosis) phenotype was transplantable into wild type recipient mice we injected 4x106 bone marrow cells from CREB transgenic mice into wild type C57/BL6 recipient mice. Serial analysis of the peripheral blood counts and cell surface markers by FACS analysis showed earlier myeloid engraftment at 6 weeks following transplantation compared to normal control mice. The transgenic recipients showed increased monocytes and neutrophils in the peripheral blood with a corresponding increase in Mac-1 positive, Gr-1 positive cell populations at 8 weeks after transplantation. At the same time, the lymphocyte count was significantly lower in CREB transgenic recipient mice than controls. Our results suggest that CREB plays a critical role in the regulation of normal hematopoiesis and stem cell self-renewal.


Blood ◽  
2009 ◽  
Vol 114 (22) ◽  
pp. 3475-3475
Author(s):  
John S. Welch ◽  
Timothy Ley

Abstract 3475 Poster Board III-412 Acute promyelocytic leukemia (APL) is characterized by the t(15;17) translocation, which leads to expression of the fusion protein that initiates this disease (PML-RARA), the creation of a small and inconsistently expressed RARA-PML fusion protein, and haploinsufficiency for both RARA and PML. While alternative translocations and fusion proteins have been associated with APL, RARA appears to be a necessary member of each (e.g. PLZF-RARA, NPM-RARA, STAT5b-RARA, NuMA-RARA). Furthermore, RARA activity is important for myeloid maturation. We have therefore explored the effect of RARA haploinsufficiency in the murine Cathepsin-G PML-RARA (mCG-PR) model of APL. We crossed RARA+/− mice with mCG-PR mice, all on a C57/B6 background. Both mCG-PR and the mutant RARA allele were observed at normal Mendelian ratios. At 8 weeks, mCG-PR x RARA+/− mice exhibited normal peripheral blood counts, spleen sizes and total bone marrow cells. Bone marrow cells from 8-week old mCG-PR x RARA+/− and mCG-PR mice both exhibited increased self-renewal in methylcellulose replating assays, with increased average cells per colony in mCG-PR x RARA+/− and mCG-PR colonies compared to wild type (28,470 ± 6,000 and 15,400 ± 2,375 vs 7,700 ± 630, p = 0.0001 and 0.0003 following the initial plating). Five months after bone marrow transplantation at a 1:9 ratio with competitor Ly5.1 bone marrow cells, mCG-PR derived cells have engrafted and expanded in four recipients to 10%, 12.4%, 13.8% and 15.2% of peripheral blood cells, while mCG-PR x RARA+/− derived cells have more robustly expanded to 8.7%, 20.2%, 24.5%, and 30.3% of peripheral blood cells. A cohort of 29 mCG-PR x RARA+/− and 20 mCG-PR mice was subjected to a tumor watch. With an average follow up of 10 months, we have observed AML arise in 9 mCG-PR x RARA+/− mice and 11 mCG-PR mice (an additional 20 and 9 mice in each respective cohort remain at risk of leukemia and will be subsequently evaluated). Leukemia arising from mCG-PR mice exhibited leukocytosis, splenomegaly and variable myeloid differentiation arrest in the peripheral blood and spleen, as measured by manual differential counts and cKit/CD11b expression in the Gr1+ compartment. Leukemia from two mice had marked promyelocytic maturation arrest, but 5 others retained differentiation, with 30-60% bands and ring-formed neutrophils in the spleen, which correlated with similar percentages of Gr1+/cKit-/CD11b+ cells (4 mice died with splenomegaly noted during necropsy and were not further characterized). AML arising in mCG-PR x RARA+/− mice displayed similar variability in maturation arrest, but these mice had a trend towards lower peripheral blood WBC counts compared to mCG-PR AML (mean WBC 28,000/mcl ± 26,000 vs 82,000/mcl ± 71,000/mcl, p = 0.06) and larger spleen size (mean 1,425 mg ± 475 mg vs 1,017 mg ± 193 mg, p = 0.03). In addition, 6/9 mCG-PR x RARA+/− mice with AML had marked cervical lymphadenopathy caused by infiltrating AML cells. Only 1/11 mCG-PR mice with AML displayed lymphadenopathy, which previously has been observed only rarely with this mouse model (Westervelt et al, Blood 2003). Both mCG-PR x RARA+/− and mCG-PR AML cells responded to ATRA with increased maturation in vitro and in vivo, a shift from cKit+/CD11b- to cKit-/CD11b+ expression, a loss of self renewal capacity, and improved survival compared to untreated controls, as expected. In sum, mCG-PR x RARA+/− mice are different from mice with mCG-PR alone. RARA haploinsufficiency further perturbs hematopoiesis in mCG-PR mice, accelerating an early competitive advantage conferred by PR expression, and possibly altering the pattern of homing and peripheralization of AML cells. Lower peripheral white blood cell counts in these animals more accurately reproduce the relatively low white counts of most patients with APL. These data strongly suggest that RARA haploinsufficiency contributes to the overall phenotype of APL initiated by PML-RARA, and thus is a relevant mutation created by the t(15;17) translocation. Disclosures: Welch: Cephalon: Research Funding; Eisai: Research Funding.


Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 551-551
Author(s):  
Zeenath Unnisa ◽  
Jason P Clark ◽  
Elizabeth Wojtowicz ◽  
Lino Tessarollo ◽  
Neal G. Copeland ◽  
...  

Abstract Abstract 551 Normal hematopoiesis is maintained by long-term hematopoietic stem cells (LT-HSCs) that are defined by their extensive self-renewal and multipotency. Self-renewal of LT-HSCs in turn is regulated by a complex network of intrinsic and extrinsic factors. The transcription factor MEIS1 is highly expressed in hematopoietic stem and progenitor cells and also in several leukemias, suggesting that MEIS1 might be important in regulating self-renewal. However, the role of MEIS1 in normal hematopoiesis has not been defined. To determine the role of MEIS1 in hematopoiesis, we studied conditional knockout mice. We generated transgenic mice bearing loxp sites flanking the homeodomain of MEIS1. The MEIS1-floxed mice were then bred to Rosa26-CreERT2 mice, the latter expressing cre-recombinase ubiquitously, that can be activated by estrogen or its analog Tamoxifen (Tam). Efficient, complete recombination was achieved in vivo by treating MEIS1-f/f-Cre (homozygous for MEIS1-flox) mice with Tam and in vitro by treating bone marrow cells with 4-hydroxy tamoxifen. Loss of MEIS1 expression was detected by QRT-PCR and western blotting. To determine the role of MEIS1 in the maintenance of adult hematopoiesis, MEIS1-f/f-Cre and control mice were treated with Tam and MEIS1 deletion confirmed by PCR. At three weeks post deletion, bone marrow analysis showed a significant reduction in the number of LT-HSCs defined as lin-/c-Kit+/Sca1+/CD48−/CD150+ in the MEIS1-depleted mice compared to controls (0.012% compared to 0.037%, N=6, p<0.05, t-test). However, the progenitor populations were unaffected by MEIS1 deletion. Over a period of 12 weeks of observation, the mice did not show any signs of distress and the peripheral blood counts of the experimental and control mice remained normal, indicating that short term hematopoiesis was not affected. Cell cycle analysis of LT-HSCs showed that MEIS1 deletion resulted in a significant shift of cells from G0 to G1 phase (G0 and G1 proportions respectively, 81.75±3.25% and 9.40±3% for control and 56.10±0.873% and 31.17±1.5% for MEIS1-deleted). To determine the effects of MEIS1 loss on intrinsic hematopoietic stem cell function, we performed competitive repopulation assays. Bone marrow cells harvested from MEIS1-f/f-Cre or MEIS1-f/+-Cre (control) mice were combined with equal numbers of bone marrow cells from BoyJ mice and transplanted via tail vein injection into lethally irradiated BoyJ mice. Four weeks after transplant, recipients were treated with Tam or vehicle for 5 days and deletion of MEIS1 confirmed by PCR on peripheral blood. Peripheral blood of recipient mice was analyzed at 1, 4, 8, 12 and 16 weeks after treatment and relative chimerism assessed by flow cytometry. At 1 and 4 weeks after treatment, the chimerism in the MEIS1 deleted group (Tam treated MEIS1-f/f-CreER) and the control groups (Tam treated MEIS1-f/+-CReER and vehicle treated MEIS1-f/f-CreER) was comparable (41%, 40.5% and 41.5% respectively, average, N=5 to 8). However, by 8 weeks after treatment, the MEIS1 deleted group showed a significant decline in chimerism compared to controls (18.2% compared to 43.1% and 35.1% respectively, p<0.02, t-test) and at 16 weeks the chimerism in the MEIS1-deleted group declined further (11.1% compared to 40.2% and 35.0% respectively, p<0.001). Subpopulation analysis showed loss of chimerism in granulocytes and in B and T lymphocytes. The latency and breadth of the effect of MEIS1 loss suggested an effect on the hematopoietic stem cell population. Indeed, bone marrow analysis of transplant recipients showed near complete loss of LT-HSC chimerism (3% compared to 70.25% and 75.6% respectively, p<0.001). Finally, we performed gene expression profiling on lineage negative bone marrow cells with and without MEIS1 deletion. Results showed that loss of MEIS1 was associated with decreased expression of hypoxia-responsive genes. Collectively, these results indicate that MEIS1 is required for the maintenance of the pool of LT-HSCs. Loss of MEIS1 promotes cycling and exhaustion of LT-HSCs. Further, we propose that activation of the hypoxia-response pathway may be one of the mechanisms by which MEIS1 exerts its effects on hematopoietic stem cells. Disclosures: No relevant conflicts of interest to declare.


Sign in / Sign up

Export Citation Format

Share Document