Krüppel-like Factors KLF4 and KLF2 Regulate microRNA-150 Expression in Myeloid Leukemias

Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 874-874
Author(s):  
Valerie A. Morris ◽  
Carrie Cummings ◽  
Brendan Korb ◽  
Sean M. Boaglio ◽  
Vivian Oehler

Abstract Background: Acute myeloid leukemia (AML) is characterized by increased self-renewal of leukemia stem/progenitor cells and failure of differentiation to mature myeloid cells. MicroRNAs (miRNAs) are small single stranded non-coding RNAs 19 to 24 nucleotides in length that regulate expression of tens to hundreds of genes via mRNA degradation or translational repression. MiRNA contributions to normal hematopoiesis have been described and deletion of key miRNA processing enzymes in murine and human cells suggests that miRNA loss contributes to the cancer phenotype and aberrant differentiation in leukemia. By combining observations of miRNA expression in normal hematopoietic progenitor cells and patient AML cells and high-throughput lentiviral expression library screening approaches in AML cell lines we have identified candidate miRNAs that contribute to altered proliferation and differentiation in AML cells. We have previously established 1) that miR-150 expression is decreased in a large subset of primary patient AML samples, in particular poor risk cytogenetic groups, 2) and that miR-150 re-expression induces myeloid differentiation and decreases cell proliferation of normal hematopoietic progenitor cells and AML cell lines and primary patient cells in part through downregulation of MYB expression. MiR-150 loss is relevant in other hematopoietic and solid tumor malignancies where re-expression inhibits cell proliferation, promotes apoptosis and induces reversal of endothelial to mesenchymal transition. Transcription factors are important regulators of myeloid differentiation and cell proliferation. Moreover, as highlighted by recent sequencing of the AML genome, alterations in myeloid transcription factors through mutation, gene rearrangement, and altered expression play a significant role in leukemogenesis. Consequently, we have focused on how myeloid transcription factors regulate miRNA expression, specifically for miR-150. Results: Using 5’RACE from healthy bone marrow RNA, we identified a major transcription start site at 214 basepairs upstream of the pre-miR-150 hairpin. We identified the minimal miR-150 promoter region as -266 to +259 basepairs from the major transcription start site using miR-150 promoter truncation luciferase constructs assayed in myeloid leukemia cell lines (THP-1, K562, and KG1a) and a lymphoid leukemia cell line (Jurkat). We identified DNA binding sites for the Krüppel-like factor (KLF) family of transcription factors that are necessary for miR-150 promoter activity using site-directed DNA mutagenesis of the luciferase reporters. KLFs regulate proliferation, differentiation, pluripotency, migration and inflammation. Depending on cell type and context, KLFs can function as tumor suppressors or oncogenes. To identify which KLF isoforms regulate miR-150 expression, we assayed the ability of KLFs 2, 3, 4, 5, 6, 7, 9, and 10 to induce miR-150 promoter activity using the luciferase reporters and endogenous miR-150 expression by quantitative PCR. KLF2 and KLF4 overexpression increased miR-150 promoter activity in luciferase assays 50-fold and 450-fold respectively in K562 cells. Furthermore, KLF2 and KLF4 induced endogenous miR-150 expression 20-fold and 100-fold respectively as detected by quantitative PCR in both THP-1 and K562 cells. Prior work has established that KLF2 and KLF4 regulate the differentiation of monocytes. We then confirmed that KLF2 and KLF4 overexpression promotes myeloid differentiation of THP-1 cells by flow cytometry and gene expression that was partially reversed by inhibition of miR-150 expression. Conclusions: Previous studies have determined that KLF2 and KLF4 expression are decreased or absent in a significant subset of AML cases. Our observations suggest that loss of KLF2 and KLF4 expression contributes to decreased miR-150 expression which in turn alters cell proliferation and differentiation. Other studies have implicated the cell cycle inhibitor p21WAF1/CIP1 and altered PPAR gamma signaling downstream of KLF4. Nonetheless, our mechanistic understanding is limited. Our work suggests that the loss of miR-150 and other miRNAs downstream of these transcription factors also contributes. Understanding the interactions between KLFs, miR-150 and other miRNAs has broader significance as KLF2 and KLF4 expression is altered in other hematopoietic and solid tumors. Disclosures No relevant conflicts of interest to declare.

Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 4341-4341
Author(s):  
Mingying Zhang ◽  
Fangnan Xiao ◽  
Yunan Li ◽  
Zizhen Chen ◽  
Xiaoru Zhang ◽  
...  

Abstract Introduction: Chemoresistance and disease relapse remain the main obstacles responsible for treatment failure in leukemia. MicroRNAs (miRNAs) play essential roles in various physiological and pathological processes, including cell proliferation, differentiation, metabolism, and cancer development. The miR-106b-25 cluster consists of three miRNAs: miR-106b, miR-93 and miR-25. We have previously reported that miR-106b-25 was associated with chemoresistance by negatively regulated EP300 in breast cancer, but its role in hematological malignancies has not yet been elucidated. Here, we aim to clarify the biological role and underlying mechanisms of miR-106b-25 on drug resistance in leukemia. Methods: To see whether the miR-106b-25 was associated with the poor prognosis of AML patients, enriched LSCs (CD34 + cells) were isolated from the bone marrow of 18 newly diagnosed AML patients, the expression of miR-106b, miR-93, and miR-25 were examined, respectively. The expression levels of miR-106b, miR-93 and miR-25 were further determined in the doxorubicin-resistant leukemia cell line K562/A02 and HL60/ADR, compared with their parental cell lines. In addition, K562 cells were transduced with lentiviral vectors carrying miR-106b-25, and cell proliferation, drug resistance, colony-forming assay, apoptosis assays were performed to explore the function of miR-106b-25 overexpression in leukemia cells in vitro. To investigate the role of miR-106b-25 on tumor growth and overall survival after drug treatment, we performed xenotransplantation in nude mice using miR-106b-25 overexpressed K562 cells. To further clarify the function of each microRNA function in this cluster, K562 cells were also transduced with lentiviral vectors carrying individual miR-25, miR-93, or miR-106b separately. Cell proliferation, colony forming assay and cell apoptosis assay were also carried out subsequently. Simultaneously, RNA-sequencing was performed to reveal the underlying mechanisms of miR-106b-25 in the chemoresistance of myeloid leukemia. To experimentally confirm the direct target of the miR-106b-25 cluster in AMLs, we further performed a dual-luciferase reporter assay. Results: Upregulated miR-106b, miR-93 and miR-25 expression in enriched LSCs were significantly associated with shortened overall survival of AML patients. We also found miR-106b, miR-93 and miR-25 were significantly upregulated in drug-resistant leukemia cell lines compared with its parental cell lines. Overexpression of miR-106b-25 cluster promoted cell proliferation, led to resistance of K562 cells to doxorubicin, imatinib and ABT-737 (BCL-2 inhibitor) in liquid culture and drug-resistant colony-forming assays. Overexpression of miR-93 or miR-106b accelerated cell growth, and all the three miRNAs can promote drug-resistant colony-forming and inhibit cell apoptosis. RNA-sequencing (RNA-Seq) data revealed that multiple critical genes related to apoptotic pathways were downregulated after overexpressing miR-25, miR-93, miR-106b as well as the whole cluster, such as TP73, BAX, BAK1, Caspase-7, CDKN1A and BTG2. RT-qPCR confirmed that these genes are reduced with or without ABT-737 treatment. Luciferase assay further identified TP73 was a direct target of miR-93 and miR106b, BAK1 was a direct target of miR-25, and CASPASE-7 was a direct target of all these three miRNAs. Conclusions: In summary, we made the novel observation that miR-106-25 is associated with AML drug-resisitance and disease prognosis and identified TP73, BAK1 caspase-7 as a novel direct target of this cluster. Further studies revealed that the biological effects of miR-106b-25 cluster on leukemic cell proliferation, chemoresistance and apoptosis were mediated through regulation of apoptotic pathway. These findings indicate a promising diagnostic biomarker and a potential target therapeutic strategy for AML patients. Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 5201-5201
Author(s):  
Shaowei Qiu ◽  
Jing Yu ◽  
Tengteng Yu ◽  
Haiyan Xing ◽  
Na An ◽  
...  

Abstract Introduction: As the important suprressor of P53, iASPP was found to be overexpressed in leukemia, and functioned as oncogene that inhibited apoptosis of leukemia cells. Sertad1 is identified as one of the proteins that can bind with iASPP in our previous study by two-hybrid screen. Sertad1 is highly expressed in carcinomas from pancreatic, lung and ovarian tissues, which considered Sertad1 as an oncoprotein. In this study, our findings revealed that Sertad1 could interact with iASPP in the cytoplasm near nuclear membrane, which could block iASPP to enter into nucleus to interact with P53, and inhibited the function of iASPP eventually. Methods: Co-immunoprecipitation and fluorescence confocal microscopic imaging were used to confirm the interaction between iASPP and Sertad1, the exact binding domains and the subcellular colocalization.The plasmids of iASPP and Sertad1 were transfected alone or co-transfected into K562 cells, the stable subclones that highly expressed iASPP, Sertad1 or both of them were then established by limiting dilution and named as K562-iASPPhi, K562-Sertad1hi, and K562-Douhi, respectively. The cell proliferation, cell cycle and apoptosis of above subclones were investigated by flow cytometry. Further, silence of the above two proteins was performed to confirm their functions. Immunoblotting analysis and immunofluorescence were performed to explore the possible mechanisms of difference between the biological functions of the above subclones. Results: Sertad1 expression level varied in leukemic cell lines and AML patients irrespectively of iASPP and P53. Interaction between iASPP and Sertad1 did exist in 293 cell and leukemic cells, both iASPP and Sertad1 scattered in the cytoplasm and nucleus, and their colocalizations were mainly in the cytoplasm, which encircled the nucleus. iASPP binds directly to Sertad1 through its PHD-bromo domain, C-terminal domain and Cyclin-A domain in a reduced order, and Serta domain failed to bind to iASPP. Overexpression of iASPP in K562 cells (iASPPhi) could result in the increased cell proliferation, cell cycle arrest in G2/M phase and resistance to apoptosis induced by chemotherapy drugs. While overexpression of iASPP and Sertad1 at the same time (Douhi) could slow down the cell proliferation, lead the cells more vulnerable to the chemotherapy drugs. As figure showed, in K562-Douhi cells, both iASPP and Sertad1 were obviously located in the cytoplasm, which encircled the nuclei, the subcellular colocalization was nearly outside the nuclei. The immunoblotting analysis further supported the conclusions. The resistance of iASPP to chemotherapeutic drug was accompanied by Puma protein expression in a p53-independent manner. By knocking down the expersssion of iASPP and Sertad separately, we found that iASPP is dispensable for maintenance of anti-apoptotic function and Sertad1 is indispensable for cell cycle in leukemic cells. Conclusions: In normal situation, the protein iASPP and Sertad1 scatter in the nucleus and cytoplasm, mainly in the cytoplasm. As convinced by our study, iASPP was overexpressed in the leukemia cell lines and primary AML patients, it could function as oncogene through its binding with P53 protein in the nucleus, inhibit the function of P53. When iASPPhi cells were exposed to apoptosis stimuli, Puma protein could play an important role in this process, irrespective of the expression level of P53. But when iASPP and Sertad1 were both overexpressed in the leukemic cells, Sertad1 could tether iASPP outside the nucleus mainly through its PHD-bromo domain, prevent it from inhibiting P53 function, suppress the leukemic cell growth and stimulate cell apoptosis by rescuing the P53 eventually. Our data provided a new insight to overcome iASPP protein, namely through its binding partners, when the similar proteins or drugs that can tether iASPP outside the nucleus such as Sertad1 are transfected into the leukemic cells, it may restore p53 function to eliminate the leukemic cells. Figure 1 Figure 1. Disclosures Wang: Novartis: Consultancy; Bristol Myers Squibb: Consultancy.


Blood ◽  
2004 ◽  
Vol 104 (11) ◽  
pp. 1532-1532
Author(s):  
Suman Malempati ◽  
Rosalie C. Sears

Abstract The c-Myc oncoprotein is a key regulator of cell fate decisions including proliferation, differentiation, and apoptosis. Orderly control of c-Myc protein levels is important in maintaining regulated cell proliferation in normal cells. While c-Myc overexpression is seen in many hematopoietic malignancies, the reason for high protein levels in most cases is unknown and, in general, is not the result of translocations or gene amplification. C-Myc levels vary with cell cycle and are kept very low in quiescent cells. Protein half-life is controlled by phosphorylation at two specific N-terminal sites, Serine 62 and Threonine 58, which regulate c-Myc degradation by the ubiquitin proteasome pathway. Two Ras-dependent signaling pathways (Raf/MEK/ERK and PI(3)K/Akt) modulate phosphorylation at Serine 62, which stabilizes the protein, and Threonine 58, which targets Myc for ubiquitination and subsequent degradation. We recently reported that a stabilized form of c-Myc (c-Myc T58A) contributes to oncogenic transformation of human cells in culture (Yeh et al, Nat. Cell Bio.6:308–318, 2004). Here we describe the role of c-Myc protein stabilization in 2 pediatric ALL cell lines (REH and Sup-B15), 1 AML cell line (HL-60), and 1 CML cell line (K562). Markedly higher expression of c-Myc protein was seen in all 4 cell lines as compared to normal peripheral blood mononuclear cells (PBMCs). FISH analysis demonstrated amplification of the c-myc gene in HL-60 cells as has been previously reported, but not in REH, Sup-B15, or K562 cells. Using [35S]methionine pulse-chase analysis we demonstrate that the half-life of c-Myc in REH (55 minutes), Sup-B15 (47 minutes), and K562 (40 minutes) cells is longer than in normal PBMCs (9 to 15 minutes), but is not significantly prolonged in HL60 cells. We provide additional functional evidence for aberrant protein stabilization based on greater elevation of c-Myc protein after proteasome inhibition in PBMCs and HL-60 cells than in REH, Sup-B15, or K562 cells. These results suggest that that abnormalities in c-Myc degradation exist upstream of ubiquitination in the ALL and CML cell lines. Consistent with this hypothesis, experimental inhibition of the PI(3)K pathway knocked down c-Myc levels in REH and Sup-B15 cells, an effect that was abrogated by concomitant proteasome inhibition. This result suggests that abnormal activation of the PI(3)K pathway could participate in c-Myc stabilization in these cells. In addition, destabilization of c-Myc by PI(3)K inhibition correlated with a significant decrease in cell proliferation. In conclusion, we demonstrate that aberrant stabilization of c-Myc protein occurs in human leukemia cell lines. Affecting the c-Myc degradation pathway in hematopoietic malignancies that have stabilized c-Myc may constitute a novel therapeutic target. Additional experiments are ongoing to assess c-Myc stability in primary cells from leukemic bone marrow samples.


Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 3387-3387
Author(s):  
Teresa V. Bowman ◽  
Eirini Trompouki ◽  
Lee N Lawton ◽  
Zi Peng Fan ◽  
Dai-chen Wu ◽  
...  

Abstract Abstract 3387 BMP and Wnt signaling pathways control essential cellular responses through activation of the transcription factors SMAD (BMP) and TCF (Wnt). Here, we have evaluated their function during hematopoietic regeneration after irradiation. Using heat-shock inducible transgenic zebrafish lines that overexpress BMP2 or Wnt8, we demonstrated accelerated marrow recovery following irradiation. Heat-shock induced overexpression of the respective inhibitors Chordin and DKK1 blunted the recovery. Surprisingly, gene expression profiling after induction of BMP or Wnt signaling in zebrafish marrow cells post-irradiation revealed increased expression of the key hematopoietic genes scl, runx1, and gata2. To determine if the effect of BMP and Wnt signaling on hematopoietic genes during regeneration was direct, we performed ChIP-PCR for Smad1 and the hematopoietic regulator Gata2 in murine lineage-negative progenitors seven days after a sublethal irradiation. We found that Smad1 and Gata2 co-occupy hematopoietic genes including Cd9, Il13, Mapk6, and Meis1. To examine the binding of SMAD1 and TCF7L2 throughout the genome of hematopoietic cells, we employed ChIP-seq in human erythroid and myeloid leukemia cell lines, K562 and U937, respectively. More than 70% of the genes bound by SMAD1 and TCF7L2 were co-occupied with the lineage transcription factors GATA1 and GATA2 in erythroid cells, and with C/EBPα in myeloid cells. This finding suggests that signaling transcription factors control hematopoietic gene programs by binding DNA adjacent to lineage-specific transcription factors. The transcriptional output of BMP and Wnt activity was tested on an LMO2 enhancer reporter construct. Expression of SMAD1 or TCF7L2 alone had little effect, but markedly increased reporter activity in conjunction with GATA2, indicating that BMP and Wnt signaling cooperate with lineage regulators to enhance transcription of cell-type specific target genes. To establish the order of transcription factor occupancy, we utilized estrogen-inducible C/EBPα-ER in K562 cells or GATA1 induction in murine G1ER cell lines, and assessed SMAD1 occupancy before and after induction of each respective lineage regulator. Induction of the myeloid lineage regulator C/EBPα in K562 cells shifted binding of SMAD1, such that SMAD1 co-occupancy with C/EBPα changed from 6% to 15% of C/EBPα targets. In contrast, expression of the erythroid regulator GATA1 promoted loss of SMAD1 on 82% of its targets, and restricted more than 98% of the remaining SMAD1 sites to erythroid targets adjacent to GATA1. Co-occupancy of signaling factors and lineage regulators was further tested in primary human CD34+ multipotent hematopoietic progenitors and CD34+ cells directed to the erythroid lineage. Both SMAD1 and TCF7L2 co-localized with GATA2 on greater than 75% of bound genes in multipotent CD34+ progenitor cells. Similar to our results following GATA1 induction in G1ER cells, SMAD1 occupancy shifted to 65% erythroid targets upon differentiation of progenitors to the erythroid lineage. These data provide strong evidence that the binding of signaling factors follows the genomic occupancy of the dominant lineage regulator during differentiation. Together, our findings demonstrate that hematopoietic regeneration is driven by collaboration of master regulators and signaling transcription factors to control the entire hematopoietic program. Disclosures: Daley: Verastem, Inc: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; iPierian, Inc: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; Epizyme, Inc: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; Solasia, KK: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; MPM Capital, Inc: Consultancy, Membership on an entity's Board of Directors or advisory committees; Johnson & Johnson: Membership on an entity's Board of Directors or advisory committees. Zon:Fate Therapeutics:; Stemgent: Consultancy.


Blood ◽  
1991 ◽  
Vol 78 (11) ◽  
pp. 3012-3020 ◽  
Author(s):  
H Hirvonen ◽  
V Hukkanen ◽  
TT Salmi ◽  
TP Makela ◽  
TT Pelliniemi ◽  
...  

Abstract The myc proto-oncogenes encode nuclear phosphoproteins, which are believed to participate in the control of cell proliferation and differentiation. Deregulated expression of c-myc has been implicated in several human hematopoietic malignancies. We have studied the expression and mRNA processing of human L-myc, N-myc, and c-myc genes in a panel of human leukemias, leukemia cell lines, and normal hematopoietic cells. L-myc mRNA was expressed in three acute myeloid leukemias (AML) studied and in several myeloid leukemia cell lines. Only low expression levels were observed in adult bone marrow and in fetal spleen and thymus. The K562 and Dami leukemia cell lines showed a unique pattern of L-myc mRNA processing, with approximately 40% of L- myc mRNA lacking exon III and intron I. N-myc was expressed in five of six AML cases studied, in one of nine acute lymphocytic leukemia (ALL) cases, and in several leukemia cell lines, while c-myc mRNA was detected in all leukemias and leukemia cell lines studied. Coexpression of all three myc genes was observed in Dami and MOLT-4 cell lines and in two AMLs, and either L-myc or N-myc was coexpressed with c-myc in several other cases. These results show that in addition to c-myc, the L-myc and N-myc genes are expressed in some human leukemias and leukemia cell lines, and suggest a lack of mutually exclusive cross- regulation of the myc genes in human leukemia cells.


Blood ◽  
2006 ◽  
Vol 108 (11) ◽  
pp. 1986-1986
Author(s):  
Xuejun Zhang ◽  
Li Wen ◽  
Fuxu Wang ◽  
Ling Pan ◽  
Jianmin Luo ◽  
...  

Abstract Tumor Necrosis factor (TNF)-related apoptosis- inducing ligand (TRAIL) is a new member of TNF superfamily discovered recently. Several studies showed that TRAIL can preferentially induce apoptosis in a variety of tumor cells, while most normal cells tested do not appear to be sensitive to TRAIL. In the present study, we treated K562 and U937 leukemia cell lines with recombinant mutant human TRAIL (rmhTRAIL) alone or together with daunorubicin (DNR) to investigate the apoptosis of the treated cells and the synergistic reaction of rmhTRAIL and DNR. The normal cell line MRC-5 was used as control. The expression of four TRAIL receptors mRNA (death receptor DR4 and DR5, decoy receptor DcR1 and DcR2) in the cells lines were detected before and after the treatment by DNR. (1) AO-EB double staining and TUNEL staining were used to evaluate the morphological change of leukemia cell lines before and after the treatment. The results showed that rmhTRAIL could induce the apoptosis of leukemia cell lines and a dose-dependent manner was found in leukemia cell lines but not in MRC-5 cell lines. (2) The growth inhibition rate of leukemia cell lines induced by rmhTRAIL alone or combined with DNR was examined with MTT assays. Different concentrations of rmhTRAIL(8, 40, 200, 1000ng/mL)alone or combined with DNR(8, 40, 200, 1000ng/mL) was used. The result showed a dose-dependent growth inhibition by rmhTRAIL alone for K562- and U937-cell line (P<0.05) also, but not for MRC-5 cell line (P>0.05). The IC50 for K562 cells and for U937 cells had no statistic difference (538.80 vs 301.56ng/mL, P>0.05). In leukemia cell lines, the growth inhibition rates in combination groups were much higher than in rmhTRAIL or DNR alone groups (P<0.05), and no synergistic killing effects was found in MRC-5 cells (P<0.05). It was concluded that rmhTRAIL had synergistic effects with DNR in the growth inhibition of K562 and U937 cells. (3). To explore the antitumor mechanisms of rmhTRAIL combined with DNR, the expression level of the DR4, DR5 and DcR1, DcR2 mRNA in these three cell lines was examined by semi-quantitative reverse transcription polymerase chain reaction (RT-PCR) before and after the treatment with DNR. The high expression of DR4,DR5 mRNA in the tested cells were observed before the treatment of DNR, while very low or even undetectable expression level of DcR1 and DcR2 mRNA were observed in U937 and K562 cells, and a high expression level of DcR1 and DcR2 mRNA in MRC-5 cells were observed. After 24 hours treatment of three cell lines with DNR (200ng/ml), the expression level of DR5 mRNA increased in K562 and U937 cells (P<0.05). DR4 mRNA also increased in K562 cells but not in U937 cells. There was no change in DcR1 and DcR2 mRNA level in three cell lines. The four receptors’ mRNA level in MRC-5 cells was not influenced by DNR. Our results indicated that rmhTRAIL could induce the apoptosis of leukemia cell lines, and DNR could enhance significantly the sensitivity of K562 and U937 cells to apoptosis induced by rmhTRAIL through up-regulation of death receptors. Therefore, we presumed TRAIL might be act as a new agent for biological therapy in leukemia.


Blood ◽  
2013 ◽  
Vol 122 (21) ◽  
pp. 1558-1558
Author(s):  
Pavel Burda ◽  
Nikola Curik ◽  
Nina Dusilkova ◽  
Giorgio L Papadopoulos ◽  
John Strouboulis ◽  
...  

Abstract Introduction Myelodysplastic syndrome (MDS) is often manifested by anemia due to ineffective erythropoiesis. Upon transformation to MDS/AML the uniform population of leukemic blasts overgrow dysplastic bone marrow. Hematopoiesis is regulated by transcription factors GATA-1 and PU.1 that interact and mutually inhibit each other in progenitor cells to guide multilineage commitment and subsequent lineage differentiation. Expression of PU.1 is controlled by several transcription factors including PU.1 itself at distal URE enhancer. It has been well established that underexpression of PU.1 in progenitor cells leads to AML (Rosenbauer F et al. 2004). In addition, co-expression of PU.1 and GATA-1 in AML-erythroleukemia (EL) blasts prevents induction of differentiation programs regulated by these transcription factors. In our laboratory, we recently observed that MDS/AML erythroblasts display repressive histone modifications and methylation status of PU.1 gene that respond to 5-azacitidine leading to inhibited blast cell proliferation and stimulated myeloid differentiation (Curik N et al. 2012). Inhibition of transcriptional activity of PU.1 protein by GATA-1 has been reported (Nerlov C et al. 2000) however it is not known whether GATA-1 can inhibit PU.1 gene in human early erythroblasts directly. Hypothesis GATA-1 inhibits PU.1 levels directly and modulates its transcriptional outcome in early erythroblasts. We also hypothesize that GATA-1-mediated repression of PU.1 transcription is delayed and this may play a role in ineffective erythropoiesis. Material and Methods Cell lines: MDS-derived OCI-M2 EL and other two human ELs (HEL, K562) and one murine EL (MEL); all co-expressing GATA-1 and PU.1. Patients: MDS patients (N=5) with rather advanced disease; MDS/AML (4) and RAEBI (1). Four received AZA; response: PR (2), SD (2) with HI. Median OS>24 Mo. For chromatin immunoprecipitation (ChIP) analysis either cell lines or CD19/CD3-depleted bone marrow cells were used. Results Direct association of GATA-1 with PU.1 gene was demonstrated in all three human ELs using ChIP. Occupancy of GATA-1 was detected upstream the PU.1 promoter and distally at GATA-1 binding sites or at PU.1 binding sites together with PU.1. Comparable data documenting occupancy of GATA-1 at PU.1 gene were observed also in MEL cells and in normal murine fetal erythroblasts using ChIP-sequencing. To test how GATA-1 regulates PU.1 expression we overexpressed GATA-1 in erythroblasts and tested expression of PU.1, histone H3 modification (near GATA-1 occupancy) and cell growth. We found that GATA-1 inhibited PU.1 expression, facilitated enrichment of repressive modifications at PU.1 gene (H3K9Me, H3K27Me) while depleted activation modifications (H3K9Ac, H3K4Me), and also inhibited cell growth. Next, we tested effects of GATA-1 knockdown using siRNA. Indeed, inhibition of GATA-1 expression in erythroblasts leads to increase in PU.1 level as well as of its targets (CEBPA, MAC1). Using Luciferase assay we confirmed that both endogenously produced PU.1 and GATA-1 are capable to stimulate exogenously inserted reporters. Next, we compared chromatin structure of PU.1 gene between data from ELs, normal controls and high risk MDS. Our data revealed that PU.1 gene in MDS is enriched with repressive modifications (H3K9Me, H3K27Me) while depleted with activation modifications (H3K9Ac, H3K4Me) suggesting defects in dynamic regulation of PU.1 expression in MDS. Conclusion Our data from ELs provide a) evidence of GATA-1-mediated repression of PU.1 gene in erythroblasts and that b) manipulation of GATA-1 affected PU.1 level in opposite direction. In high risk MDS, the chromatin structure of PU.1 gene displays accumulation of repressive epigenetic marks that are responsive to AZA. We think that during early erythroid differentiation GATA-1 binds and represses PU.1 gene, however this is not fully completed in MDS and therefore erythroid differentiation is not efficient. Grants: P301/12/P380, P305/12/1033, NT14174-3/2013, UNCE204021, FR-TI2/509, SVV-2013-266509, PRVOUK-P24/LF1/3 Disclosures: No relevant conflicts of interest to declare.


Sign in / Sign up

Export Citation Format

Share Document