GLORIA

GEOMAR Library Ocean Research Information Access

Your email was sent successfully. Check your inbox.

An error occurred while sending the email. Please try again.

Proceed reservation?

Export
  • 1
    In: Blood, American Society of Hematology, Vol. 132, No. Supplement 1 ( 2018-11-29), p. 2617-2617
    Abstract: #Michael Heuser and Anuhar Chaturvedi share senior authorship Background: Isocitrate dehydrogenase-1 (IDH1) is mutated in about 6% of AML patients. Mutant IDH produces R-2-hydroxyglutarate (R-2HG), which induces histone and DNA hypermethylation through inhibition of epigenetic regulators, thereby linking metabolism to tumorigenesis. We recently reported that at comparable intracellular R-2HG levels, mice receiving transplants of IDH1 mutant cells died significantly earlier than R-2HG treated mice in the context of HOXA9 overexpression. This suggests oncogenic functions of mutant IDH1 beyond R-2HG production. We employed a splice variant of mutated IDH1 that does not produce R-2HG (IDH1mutantΔ7) to decipher R-2HG independent signaling pathways that may contribute towards leukemogenesis. Methods: Bone marrow cells from mice were immortalized with HoxA9, and IDH1wildtype (IDH1wt), IDH1mutant (IDH1mut), IDH1wildtypeΔ7 (IDH1wtΔ7) and IDH1mutΔ7, were constitutively expressed and the leukemogenic potential was evaluated in vivo. Intracellular R-2HG was measured by enantiomer-specific quantification. Deletion of exon 7 from IDH1mut leads to a frameshift that creates a premature stop codon in the 9th exon, finally producing a 119 amino acids truncated protein, IDH1mutΔ7. This splice variant does not produce increased levels of R-2HG. The signaling pathways were explored by immunoblotting and immunofluorescence. Results: Mice receiving cells with IDH1mutΔ7 had the same short latency to leukemia as mice receiving cells with full-length mutant IDH1, while IDH1wt and IDH1wtΔ7 cells died with significantly longer latency. The WBC count increased over time in IDH1mutΔ7 mice similar to IDH1mut mice, whereas WBC counts in IDH1wtΔ7 mice remained normal. IDH1mutΔ7 mice died from monocytic leukemia that was phenotypically and morphologically indistinguishable from IDH1mut mice. HoxA9 IDH1mutΔ7 cells were readily transplantable into secondary recipients. During in vivo cell cycle analysis, we observed that the proportion of cells in S/G2/M phases was significantly higher in bone marrow cells transduced with IDH1mut or IDH1mutΔ7 when compared to cells transduced with IDH1wt or CTL. These data suggest that mutant IDH1 enhances myeloproliferation even in the absence of R-2HG. To identify R-2HG independent signaling pathways mediated by the mutant IDH1 protein, we first analyzed the gene expression of important regulators of cell cycle, differentiation, cell signaling and transcription by quantitative RT-PCR. Several genes (Ccnd1, Slc2a, Hdac3, Tgif2,and c-myc) were upregulated in IDH1mut and IDH1mutΔ7 cells compared to IDH1wt cells. Interestingly, we found a specific up-regulation of Ctnnb1 and Nfkb genes in IDH1mutΔ7 cells over both IDH1mut and IDH1wt cells. We next validated our mRNA expression results by immunoblotting and found that NFKB and ERK signaling were upregulated in both IDH1mut and IDH1mutΔ7 compared to IDH1wt and IDH1wtΔ7 cells. Interestingly, the protein level of β-catenin, STAT3 and STAT5 were many fold higher in IDH1mutΔ7 compared to IDH1mut and IDH1wt cells. β-catenin is known to be transactivated via c-Src, which is phosphorylated by EGFR to promote β-catenin nuclear localization and signaling. We traced this pathway for its relevance in our cells and found that IDH1mutΔ7 cells indeed showed higher levels of both EGFR and c-Src phosphorylation compared to IDH1mut cells. We performed immunofluorescence and cellular fractionation for β-catenin and found it to be partially localized in the nucleus in IDH1mutΔ7 but not in IDH1mut cells. We also observed an up-regulated STAT3 phosphorylation in IDH1mutΔ7 cells over IDH1mut. Conclusions: In summary, mutant IDH1 activates ERK and NFKB signaling, which is attributed to both R-2HG dependent and independent mechanisms of leukemogenesis. Interestingly, IDH1mutΔ7 employs R-2HG independent EGFR/β-catenin and JAK/STAT signaling for oncogenesis. This R-2HG-independent leukemogenesis reveals a novel signaling dynamic of IDH1mut which should be evaluated for its therapeutic potential. Disclosures Ganser: Novartis: Membership on an entity's Board of Directors or advisory committees. Heuser:Astellas: Research Funding; Karyopharm: Research Funding; Novartis: Consultancy, Honoraria, Research Funding; Pfizer: Consultancy, Honoraria, Research Funding; Janssen: Consultancy; StemLine Therapeutics: Consultancy; Bayer Pharma AG: Consultancy, Research Funding; Sunesis: Research Funding; BergenBio: Research Funding; Tetralogic: Research Funding; Daiichi Sankyo: Research Funding.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2018
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 2
    In: Blood, American Society of Hematology, Vol. 124, No. 21 ( 2014-12-06), p. 3171-3171
    Abstract: Background: Recurrent somatic mutations in splicing machinery components, including SRSF2, SF3B1 and U2AF1 have been recently identified and frequently found in a large spectrum of myeloid malignancies. The highest reported frequencies were 85% for myelodysplastic syndromes associated with ring sideroblasts (MDS-RS), 44% for MDS without RS and 55% for chronic myelomonocytic leukemia (CMML), 26% in therapy-related MDS or acute myeloid leukemia, 7% in primary acute myeloid leukemia (AML), and 9% in myeloproliferative neoplasms (MPNs). In primary myelofibrosis (PMF), SRSF2 mutations are associated with advanced age, higher DIPSS scores and poor prognosis. On the other hand, SF3B1 and U2AF1 mutations had no prognostic relevance for either overall or leukemia-free survival. Aim: The aim of this study was to investigate the clinical characteristics and prognosis of splicing factor mutations in SRSF2, SF3B1 and U2AF1 in patients with primary myelofibrosis or post ET/PV myelofibrosis that underwent allogeneic stem cell transplantation (alloHSCT). Methods: 158 patients with WHO defined primary myelofibrosis (n=112, 71%) or post-ET/post-PV myelofibrosis (n=46, 29%) underwent allogeneic hematopoietic stem cell transplantation. Bone marrow or peripheral blood samples were obtained before transplantation. U2AF1, SRSF2, and SF3B1 were amplified by PCR and were sequenced and confirmed with Sanger sequencing besides JAK2, CALR and MPL. Results: Mutations in SRSF2, SF3B1 and U2AF1were detected in 12 (8%), 6 (4%) and 9 (6%) patients, respectively. In total, 27 patients (18%) carried at least one of the three main spliceosome pathway mutations. Baseline characteristics were similarly distributed between SRSF2, SF3B1 and U2AF1 mutated and wildtype patients, respectively (primary vs secondary myelofibrosis, donor match, cytogenetic characteristics), except for higher median age of U2AF1 mutated compared to wildtype patients (67 vs. 58 years, P=0.023). We observed no associations between the presence of SRSF2/SF3B1/U2AF1 mutations and the presence of JAK2, CALR and MPL mutation status, respectively. The majority of patients (n=110, 70%) were transplanted with a matched donor (33 from related and 77 from unrelated donors). 48 patients (30%) received a mismatched unrelated allograft. Univariate analysis disclosed no differences in cumulative incidence of relapse (CIR) between mutated and wildtype patients (CIR: SRSF2, P=0.78; SF3B1, P=0.19; U2AF1, P=0.51; all three splicing genes, P=0.83), although none of the six SF3B1 mutated patients relapsed after alloHSCT. Non-relapse mortality (NRM) was not significantly different between mutated and wildtype patients (NRM: SRSF2, P=0.58; SF3B1, P=0.6; U2AF1, P=0.11; all three splicing genes, P=0.77). Furthermore, patients with mutations in one or any of the three main spliceosome genes had a similar 2-year overall survival after allogeneic HSCT as wildtype patients (2y-OS mutant vs wildtype: SRSF2, 75% vs 68%, P=0.8; SF3B1, 83% vs 69%, P=0.4; U2AF1, 38% vs 70%, P=0.13; all three splicing genes, 68% vs 70%, P=0.78). Conclusions: In our study we show that (i) splicing gene mutations are recurrent molecular aberrations in myelofibrosis patients, (ii) are not associated with an adverse prognostic outcome in patients with myelofibrosis undergoing allogeneic transplantation, and (iii) the potential negative prognostic effect of SRSF2 mutations can be overcome by alloHSCT. Splicing gene mutations do not constitute a risk factor for alloHSCT in patients with myelofibrosis. Disclosures No relevant conflicts of interest to declare.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2014
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 3
    In: Blood, American Society of Hematology, Vol. 124, No. 21 ( 2014-12-06), p. 366-366
    Abstract: Mutations in the metabolic enzymes IDH1 and IDH2 are frequently found in several tumors including glioma and acute myeloid leukemia (AML). Mutant IDH produces R-2-hydroxyglutarate (R2HG), which induces histone- and DNA-hypermethylation through inhibition of epigenetic regulators, thus linking metabolism to tumorigenesis. However, it is unknown whether R2HG alone is sufficient to recapitulate the biologic effects of mutant IDH1 in vivo. Recently, we have shown that IDH1mut cooperates with HoxA9 and induces a monocytic leukemia in mice. In order to evaluate the effects of R2HG independently of the mutated IDH1 protein and to determine whether the effects are specific to the R-enantiomer of 2HG, we treated mice transplanted with HoxA9 immortalised bone marrow cells with R2HG, S-2-hydroxyglutarate (S2HG), alpha-ketoglutarate (aKG) and phosphate buffered saline (PBS). The mice in the metabolite cohorts received an intraperitoneal dose of 1 mg per day. Mice treated with R2HG had higher engraftment levels at 16 and 20 weeks post transplantation than the mice treated with S2HG, αKG and PBS respectively (P 〈 .01). High WBC counts (70±16 /nl) and lower platelet counts than in control mice were observed in the cohort receiving R2HG after 16 to 20 weeks of treatment, while the S2HG, αKG and PBS cohorts had normal blood counts even at 20 weeks (P 〈 .05). Peripheral blood from R2HG treated mice revealed significantly more immature Mac1+Gr1- and less mature Mac1+Gr1+ cells at 12 and 16 weeks after treatment than S2HG, αKG and PBS treated mice (P 〈 .001). In addition, the R2HG treated mice died with a median latency of 137 days post transplantation from monocytic leukemia, while mice treated with S2HG, αKG and PBS died with a median latency of 223, 202 and 184 days respectively (P 〈 .001). Further, in order to assess whether R2HG alone was sufficient as a single hit to induce myeloproliferation, normal C57BL/6 mice (without HoxA9) were treated with R2HG, S2HG and PBS for eight months. No differences were observed for survival, blood counts, immunophenotype and frequencies of progenitor cells (lin-ckit+sca1+, CMP, GMP and MEP) between treatment groups and control. This data shows that the metabolite R2HG like the IDH1 mutant protein cooperates with HoxA9 to induce monocytic leukemia. We next compared mice receiving transplants of HoxA9+IDH1mut cells with mice receiving HoxA9 cells that were then treated with R2HG. Both cohorts developed monocytic leukemia, albeit with different kinetics. The Hoxa9+IDH1mut mice died with a median latency of 83 days while the R2HG cohort died with a median latency of 137 days post transplantation (P 〈 0.001). Also, while the former cohort developed severe leukocytosis, anemia and thrombocytopenia at 12 weeks, the R2HG treated mice had high WBC counts and lower platelet counts than control mice at 16 to 20 weeks after treatment. The faster disease kinetics in IDH1mut mice could be attributed to a significantly lower proportion of cells in G0/G1 and higher proportion of cells in S phase when compared to cells from mice treated with R2HG at 9 weeks after transplantation (P 〈 .001). This resulted from a marked downregulation of cyclin-dependent kinase inhibitors (Cdkn) 1A (p21), 1B (p27), 2A (p16), and 2B (p15) in HoxA9+IDH1mut cells as compared to HoxA9 cells treated with R2HG or PBS. In order to rule out an influence of differential R2HG levels on the disease progression between the two cohorts, levels of R2HG were quantified. IDH1mut expressing and R2HG treated bone marrow cells from mice had similarly high ratios of R2HG/S2HG. The quantified R2HG levels were also comparable to that of primary AML patient cells harbouring mutated IDH1. In unsupervised hierarchical clustering mutated cells clustered together with R2HG and separated from PBS treated mice. 69 of the top 100 enriched Gene Ontology gene sets of HoxA9+IDH1mut were also found in HoxA9+R2HG, suggesting largely redundant but also non-overlapping functions of the mutant IDH1 protein and the oncometabolite R2HG. In summary, we show that R2HG, similar to the mutant IDH1 protein, promotes leukemogenesis in cooperation with HoxA9, although with delayed kinetics. Our data proves that R2HG acts as an oncometabolite in vivo in a murine model of leukemogenesis. Disclosures No relevant conflicts of interest to declare.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2014
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 4
    In: Biology of Blood and Marrow Transplantation, Elsevier BV, Vol. 23, No. 7 ( 2017-07), p. 1095-1101
    Type of Medium: Online Resource
    ISSN: 1083-8791
    Language: English
    Publisher: Elsevier BV
    Publication Date: 2017
    detail.hit.zdb_id: 3056525-X
    detail.hit.zdb_id: 2057605-5
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 5
    In: Blood, American Society of Hematology, Vol. 132, No. Supplement 1 ( 2018-11-29), p. 759-759
    Abstract: Background: About 40% of IDH1 mutated (IDH1mut) acute myeloid leukemia (AML) patients respond to IDH1 inhibitors with a median duration of response of 8.2 months. A better understanding of the biology of IDH1mut leukemia may further improve the treatment of these patients. IDH1mut produces R-2-hydroxyglutarate (R-2HG), which activates PHD1 and PHD2 but have negligible effects on PHD3. In the present study we assessed whether PHD3 plays a role in the pathogenesis of IDH1 mutated leukemia and can be targeted in a patient-derived xenograft (PDX) model of IDH1 mutated AML. Methods: Bone marrow cells from Phdwt and Phd3ko mice were immortalized with HoxA9, and IDH1wildtype (IDH1wt) and IDH1mut respectively, were constitutively expressed. The effects on cell proliferation, apoptosis and colony formation were evaluated in vitro, whereas the leukemic potential was evaluated in vivo by transplantation in syngeneic mice. To show that PHD3 is a therapeutic target, either IDH1mut cells from AML patients were transduced with shRNA against PHD3 and transplanted in immunocompromised mice, or leukemic cells from an AML patient with mutated IDH1 were xenografted in immunocompromised mice and treated with the PHD inhibitor molidustat. Results: In in-vitro functional assays loss of Phd3 specifically impaired proliferation, apoptosis and clonogenic capacity of HoxA9-IDH1mut but not HoxA9-IDH1wt cells. Likewise, in mouse transplantation assays, loss of Phd3 eliminated HoxA9-IDH1mut induced leukemia. However, Phd3 was dispensable to the engraftment and proliferation of HoxA9-IDH1wt cells. Additionally, the IDH1-independent model of MN1-induced leukemia remained unaltered in the absence of Phd3, indicating the specificity of the role of Phd3 in mutant IDH1-induced transformation. To identify molecular pathways that might explain in vitro and in vivo phenotypes gene expression profiling was performed. Immune and stress-response pathways as well as metabolism-related genes were most prominently dysregulated in Phd3ko IDH1-mutant cells. Analysis of dysregulated transcription factors by gene set enrichment analysis revealed a depletion of key oncogenic transcription factors (Myc, Rb, Stk33, and Rps14) in Phd3ko IDH1mut cells compared to Phd3ko IDH1wt cells. To study if IDH1mut signals to Phd3 through R-2HG, we transduced Phd3kocells, with a splice variant of mutant IDH1, which does not produce R-2HG but causes leukemia in mice with similar kinetics as in mice with the full-length IDH1 mutant protein. Interestingly, loss of Phd3 also eliminated leukemia in these mice, which demonstrates that mutant IDH1 signals through Phd3 independently of R-2HG. To study the functional relevance of PHD3 inhibition in patients, cells from an IDH1 mutated AML patient were transduced with an shRNA against PHD3 and were transplanted in immunodeficient NSG mice. Inhibition of PHD3 depleted human AML cells in the IDH1-mutated PDX model. Moreover, the PHD inhibitor molidustat was 50-fold more active in IDH1mut (80 nM) compared to IDH1wt AML patient cells (4000 nM) in colony-forming unit assays. In a xenograft model of IDH1 mutated AML, molidustat significantly prolonged survival compared to control-treated mice (P 〈 .001). Conclusion: We demonstrate that the leukemogenic activity of the mutant IDH1 protein depends on PHD3 independently of R-2HG. We identified inhibition of PHD3 as a novel therapeutic strategy in IDH1 mutated AML. Since PHD3 can be targeted pharmacologically, combinatorial treatment of PHD3 and IDH1 inhibitors is warranted to improve eradication of leukemic stem cells in IDH1 mutated AML. #AC and MMAC share first authorship Disclosures Ganser: Novartis: Membership on an entity's Board of Directors or advisory committees. Heuser:Karyopharm: Research Funding; Daiichi Sankyo: Research Funding; Sunesis: Research Funding; Tetralogic: Research Funding; Bayer Pharma AG: Consultancy, Research Funding; StemLine Therapeutics: Consultancy; Janssen: Consultancy; Pfizer: Consultancy, Honoraria, Research Funding; BergenBio: Research Funding; Astellas: Research Funding; Novartis: Consultancy, Honoraria, Research Funding.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2018
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 6
    In: Blood, American Society of Hematology, Vol. 122, No. 16 ( 2013-10-17), p. 2877-2887
    Abstract: IDH1 promotes leukemogenesis in vivo in cooperation with HoxA9. Pharmacologic inhibition of mutant IDH1 efficiently inhibits AML cells of IDH1-mutated patients but not of normal CD34+ bone marrow cells.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2013
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 7
    In: Leukemia, Springer Science and Business Media LLC, Vol. 34, No. 2 ( 2020-02), p. 416-426
    Type of Medium: Online Resource
    ISSN: 0887-6924 , 1476-5551
    RVK:
    Language: English
    Publisher: Springer Science and Business Media LLC
    Publication Date: 2020
    detail.hit.zdb_id: 2008023-2
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 8
    In: Blood, American Society of Hematology, Vol. 124, No. 21 ( 2014-12-06), p. 3598-3598
    Abstract: Mutations in the metabolic enzymes isocitrate dehydrogenase 1 (IDH1) and 2 (IDH2) are frequently found in patients with glioma, acute myeloid leukemia (AML), melanoma, thyroid cancer, cholangiocellular carcinoma and chondrosarcoma. Mutant IDH produces R-2-hydroxyglutarate (R2HG), which induces histone- and DNA-hypermethylation through inhibition of epigenetic regulators, thus linking metabolism to tumorigenesis. We recently established an in vivo mouse model and investigated the function of mutant IDH1. By computational drug screening, we identified an inhibitor of mutant IDH1 (HMS-101), which inhibits mutant IDH1 cell proliferation, decreases R2HG levels in vitro, and efficiently blocks colony formation of AML cells from IDH1 mutated patients but not of normal CD34+ bone marrow cells. In the present study we investigated the effect of the inhibitor in our IDH1/HoxA9-induced mouse model of leukemia in vivo. To identify the maximally tolerated dose of HMS-101, we treated normal C57BL/6 mice with variable doses of HMS-101 for 9 days and measured the serum concentration. Mice receiving 0.5 mg and 1mg intraperitoneally once a day tolerated the drug well with mean plasma concentrations of 0.1 to 0.3 µM. To evaluate the effect of HMS-101 in the IDH1 mouse model, we transduced IDH1 R132C in HoxA9-immortalized murine bone marrow cells. Sorted transgene positive cells were then transplanted into lethally irradiated mice. After 5 days of transplantation, mice were treated with HMS-101 intraperitoneally for 5 days/week. The R/S-2HG ratio in serum was reduced 3-fold in HMS-101 treated mice after 8 weeks of treatment compared to control treated mice. HMS-101 or PBS treated mice had similar levels of transduced leukemic cells in peripheral blood at 2 and 6 weeks after transplantation. However, from week 6 to week 15 leukemic cells in peripheral blood decreased from 76% to 58, 63% to 29%, 67% to 7%, and 74% to 38% in 4/6 mice treated with HMS-101. In one mouse the percentage of leukemic cells was constant, and in one mouse it increased from week 6 to week 15 after transplantation. Leukemic cells increased constantly in peripheral blood until death in control treated mice. While the control cohort developed severe leukocytosis, anemia and thrombocytopenia around 8 to 10 weeks post transplantation, mice treated with HMS-101 still had normal WBC, RBC and platelet counts at 15 weeks after transplantation. Moreover, the HMS-101 treated mice had significantly more differentiated Gr1+CD11b+ cells in peripheral blood than control mice at 6 weeks and 15 weeks after transplantation and at death (P=.01). Morphologic evaluation of blood cells at 15 weeks or death from HMS-101 treated mice revealed a high proportion of mature neutrophils that were GFP positive and thus derived from IDH1 transduced cells, whereas control treated mice had monocytic morphology with a high proportion of immature cells. Importantly, HMS-101 treated mice survived significantly longer with a median latency of 87 days (range 80-118), whereas PBS-treated mice died with a median latency of 66 days (range 64-69) after transplantation (P 〈 .001). Of note, HMS-101 was found to be specific for mutant IDH1, as mutant IDH2 cells were not preferentially inhibited over IDH2 wildtype cells in vitro. This data demonstrates that HMS-101 specifically inhibits R2HG-production of mutant IDH1 in vivo, inhibits proliferation, induces differentiation in leukemic cells, and thus prolongs survival of IDH1mutant leukemic mice. Therefore, HMS-101 - a novel inhibitor of mutant IDH1 - shows promising activity in vivo and warrants further development towards clinical use in IDH1 mutated patients. Disclosures Chaturvedi: Hannover Medical School: Patents & Royalties. Preller:Hannover Medical School: Patents & Royalties. Heuser:Hannover Medical School: Patents & Royalties.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2014
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 9
    In: Blood, American Society of Hematology, Vol. 124, No. 21 ( 2014-12-06), p. 4643-4643
    Abstract: Background and Aim: Deletion of 5q is the most frequent cytogenetic aberration in MDS and is associated with distinct clinical characteristics, disease course and sensitivity to lenalidomide. The serine-threonine kinase CSNK1A1 is located in the commonly deleted region at 5q32 and has been described as a tumor-suppressor gene in colon cancer and acute myeloid leukemia through regulation of ß-catenin and p53. Recently, missense mutations in CSNK1A1 have been described in individual patients with del(5q) MDS. The aim of our study was to characterize the frequency and potential prognostic impact of CSNK1A1mutations in MDS and AML following MDS. Methods: 192 patients with MDS or AML following MDS (sAML) and deletion of chromosome 5q and 406 patients with MDS/sAML without deletion of chromosome 5q were included in the current analysis (n=598 in total). Patients with MDS (n=442) or sAML (n=156) were cytogenetically characterized by chromosome banding analysis and molecularly analyzed for mutations in exon 3 and 4 of CSNK1A1, the region critical for the kinase function, by Sanger sequencing. Patients with mutated CSNK1A1 were also analyzed for mutations in TP53 by next-generation or Sanger sequencing. Results: CSNK1A1 mutations were found in 17 (8.9%) of 192 MDS patients with del(5q). The mutation frequency was similar between patients with isolated del(5q) (n=153) and patients with concurrent cytogenetic aberrations or missing additional cytogenetic information (n=39)(9.2% vs 7.7%, P=.7). No mutation of CSNK1A1 was found in any of 406 MDS/sAML patients without del(5q). Thirteen patients (76%) had missense mutations affecting amino acid E98 in exon 3 of CSNK1A1. Of these, the glutamic acid to lysin substitution was the most frequent amino acid substitution (n=7). All mutations of glutamic acid 98 had a high probability to be damaging to the protein based on PolyPhen2 predictions (scores 0.922 to1). One patient had an Asn86Tyr mutation concurrently with the Glu98Ala mutation. Four patients (24%) had missense mutations affecting aspartic acid 140 in exon 4 of CSNK1A1. These mutations had moderate PolyPhen2 prediction scores (0.558-0.798). Three of the 17 CSNK1A1 mutated patients had additional cytogenetic aberrations besides del(5q), i.e. one trisomy 8, one trisomy 11, and one monosomy 7. None of the CSNK1A1 patients had a concurrent TP53 mutation. Del(5q) patients with wildtype or mutated CSNK1A1 had a similar median age (73.3 vs 77.5 years, P=.15). 70% and 59% of wildtype and mutated CSNK1A1 patients had female sex, respectively (P=.33). The WBC count was similar between wildtype and mutated CSNK1A1patients (3.9 vs 4.6, P=.47). Survival information was available for 155 patients with del(5q) (81%) including 16 patients (94%) with mutated CSNK1A1. Median follow-up from the time of sample harvest was 2.02 years. The probability of survival at 2 years was 41% for CSNK1A1 mutated and 72% for CSNK1A1wildtype patients (P=.059, log-rank test), suggesting a potential negative prognostic impact of CSNK1A1 mutations in del(5q) MDS patients. Conclusion: CSNK1A1 mutations are highly specific for MDS patients with del(5q) and are one of the most frequent recurrent genetic aberrations in these patients. Our survival analysis suggests that CSNK1A1 mutations have an unfavorable prognostic effect in patients with MDS and del(5q); however, the prognostic impact has to be confirmed in additional patients. Mutation analysis of exon 3 and 4 of CSNK1A1 should be included in the routine workup of MDS patients with deletion of 5q. Disclosures Meggendorfer: MLL Munich Leukemia Laboratory: Employment. Haferlach:MLL Munich Leukemia Laboratory: Equity Ownership. Kobbe:Celgene: Honoraria, Research Funding; Amgen: Honoraria, Research Funding; Medac: Other; Astellas: Honoraria, Research Funding; Novartis: Honoraria, Research Funding; Neovii: Other. Haferlach:MLL: Equity Ownership.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2014
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 10
    In: Blood, American Society of Hematology, Vol. 126, No. 23 ( 2015-12-03), p. 352-352
    Abstract: Introduction Primary or post-ET/post-PV myelofibrosis is one of the Philadelphia chromosome-negative chronic myeloproliferative neoplasia characterized by significantly reduced overall survival. More recently several mutated genes have been detected, which allow, in addition to clinical factors, to identify patients with a significantly shorter overall survival and a higher risk of transformation into acute leukemia. Allogeneic stem cell transplantation is still the only curative treatment option for patients with myelofibrosis, and due to the inherent risk of the treatment procedure careful selection of the patients is required. Molecular genetics may help to select patients for allogeneic stem cell transplantation. Patients and methods To determine the impact of mutated genes in patients with myelofibrosis who underwent allogeneic stem cell transplantation we analyzed samples from 169 patients with a median age of 58 years (r: 18 - 75) who received allogeneic stem cell transplantation either from related (n = 36) or unrelated (n = 133) donor. Stem cell source were more often peripheral blood stem cells (n = 165) than bone marrow (n = 4). The intensity of conditioning was mainly reduced intensity (n = 166), rather than myeloablative conditioning (n = 3). Patients suffered from primary myelofibrosis (n = 110), post-ET/PV myelofibrosis (n = 46), while 13 patients were in acceleration or had transformed into acute myeloid leukemia. According to dynamic IPSS (DIPSS) (n = 165) the patients were either low (n = 7), intermediate-1 (n = 35), intermediat-2 (n = 91), or high risk (n = 32). Regarding molecular genetics we found JAK2V617F mutations in 62%, calreticulin (CALR) mutations in 20%, MPL mutations in 4%, U2AF1 in 7%, SRSF2 in 10%, SF3B1 in 4%, ASXL1 in 29%, IDH1 in 2%, IDH2 in 3%, CBL in 1%, DNMT3A in 4%, TET2 in 10%, EZH2 in 4%, while none of the patients showed mutations in ETV6 and PTPN11. Overall, only in 11 patients no mutation could be detected. One mutation could be detected in 41%, 2 mutations in 30%, 3 mutations in 11%, 4 mutations in 5%. Results During follow-up 39 patients experienced relapse and 46 patients experienced non-relapse mortality. From the non-molecular factors regarding disease-free survival in univariate analysis age 〈 58 (p 〈 0.01), intermediate-1 and low risk according to DIPPS (p = 0.002), HLA-matched vs. mismatched (p = 0.04) were significant factors for improved disease-free survival. Regarding molecular markers improved disease-free survival was seen for patients with mutations in CALR (p = 0.005), while negative impact on disease-free survival was seen for mutations in U2AF1 (p = 0.035), ASXL1 (p = 0.05), IDH2 (p = 0.006), DNMT3A (p = 0.029). No significant difference could be seen for patients with EZH2, IDH1, SRSF2, and SF3B1 mutations. There was no difference in disease-free survival for patients without any mutation vs. 1, and more than 1 mutation (p = 0.12). Regarding the previously described unfavorable mutations ASXL1, SRSF2, EZH2, IDH1, and IDH2, we found 40 patients who had at least 1 of these unfavorable mutations, 11 had 2 of these mutation, and 1 had 3 of these unfavorable mutations. However, the estimated 5-year disease-free survival did not differ significantly between patients without any of these unfavorable mutations, with 1 or with 2 of them (47 vs. 40 vs. 41%, p = 0.5). Conclusions These results suggest that some molecular marker such as ASXL1, U2AF1, IDH2 and DNMT3A negatively influence DFS in myelofibrosis after allogeneic stem cell transplantation in a univariate analysis. In contrast, the poor prognosis of the recently described unfavorable mutated genes SRSF2, EZH2, and IDH1 was not observed and may therefore be overcome by allogeneic stem cell transplantation. Disclosures No relevant conflicts of interest to declare.
    Type of Medium: Online Resource
    ISSN: 0006-4971 , 1528-0020
    RVK:
    RVK:
    Language: English
    Publisher: American Society of Hematology
    Publication Date: 2015
    detail.hit.zdb_id: 1468538-3
    detail.hit.zdb_id: 80069-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
Close ⊗
This website uses cookies and the analysis tool Matomo. More information can be found here...