scholarly journals CREB3L1 Overexpression Can Reliably Discriminate Ph-MPNs from Reactive Cases

Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 41-41
Author(s):  
Soji Morishita ◽  
Hajime Yasuda ◽  
Saya Yamawaki ◽  
Hideya Kawaji ◽  
Masayoshi Itoh ◽  
...  

Discrimination of Philadelphia-negative myeloproliferative neoplasms (Ph-MPNs) from reactive hypercytosis and myelofibrosis is imperative because treatment strategies differ greatly, and an exhaustive search for the underlying cause becomes mandatory in reactive cases. However, discrimination is not necessarily easy in the real-world setting, and a simple and universally utilizable method that can efficiently differentiate Ph-MPNs from reactive cases is awaited. We extracted platelet rich plasma (PRP) derived RNAs from 9 essential thrombocythemia (ET) patients (3 patients with JAK2V617F, 3 patients with MPLW515L/K, and 3 patients with CALR exon 9 frameshift mutation) and 6 patients with reactive thrombocytosis (3 cases due to chronic inflammation, and 3 cases due to rebound thrombocytosis) and performed RNA-seq to identify a gene expressed specifically in ET. RNA-seq analysis followed by differential expression and principal component analysis revealed that CREB3L1 had the highest impact in discriminating ET from reactive cases. Subsequently, expression levels of CREB3L1 in PRP were quantified by reverse transcription quantitative PCR and compared within patients with various Ph-MPNs harboring either JAK2, MPL, or CALR mutations, chronic myeloid leukemia (CML), and reactive cases, and found that CREB3L1 expression levels were significantly higher in 66 ET compared to 33 reactive thrombocytosis (p < 0.0001), 26 polycythemia vera (PV) compared to 23 reactive erythrocytosis (p < 0.0001), 22 primary myelofibrosis and 15 post-PV/ET myelofibrosis (MF) compared to 3 reactive MF (p < 0.001, and p < 0.01, respectively), and the entire cohort of 129 Ph-MPN compared to 5 CML patients (p < 0.001). A clear cut-off value discriminating Ph-MPNs and non-Ph-MPNs was determined, and sensitivity and specificity were both 1.0000. Furthermore, when we tested CREB3L1 expression levels of triple-negative cases with thrombocytosis, all patients with CREB3L1 overexpression were pathologically diagnosed as ET by bone marrow biopsy. We demonstrate that CREB3L1 overexpression can single-handedly and reliably discriminate Ph-MPNs from reactive hypercytosis, reactive myelofibrosis, and CML. Early utilization of this method in the diagnostic process can guide patients to an efficient diagnosis and free many patients from unnecessary testing. Disclosures Komatsu: Otsuka Pharmaceutical Co., Ltd., PharmaEssentia Japan KK, AbbVie GK, Celgene KK, Novartis Pharma KK, Shire Japan KK, Japan Tobacco Inc: Consultancy; Takeda Pharmaceutical Co., Ltd, Novartis Pharma KK, Shire Japan KK: Speakers Bureau; AbbVie: Other: member of safety assessment committee in M13-834 clinical trial.; PPMX: Consultancy, Research Funding; Otsuka Pharmaceutical Co., Ltd., Shire Japan KK, Novartis Pharma KK, PharmaEssentia Japan KK, Fuso Pharmaceutical Industries, Ltd., Fujifilm Wako Pure Chemical Corporation, Chugai Pharmaceutical Co., Ltd., Kyowa Hakko Kirin Co., Ltd., Takeda Pharmaceutica: Research Funding; Meiji Seika Pharma Co., Ltd.: Patents & Royalties: PCT/JP2020/008434, Research Funding.

Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 1179-1179
Author(s):  
Hideaki Mizuno ◽  
Akira Honda ◽  
Mineo Kurokawa

Abstract Resistance to anthracycline and cytarabine based conventional chemotherapy often occurs and results in extremely poor prognosis in patients with acute myeloid leukemia (AML). Although chemotherapy resistance is the most critical clinical problem, the mechanisms by which AML confers resistance to conventional chemotherapy are not yet fully understood. In this study, we investigated the key mechanisms of chemotherapy resistance through single cell RNA-sequencing analysis using paired bone marrow AML cells longitudinally collected from two AML-MRC patients at diagnosis and relapse after anthracycline-based chemotherapy. AML blasts were sorted by CD45/SSC gating and subjected to single cell RNA-seq analysis. Single cell RNA-seq was performed using 10x Genomics' Chromium System. Mean estimated number of cells per sample was 3.403 (2,731-4,200) and median detected genes per cell ranged 3,030 to 3,918 among four samples. Data collected from paired samples were combined in following analysis. Transcriptome based clustering following UMAP dimensionality reduction distinguished 5 and 9 cluster groups in each paired sample. Chemotherapy sensitive cluster groups dominant at diagnosis and chemotherapy resistant cluster groups dominant at relapse were clearly divided. In each paired sample, a few AML cells at diagnosis were allocated to chemotherapy resistant cluster groups. This suggested that transcriptionally identifiable less frequent cells resistant to chemotherapy existed at diagnosis and may expand during and/or after chemotherapy maintaining its transcriptional features. Next, to determine whether these transcriptional features are correlated with DNA mutation profiles, we labeled DNA mutation status to each cell and compared frequencies of mutation. As far as we detected, AML recurrent mutations such as DNMT3A R882C and TP53 missense mutation were not related to chemotherapy resistant cluster groups, although this method was relatively limited by the nature of RNA-seq-based mutation detection. Then we sought to determine transcriptional features of resistant clones. Gene set enrichment analysis identified some gene groups such as E2F signaling pathway, MYC signaling pathway, hedgehog signaling pathway and TNFA signaling pathway as transcriptional signatures related to emergence after chemotherapy. Analysis of known hematopoietic differentiation gene signatures showed distinct differentiation profiles in each cluster groups, whereas resistant cluster groups were not necessarily related to hematopoietic stem cell signatures. Intrapatient variations of transcriptional signatures among the resistant cluster groups were detected, which indicated that accurate detection of transcriptional features related to chemotherapy resistance may be difficult by using bulk RNA-seq method. As for other cluster groups which were not dominant both at diagnosis and relapse, these cluster groups hardly changed its frequencies between at diagnosis and relapse, which suggested less proliferative leukemia cells persisted during chemotherapy and have various transcriptional features although whether these persisting cells contribute to relapse was unclear. Since enriched transcriptional signatures in resistant cluster groups were not consistent between the two patients, further analysis using samples collected from more patients would be needed to determine common critical chemotherapy resistant transcriptional signature. In conclusion, our analysis suggested that a transcriptionally identifiable small fraction of cells showing gene signatures related to chemotherapy resistance at diagnosis may expand during chemotherapy and revealed intrapatient transcriptional complexity of response to chemotherapy, which cannot be uncovered by bulk RNA-sequencing. Disclosures Honda: Takeda Pharmaceutical: Other: Lecture fee; Otsuka Pharmaceutical: Other: Lecture fee; Chugai Pharmaceutical: Other: Lecture fee; Ono Pharmaceutical: Other: Lecture fee; Jansen Pharmaceutical: Other: Lecture fee; Nippon Shinyaku: Other: Lecture fee. Kurokawa: MSD K.K.: Research Funding, Speakers Bureau; Kyowa Hakko Kirin Co., Ltd.: Research Funding, Speakers Bureau; Daiichi Sankyo Company.: Research Funding, Speakers Bureau; Astellas Pharma Inc.: Research Funding, Speakers Bureau; Pfizer Japan Inc.: Research Funding, Speakers Bureau; Nippon Shinyaku Co., Ltd.: Research Funding, Speakers Bureau; Sumitomo Dainippon Pharma Co., Ltd.: Research Funding, Speakers Bureau; Otsuka Pharmaceutical Co., Ltd.: Research Funding, Speakers Bureau; Eisai Co., Ltd.: Research Funding, Speakers Bureau; ONO PHARMACEUTICAL CO., LTD.: Research Funding, Speakers Bureau; Teijin Limited: Research Funding, Speakers Bureau; Takeda Pharmaceutical Company Limited.: Research Funding, Speakers Bureau; Chugai Pharmaceutical Company: Research Funding, Speakers Bureau; AbbVie GK: Research Funding, Speakers Bureau.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 4203-4203
Author(s):  
Nicole Kucine ◽  
Amanda R. Leonti ◽  
Aishwarya Krishnan ◽  
Rhonda E. Ries ◽  
Ross L. Levine ◽  
...  

Introduction : Myeloproliferative neoplasms (MPNs) are rare clonal bone marrow disorders in children characterized by high blood counts, predisposition to clotting events, and the potential to transform to myelofibrosis or acute myeloid leukemia (AML). Children with MPNs have lower rates of the known driver mutations (in JAK2, MPL, and CALR) than adult patients, and the underlying pathways and molecular derangements in young patients remain unknown. Given the lack of knowledge about pediatric MPNs, it is critical that we gain a better understanding of the dysregulated pathways in these diseases, which is necessary for improving disease understanding and broadening treatment options in children. Therefore, the objective of this work was to identify differentially expressed genes and pathways between children with MPNs and healthy controls, as well as children with AML, to guide further study. Methods : Mononuclear cells were extracted from peripheral blood of pediatric MPN patients (n=20) and pediatric and young adult AML patients (n=1410), and bone marrow of normal controls (NC, n=68). AML patient samples were being evaluated as part of a Children's Oncology Group planned analysis. To identify an expression profile unique to MPNs, transcriptome data from MPN patients was contrasted against NC and AML patients. All samples were ribodepleted and underwent Illumina RNA-Seq to generate transcriptome expression data. All analyses were performed in R. Differentially expressed genes were identified using the voom function from the limma package (v. 3.38.3), and enriched pathways were identified using the pathfindR package (v. 1.3.1). Unsupervised hierarchical clustering and heatmap generation was performed using the ComplexHeatmap package (v. 1.20.0). Results : MPN patient samples showed a unique expression signature, distinct from both AML patients and normal controls. Unsupervised PCA plot (Figure 1A) and heatmaps (Figure 1B) show that MPN samples cluster together. There were 4,012 differentially expressed (DE) genes in MPNs compared to NC and 6,743 DE genes in MPNs compared to AML patients. There were 2,493 shared genes between the 2 groups (Figure 1C.) Significantly DE genes between MPNs and other groups included multiple platelet-relevant genes including PF4 (CXCL4), PF4V1, P2RY12, and PPBP (CXCL7). Interestingly, PF4V1 was the most DE gene in MPNs compared to AML, and third highest versus NC. Dysregulation of some of these genes has been seen in adult MPNs, as well as thrombosis. Further comparison of transcriptome profiles between children with (n=13) and without (n=7)JAK2 mutations showed upregulation of three genes, CFB, C2, and SERPING1, which are all known complement genes, implicating complement activation in JAK2-mutated MPN patients. Complement activation has previously been reported in adult MPNs. Pathway enrichment analysis shows a number of immune and inflammatory pathways as enriched in MPN patients compared to both AML and NC. There were 179 enriched pathways in MPNs compared to AML and 142 compared to NC, with 134 common pathways (Figure 1D.) The systemic lupus erythematosus pathway was the most heavily enriched pathway in MPNs compared to both AML and NC. Additional pathways with significant enrichment include hematopoietic cell lineage, cytokine-cytokine interactions, DNA replication, and various infection-relevant pathways. The JAK-STAT signaling pathway was also enriched in MPNs compared to both AML and NC, as was the platelet activation pathway. Conclusion: Transcriptome evaluation of childhood MPNs shows enrichment of numerous inflammatory and immune pathways, highlighting that, as in adult MPNs, inflammation is implicated in pediatric MPNs. Furthermore, specific complement genes were upregulated in JAK2-mutant MPN. Upregulation of platelet-specific genes implies potential insights into disease mechanisms and warrants more study. Variations in the cell populations may account for some of the differences seen, however all samples were largely mononuclear cells, making their comparisons reasonable. Further analysis of this early data is needed to better assess inflammatory changes and platelet activation in pediatric MPNs, as are larger sample sizes. Individual cells may have differential expression of various genes, and future experiments with single-cell RNA-seq would be helpful to further elucidate differences. Disclosures Levine: Novartis: Consultancy; Loxo: Membership on an entity's Board of Directors or advisory committees; Celgene: Consultancy, Research Funding; Gilead: Consultancy; Roche: Consultancy, Research Funding; Lilly: Honoraria; Amgen: Honoraria; Qiagen: Membership on an entity's Board of Directors or advisory committees; Imago Biosciences: Membership on an entity's Board of Directors or advisory committees; C4 Therapeutics: Membership on an entity's Board of Directors or advisory committees; Prelude Therapeutics: Research Funding; Isoplexis: Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 1469-1469
Author(s):  
Zhu Shen ◽  
Wenfei Du ◽  
Cecelia Perkins ◽  
Lenn Fechter ◽  
Vanita Natu ◽  
...  

Abstract Predicting disease progression remains a particularly challenging endeavor in chronic degenerative disorders and cancer, thus limiting early detection, risk stratification, and preventive interventions. Here, profiling the spectrum of chronic myeloproliferative neoplasms (MPNs), as a model, we identify the blood platelet transcriptome as a proxy strategy for highly sensitive progression biomarkers that also enables prediction of advanced disease via machine learning algorithms. Using RNA sequencing (RNA-seq), we derive disease-relevant gene expression in purified platelets from 120 peripheral blood samples constituting two time-separated cohorts of patients diagnosed with one of three MPN subtypes at sample acquisition - essential thrombocythemia, ET (n=24), polycythemia vera, PV (n=33), and primary or post ET/PV secondary myelofibrosis, MF (n=42), and healthy donors (n=21). The MPN platelet transcriptome reveals an incremental molecular reprogramming that is independent of patient driver mutation status or therapy and discriminates each clinical phenotype. Differential markers in each of ET, PV and MF also highlight candidate genes as potential mediators of the pro-thrombotic and pro-fibrotic phenotypes in MPNs. In ET and PV, a strong thromboinflammatory profile is revealed by the upregulation of several interferon inducible transmembrane genes (IFITM2, IFITM3, IFITM10, IFIT3, IFI6, IFI27L1, IFI27L2), interleukin receptor accessory kinases/proteins (IRAK1, IL15, IL1RAP, IL17RC) and several solute carrier family genes (SLC16A1, SLC25A1, SLC26A8, SLC2A9) as glucose and other metabolic transport proteins, and coagulation factor V (F5). In MF, fibrosis-specific markers were identified by an additional focused comparison of MF patients versus ET and PV, showing increased expression of several pro-fibrotic growth factors (FGFR1, FGFR3, FGFRL1), matrix metalloproteinases (MMP8, MMP14), vascular endothelial growth factor A (VEGFA), insulin growth factor binding protein (IGFBP7), and cell cycle regulators (CCND1, CCNA2, CCNB2, CCNF). Also, focusing on the JAK-inhibitor ruxolitinib/RUX-specific signatures, we not only confirm previous observations on its anti-inflammatory and immunosuppressive effects (e.g. downregulation in our RUX-treated cohort of IL1RAP, CXCR5, CPNE3, ILF3) but also identify new gene clusters responsive to RUX - e.g. inhibition of type I interferon (e.g. IFIT1, IFIT2, IFI6), chromatin regulation (HIST2H3A/C, HIST1H2BK, H2AFY, SMARCA4, SMARCC2), epigenetic methylation in mitochondrial genes (ATP6, ATP8, ND1-6 and NDUFA5), and other proliferation, and proteostasis-associated markers as putative targets for MPN-directed therapy. Mechanistic insights from our data highlight impaired protein homeostasis as a prominent driver of MPN evolution, with a persistent integrated stress response. Preliminary ex vivo data on MPN patient bone-marrow-derived CD34+ cells and cultured megakaryocytes validate a proteostasis-focused subset of our peripheral platelet RNA-seq signatures. Further leveraging this substantive dataset, and in particular a progressive expression gradient across MPN, we develop a machine learning model (Lasso-penalized regression) predictive of the advanced subtype MF at high accuracy and under two conditions of validation: i) temporal Stanford internal (AUC-ROC of 0.96) and ii) geographic external cohorts (AUC-ROC of 0.97, using independently published data of an additional n=25 MF and n=46 healthy donors). Lasso-derived signatures offer a robust core set of < 5 MPN progression markers. Together, our platelet transcriptome snapshot of chronic MPNs demonstrates a methodological avenue for disease risk stratification and progression beyond genetic data alone, with potential utility in a wide range of age-related disorders. Part of the work contributing to this abstract has been posted as a preprint at this link: https://www.biorxiv.org/content/10.1101/2021.03.12.435190v2 Figure 1 Figure 1. Disclosures Gotlib: Blueprint Medicines: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Incyte: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Deciphera: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Novartis: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Abbvie: Membership on an entity's Board of Directors or advisory committees, Research Funding; BMS: Research Funding; Kartos: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; PharmaEssentia: Honoraria, Membership on an entity's Board of Directors or advisory committees; Cogent Biosciences: Honoraria, Membership on an entity's Board of Directors or advisory committees, Other: Chair for the Eligibility and Central Response Review Committee, Research Funding; Allakos: Consultancy.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 3072-3072
Author(s):  
Francisca Ferrer Marin ◽  
Ana Belen F Arroyo ◽  
Beatriz Bellosillo ◽  
Lurdes Zamora ◽  
Ana Kerguelen Fuentes ◽  
...  

Abstract Introduction: Transformation to secondary myelofibrosis (MF) occurs as part of the natural history of polycythemia vera (PV) and essential thrombocythemia (ET), the two more indolent Ph-negative myeloproliferative neoplasms (MPN). Once transformed, survival is remarkably shorted. Chronic inflammation plays a critical role in the progression of MPN, driving clonal expansion toward end stage disease. Importantly, MPN are characterized by the production of inflammatory cytokines, by both malignant and non-malignant clone. Inflammation and cancer share a common pathway, i.e. NF-κB. Interestingly, miR-146a regulates TLR/NF-κB pathway through the inhibition of its targets, IRAK1 and TRAF6, decreasing the production of cytokines. Based on: i) miR-146a-/- mice develop an MF-like phenotype with aging; and ii) miR-146a polymorphism (miRSNPs) rs2431697, influences its expression levels (50% decrease in TT individuals); we hypothesized that lower miR-146a-5p levels associated to this miRSNPs may result in high risk to develop MF. Objective: To evaluate the association of rs2431697 with MF transformation and to study the molecular mechanisms beyond this association. Methods: We genotyped rs2431697 in 938 patients (312 MF, 299 PV, and 327 ET) recruited from 13 tertiary Spanish institutions belonging to GEMFIN and 600 controls. The levels of miR-146a and IRAK1 were evaluated by qRT-PCR in total blood RNA of homozygous patients (TT=30, CC=25) with PV or ET and in healthy subjects (TT=7, CC=7). In miR-146a-/- mice, 2 and 9 months old, we evaluated spleen size and cellularity: degree of fibrosis in bone marrow (H&E and silver staining); and STAT3 and pSTAT3 in granulocytic lysates by western blot. Results: Association analysis, taken controls as reference, showed that TT genotype (associated in the literature with low levels of mir-146a) is associated to MF with an OR of 1.36 (1.01-1.82, p=0.04). Among MF patients, the subgroup with the greatest differences was the one of secondary MF (OR = 1.47, CI: 0.98-2.20) (Table 1 a,b). Next, we compared the genetic frequencies of rs2431697 SNPs between the secondary MF patients and the population in risk. Confirming our hypothesis, we observed an enrichment of TT genotype in the post-PV/TE MF group (n=132) compared to the PV+TE group (n=626) (OR=1.51; p<0.05), Table 1c. In patients with PV or true-ET (WHO criteria) and known clinical follow-up (n=243), excluding pre-fibrotic MF, 8.6% were transformed. The median time to transformation was 27 years, being significantly shorter in patients with PV (vs. ET); homozygous for JAK2V617F; and in TT carriers (vs. TC+CC) (Figure 1A). The mayor differences were seen at early time points (Wilconxon test, p=0.001). In fact, 7 out of 10 TT patients who progress to MF, did so in the first 10 years after diagnosis (70%) as compared with 2 out of 11 CC/CT patients (18%).Both groups (TT and TC+CC) were similar in age, sex, cell counts, initial diagnosis (PV/ET), driver mutations and fibrosis grade 1. In the multivariate analysis, TT genotype remained statistically significative [OR 2.87; CI: 1.19-6.94; p=0.019], independently of phenotype (PV/ET) or V617F allele burden. Both in JAK2 mutated and wild type patients, the time to progression to secondary MF was significantly shorter among TT patients (p=0.027) (Figure 1B). Moreover, TT genotype helped to categorize the risk of progression to MF independently of the driver mutation (JAK2 or CALR) or JAK2 burden allele (p=0.06) (Figure 1C). Consistently, TT patients showed a trends towards a lower expression of miR-146a (p=0.08) and higher IRAK1 (p=0.07) with a significant correlation between both (p<0.01). Finally, we evaluated the association between the JAK-STAT3 and TLR/NF-κB pathways in mice miR-146a-/-. We observed higher total STAT3 and pSTAT3 expression levels in miR-146a-/- than in WT mice (Figure 1D). This increase correlated with aging, and were associated with the appearance of splenomegaly, extramedular hematopoyesis and bone marrow fibrosis at 9 months of age (Figure 1E). Conclusion: rs2431697-TT is an independent marker of early progression to secondary MF. The lower expression of miR-146a that this SNP confers is associated with an increase in JAK-STAT3 signaling. Our findings include, for the first time, miR-146a in the MPN signaling pathways. Thus, miR-146a, modulating the activation of NF-kB-IRAK1, could indirectly regulates JAK-STAT3 signalling. CINC424AES05T Disclosures Ferrer Marin: Novartis: Consultancy, Research Funding; Incyte: Consultancy. Hernandez Boluda:Incyte: Consultancy; Novartis: Consultancy. García Gutiérrez:Incyte: Consultancy, Research Funding, Speakers Bureau; Novartis: Consultancy, Research Funding, Speakers Bureau; Bristol-Myers Squibb: Consultancy, Research Funding, Speakers Bureau. Gómez-Casares:Bristol-Myers Squibb: Speakers Bureau; Incyte: Speakers Bureau; Novartis: Speakers Bureau. Besses:Novartis: Honoraria, Research Funding; Shire: Honoraria.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 1042-1042
Author(s):  
Kohei Hosokawa ◽  
Sachiko Kajigaya ◽  
Keyvan Keyvanfar ◽  
Qiao Wangmin ◽  
Yanling Xie ◽  
...  

Abstract Background. Paroxysmal nocturnal hemoglobinuria (PNH) is a rare acquired blood disease, characterized by hemolytic anemia, bone marrow (BM) failure, and venous thrombosis. The etiology of PNH is a somatic mutation in the phosphatidylinositol glycan class A gene (PIG-A) on the X chromosome, which causes deficiency in glycosyl phosphatidylinositol-anchored proteins (GPI-APs). The involvement of T cells in PNH is strongly supported by clinical overlap between PNH and aplastic anemia (AA); the presence of GPI-AP deficient cells in AA associated with favorable response to immunosuppressive therapy; and an oligoclonal T cell repertoire in PNH patients. However, the molecular mechanisms responsible for the aberrant immune responses in PNH patients are not well understood. To identify aberrant molecular mechanisms involved in immune targeting of hematopoietic stem cells in BM, RNA sequencing (RNA-seq) was applied to examine the transcriptome of T cell subsets from PNH patients and healthy controls. Method. Blood samples were obtained after informed consent from 15 PNH patients and 15 age-matched healthy controls. For RNA extraction, freshly isolated peripheral blood mononuclear cells were sorted on the same day of blood draw to obtain four different T cell (CD3+ CD14- CD19- ViViD-) populations [CD4+ naïve (CD45RA+ CD45RO-), CD4+ memory (CD45RA- CD45RO+), CD8+ naïve (CD45RA+ CD45RO-), and CD8+ memory (CD45RA- CD45RO+) T cells] by fluorescence-activated cell sorter . RNA-Seq analysis from three PNH and three healthy controls was performed using the Illumina HiSeq™ 2000 platform. The Ingenuity® Pathway Analysis and Gene set enrichment analysis (GSEA) were employed to elucidate transcriptional pathways. RNA-seq data were validated by flow cytometry and quantitative real-time RT-PCR (RT-qPCR). Results and Discussion . Differentially expressed gene analysis of four T cell subsets showed distinct gene expression signatures in individual T cell subsets. In CD4+ naïve T cells, 11 gene expression levels were significantly different: five upregulated (including SRRM2 and TNFSF8) and six downregulated genes (including GIMAP6) (> 2 fold change, false discovery rate [FDR] < 0.05). In CD4+ memory T cells, 25 gene expression levels were significantly different: 15 upregulated (including JUND and TOB1) and 10 downregulated genes (including GIMAP4). In CD8+ naive T cells, only two gene expression levels were significantly different: upregulated CTSW and downregulated RPL9. In CD8+ memory T cells, seven gene expression levels were significantly different: two upregulated (CTSW and DPP4) and five downregulated genes (including SLC12A7). Further, differentially expressed gene analysis was performed by combining CD4+ naïve, CD4+ memory, CD8+ naïve, and CD8+ memory T cells from PNH or healthy controls, respectively. Out of 55 gene expression levels that were significantly different, 41 were upregulated (including TNFAIP3, JUN, JUND, TOB1, TNFSF8, and CD69) and 14 downregulated (including GIMAP4). By canonical pathways analysis, putative gene network interactions of differentially expressed genes were significantly enriched for canonical pathways of TNFR1, TNFR2, IL-17A, and CD27 signaling. By GSEA, the most significantly upregulated gene sets in CD4+ naïve, CD4+ memory, CD8+ naïve, and CD8+ memory T cells from PNH patients displayed gene signatures related to the "IGF1 pathway", "Pre-NOTCH expression and processing", "AP-1 pathway", and "ATF2 pathway", respectively. For validation of the RNA-seq data, we chose seven genes (TNFAIP3, JUN, JUND, TOB1, TNFSF8, CD69, and CTSW) because these are important mediators involved in regulation for T cells and dysregulation of these genes is associated with autoimmune diseases. Differential expression levels of TNFAIP3, JUN, and TOB1 were validated by RT-qPCR. By flow cytometry, higher expression of CD69 and TNFSF8 was confirmed in CD4+ and CD8+T cells from PNH compared to healthy controls. Conclusion. Using RNA-seq, we identified novel molecular mechanisms and pathways which may underlie the aberrant T cell immune status in PNH. Specific dysregulation of T cell intracellular signaling may contribute to BM failure and the inflammatory environment in PNH. Understanding these pathways may provide new therapeutic strategies to modulate T cell immune responses in BM failure. Disclosures Hosokawa: Aplastic Anemia and MDS International Foundation: Research Funding. Rios:GSK/Novartis: Research Funding. Weinstein:GSK/Novartis: Research Funding. Townsley:GSK/Novartis: Research Funding.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 3058-3058 ◽  
Author(s):  
Fiorella Schischlik ◽  
Roland Jäger ◽  
Felix Rosebrock ◽  
Eva Hug ◽  
Michael Schuster ◽  
...  

Abstract Background Myeloproliferative Neoplasms (MPN) encompass several disease subgroups (ET, PV, and PMF) with distinct phenotypic and clinical features. Although the landscape of somatic mutations in MPN has been mapped in detail, the current therapies show limited efficacy in eliminating MPN cells. Immunotherapy for MPN have been suggested as an alternative to standard therapies, however, a broader view on the neoantigen landscape and MHC class I (MHCI) restriction of putative MPN neoantigens is missing. Aims We aimed to map the mutation and neoantigen landscape of MPN by analyzing whole transcriptome data and examine the applicability of targeted immunotherapy by analyzing co-occurrence of specific neoantigens with high affinity MHCI molecules in individual patients. Methods We have performed whole transcriptome sequencing of granulocytes for 113 MPN patients (32 ET, 17 PV, 55 PMF, and 9 post-MPN AML) and 15 controls. We implemented customized and established algorithms for genome wide fusion gene discovery, variant calling on 87 myeloid relevant genes, detection of splicing abnormalities and inferred these mutation classes directly from RNA-seq data for each patient. We extracted HLA genotypes from the same RNA-seq dataset. We examined the impact on protein sequence for each mutation class and predicted peptide affinities to MHCI. Top predicted peptide:MHCI interactions were validated in vitro using the UV-induced peptide exchange method. Results Following data processing and filtering, we identified a total of 13 gene fusions, 231 non-synonymous SNVs, and 21 Indels in 106/113 patients. Fusions were validated with Sanger sequencing, for SNVs and Indels, we re-sequenced DNA of 77 patients using a TruSight targeted sequencing panel. An unexpected high frequency of SF3B1 (0% ET, 14.6% PMF, 0% PV) hotspot mutations (K666N/R/T, K700E) was identified. We performed differential splice junction analysis comparing SF3B1 mutated and non-mutated patients and found 895 significantly differentially spliced junctions (750/895 were mediated through alternative 3'splicing). For each junction we calculated the percent splice in value (PSI) expressing the relative abundance of the alternative junction. We identified 20 junctions with retained intronic sequences with PSI>20%. For each of the 20 splicing defects we evaluated the impact of intron retention on the protein sequence. These splicing defects caused insertion of novel amino acids either in frame or through frameshifts. From all the mutations detected, we generated a virtual peptide library. In 113 patients, we detected 541 patient specific peptides predicted to bind patients' own MHCI proteins of which 102 had a predicted strong and 439 weak binding affinity. The recurrent CALR frameshifts and MPL-W515K/L/A mutations were also a rich resource for potential neoantigens due to their unique property to bind many shared common MHCI molecules (42, and 17 predicted for CALR and MPL, respectively). SF3B1 mutated patients had the highest number of predicted recurrent neoantigens with an average of 38 weak and 16 strong peptide/MHCI pairs. In total, we could identify 149 unique neoantigens covering 62% of MPN patients as a potential target in personalized cancer immunotherapy approaches. Of the 149 predicted unique neoantigens we tested 35 top ranking peptides for binding to recombinant MHCI monomers. We validated 70% of these peptides to be strong MHCI binders. Conclusions Our results provide a framework for systematic mining of neoantigens from different mutation classes using RNA analysis. The data presented in this study may serve as a resource for development of personalized vaccines or adoptive cell-base therapies, in particular for PMF patients positive for CALR, MPL and SF3B1 mutations. Disclosures Rosebrock: MyeloPro Diagnostics and Research GmbH: Employment. Hug:MyeloPro Diagnostics and Research GmbH: Employment. Haferlach:MLL Munich Leukemia Laboratory: Employment, Equity Ownership. Meggendorfer:MLL Munich Leukemia Laboratory: Employment. Stengel:MLL Munich Leukemia Laboratory: Employment. Gisslinger:Takeda: Consultancy, Honoraria; AOP Orphan: Consultancy, Honoraria, Research Funding; Janssen: Consultancy, Honoraria; Novartis: Consultancy, Honoraria, Research Funding; Celgene: Consultancy, Honoraria; Shire: Honoraria. Kralovics:MyeloPro Diagnostics and Research GmbH: Equity Ownership.


Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 9-10
Author(s):  
Yoshihiko Kihara ◽  
Marito Araki ◽  
Misa Imai ◽  
Yosuke Mori ◽  
Mei Horino ◽  
...  

Mutant calreticulin (CALR) has been shown to play a causal role in the development of essential thrombocythemia (ET) and primary myelofibrosis (PMF) via activation of the thrombopoietin receptor MPL. The oncogenic property of mutant CALR originates from a +1 frameshift mutation in its carboxyl-terminal domain, which is found in approximately 30% of patients with ET and PMF. Because the domain is uniquely found in mutant CALR, it has been recognized as a neoantigen and can therefore be used to target CALR-mutant cells using immunotherapy. In the present study, we found that a large portion of the domain generated by the frameshift in the mutant CALR was cleaved by an endoprotease belonging to the subtilisin family in multiple cell lines and primary cells. The cleaved form of mutant CALR was detected in the cell lysate; however, it was more abundant in the culture supernatant, implying that the cleavage occurred on the cell surface and/or outside the cells. Using mass spectrometric analysis, we determined the cleavage site of mutant CALR. To examine whether the cleavage was required for the oncogenic properties of mutant CALR, we introduced point mutations at the cleavage site. The mutant CALR construct that was resistant to protease cleavage exhibited full oncogenic capacity when expressed in UT-7/TPO cells. Consistent with this observation, chemical inhibition of the protease, which blocked the cleavage of mutant CALR, did not interfere with mutant CALR-dependent cell growth in UT-7/TPO cells. Next, we generated B3, a rat monoclonal antibody that recognized the mutant-specific sequence, even after cleavage. B3 recognized both the uncleaved and cleaved forms of mutant CALR by immunoblot in cell lysates prepared from the platelets and peripheral blood cells of CALR-mutant ET and PMF patients. B3 also recognized mutant CALR expressed on the cell surfaces of monocytes and granulocytes from CALR-mutant ET and PMF patients. Based on these results, we developed B3-chimera, a mouse chimeric antibody of B3, and evaluated its therapeutic potential for ET in vivo. We used an ET mouse model created by transplantation of LSK (lin-sca1+c-kit+) cells transduced with CALR del52 into the bone marrow. Intravenous injection of B3-chimera markedly suppressed the thrombocytosis induced by CALR del52, which was associated with a significant reduction in the megakaryocyte count in the bone marrow (Figure 1). In conclusion, we demonstrated that targeting the cleaved form of mutant CALR on the cell surface was a promising strategy for the treatment of CALR-mutant myeloproliferative neoplasms. Disclosures Akazawa: ITOCHU CHEMICAL FRONTIER Corporation: Membership on an entity's Board of Directors or advisory committees. Komatsu:Otsuka Pharmaceutical Co., Ltd., Shire Japan KK, Novartis Pharma KK, PharmaEssentia Japan KK, Fuso Pharmaceutical Industries, Ltd., Fujifilm Wako Pure Chemical Corporation, Chugai Pharmaceutical Co., Ltd., Kyowa Hakko Kirin Co., Ltd., Takeda Pharmaceutica: Research Funding; AbbVie: Other: member of safety assessment committee in M13-834 clinical trial.; PPMX: Consultancy, Research Funding; Takeda Pharmaceutical Co., Ltd, Novartis Pharma KK, Shire Japan KK: Speakers Bureau; Otsuka Pharmaceutical Co., Ltd., PharmaEssentia Japan KK, AbbVie GK, Celgene KK, Novartis Pharma KK, Shire Japan KK, Japan Tobacco Inc: Consultancy; Meiji Seika Pharma Co., Ltd.: Patents & Royalties: PCT/JP2020/008434, Research Funding.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 1663-1663
Author(s):  
Sabine Kayser ◽  
Maximilian Feszler ◽  
Julia Krzykalla ◽  
Matthias Schick ◽  
Thomas Hielscher ◽  
...  

Abstract Background:Recent publications suggest important roles of lysine methyltransferase 2C (KMT2C, located on 7q) and sprouty 4 (SPRY4, located on 5q) as candidate genes in leukemogenesis of acute myeloid leukemia (AML). The prognostic impact of the gene expression levels (ELs) of both genes on outcome in AML patients (pts) is currently unclear. Aim:To evaluate the prognostic impact of KMT2C and SPRY4 expression in correlation to clinical characteristics and genetic abnormalities assessed at diagnosis in a cohort of intensively treated adult AML pts. Methods: We retrospectively studied 268 AML pts (median age, 48 years; range, 17-60 years) who had been enrolled on 2 AML SHG trials (0295 and 0199, n=148; only normal cytogenetics pts (CN)) and the SAL-AML2003 trial (n=120; only abnormal cytogenetics pts (CA)). Acute promyelocytic and core-binding factor leukemia pts were excluded. Type of AML was de novo in 235 (88%), secondary in 19 (7%) and therapy-related in 14 (5%) of the 268 pts. Regarding baseline characteristics, CN pts had significantly higher white blood counts (WBC; p=0.001) and blast cells in peripheral blood (p=0.02) as compared to CA pts; all other factors were comparable. Cytogenetic analyses could be performed in 263 (98%) of the 268 pts. Cytogenetic risk classification according to ELN guidelines was intermediate-II in 55 (47%) and adverse in 63 (53%) of the CA pts, respectively. Abnormalities (abn) of 5q were present in 21 (18%) and abn of 7q in 16 (14%) of the CA pts. NPM1 and FLT3-ITD were analyzed in 145 (98%) of the CN pts. Of those, 59 (41%) were only NPM1 positive (pos), 12 (8%) were only FLT3-ITD pos, 34 (23%) were double pos and 40 (28%) were double negative (neg). KMT2C and SPRY4 ELs, normalized to ABL1and log2-transformed for analysis, were measured in triplets on cDNA obtained at diagnosis by RT-qPCR. Based on cDNA availability, KMT2C ELs could be analyzed in 143 (97%) of the CN and in all of the 120 CA pts, respectively. SPRY4 ELs could be measured in 30 (21%) of the CN and 107 (89%) of the CA pts, respectively. Results: KMT2C ELs were significantly lower in CN pts with de novo as compared to secondary AML (p= 0.02), whereas there was no difference in CA pts. No significant association was found for SPRY4 and type of AML. KMT2C ELs were significantly lower in FLT3-ITD pos as compared to FLT3-ITD neg CN pts (p=0.046), whereas there was no difference for SPRY4 ELs between the two groups (p=0.57). In addition, there was a significantly lower KMT2C expression in CN pts with intermediate-I risk as compared to NPM1 pos / FLT3-ITD neg pts (p=0.01). Regarding CA pts, there was no difference of KMT2C or SPRY4 ELs in adverse as compared to intermediate-II risk pts (p=0.08; p=0.20, respectively). When focusing on specific subgroups, KMT2C ELs were significantly lower in abn7q CA pts as compared to those without abn7q (p=0.002), whereas there was no difference of SPRY4 ELs in CA pts with or without abn5q (p=0.27). In univariate analysis higher SPRY4 ELs showed a significant favorable impact on relapse-free (RFS, p=0.03) and a trend towards a beneficial impact on overall survival (OS, p=0.06) for CA patients. A similar effect for KMT2C was not observed (RFS, p=0.96; OS, p=0.92). In subgroup analyses of pts with adverse risk cytogenetics, there was no impact of KMT2C or SPRY4 ELs on RFS (p=0.73; p=0.39) or OS (p=0.49; p=0.46), respectively. The same was true for FLT3-ITD pos CN pts (RFS, p=0.73; p=0.37; OS, p=0.91; p=0.36, respectively). In multivariate analyses on RFS and OS in CA pts including age, gender, KMT2C and SPRY4 ELs, logarithm of WBC, blast cells in bone marrow and cytogenetic risk group as variables, only higher age (OS, Hazard ratio (HR),1.28 per 10 years; 95%-confidence interval (CI): 1.02-1.59; p=0.03) and complex karyotype as compared to intermediate-II risk cytogenetics (RFS, HR: 2.25; 95%-CI: 1.20-4.22; p=0.01; OS, HR: 2.97; 95%-CI: 1.65-5.35; p<0.001) had an adverse impact. An effect of KMT2C or SPRY4 on RFS (p=0.84; p=0.16) or OS (p=0.85; p=0.45) was not found in the multivariate setting. In addition, in a multivariate model on CN pts (risk class according to NPM1 and FLT3-ITD mutational status instead of cytogenetic risk class) neither KMT2C nor SPRY4 had an impact on RFS (p=0.13; p=0.39, respectively) or OS (p=0.36; p=0.56, respectively). Conclusions:Lower KMT2C and SPRY4 ELs are associated with distinct genetic risk groups. An impact on prognosis was evident in univariable analyses for SPRY4 but not for KMT2C ELs in CA pts. Disclosures Kayser: Novartis: Consultancy. Platzbecker:Amgen: Honoraria, Research Funding; Novartis: Honoraria, Research Funding; TEVA Pharmaceutical Industries: Honoraria, Research Funding; Janssen-Cilag: Honoraria, Research Funding; Celgene Corporation: Honoraria, Research Funding. Heuser:Tetralogic: Research Funding; Novartis: Consultancy, Research Funding; Celgene: Honoraria; Bayer Pharma AG: Research Funding; Pfizer: Research Funding; Karyopharm Therapeutics Inc: Research Funding; BerGenBio: Research Funding. Thiede:AgenDix: Employment, Other: Ownership.


Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 31-31
Author(s):  
Nami Masubuchi ◽  
Marito Araki ◽  
Kaho Shioiri ◽  
Sota Kitazawa ◽  
Yoshihiko Kihara ◽  
...  

Iron incorporation through transferrin (Tf) and its receptor, TfR1, plays a critical role in erythropoiesis. We hypothesized that blockade of TfR1 may be a promising strategy to suppress erythrocytosis observed in polycythemia vera (PV). To examine this possibility, we developed PPMX-T003, a monoclonal antibody against human TfR1, through phage display screening. PPMX-T003 was observed to bind to TfR1 with a high affinity(Kd= 3.56×10-10)and presented strong ability to suppress the growth of PV model cell lines such as HEL (EC50= 0.07 nM). PPMX-T003 induced TfR1 internalization but blocked the recycling of the receptor presumably because it was strongly bound, even in the endosome, which switched the fate of TfR1 from recycling to degradation. Consequently, the PPMX-T003 treatment greatly reduced the intracellular iron concentration, along withcell cycle arrest in the G2/M phase in HEL cells. In contrast, PPMX-T003 failed to block cell proliferation in human umbilical vein endothelial cells that expressed a subtle level of TfR1. These findings suggest that PPMX-T003 is a potent inhibitor for cells whose growth is dependent on iron incorporation by TfR1. This led us to examine whether PPMX-T003 blocked the endogenous erythroid colony (EEC) formation in erythroid progenitor cells in PV. We observed that PPMX-T003, at a concentration of 40 ng/mL, exhibited complete blockade of EEC formation inJAK2 V617F-positive PV patients' cells. Although hydroxycarbamide (HU) and phlebotomy are well-established treatment strategies for the management of hematocrit in PV patients, there is a risk of secondary leukemia associated with HU treatment and side effects such as fatigue and restless leg syndrome due to phlebotomy-induced iron deficiency. Therefore, we suggest PPMX-T003, which may preferentially suppress erythrocytosis without changing the iron level in peripheral blood, as an alternative treatment for PV patients. Disclosures Ohira: PPMX: Current Employment, Current equity holder in private company. Nomura:PPMX: Current Employment. Matsuura:PPMX: Current Employment. Komatsu:Takeda Pharmaceutical Co., Ltd, Novartis Pharma KK, Shire Japan KK: Speakers Bureau; AbbVie: Other: member of safety assessment committee in M13-834 clinical trial.; PPMX: Consultancy, Research Funding; Meiji Seika Pharma Co., Ltd.: Patents & Royalties: PCT/JP2020/008434, Research Funding; Otsuka Pharmaceutical Co., Ltd., PharmaEssentia Japan KK, AbbVie GK, Celgene KK, Novartis Pharma KK, Shire Japan KK, Japan Tobacco Inc: Consultancy; Otsuka Pharmaceutical Co., Ltd., Shire Japan KK, Novartis Pharma KK, PharmaEssentia Japan KK, Fuso Pharmaceutical Industries, Ltd., Fujifilm Wako Pure Chemical Corporation, Chugai Pharmaceutical Co., Ltd., Kyowa Hakko Kirin Co., Ltd., Takeda Pharmaceutica: Research Funding.


2021 ◽  
Vol 8 (1) ◽  
Author(s):  
Yifang Yang ◽  
Junshu Wu ◽  
Defu Wu ◽  
Qi Wei ◽  
Tan Zhong ◽  
...  

Abstract Background The use of ocular hypotensive drugs has been reported to attenuate myopia progression. This study explores whether brimonidine can slow myopia progression in the guinea pig form-deprivation (FD) model. Methods Three-week-old pigmented male guinea pigs (Cavia porcellus) underwent monocular FD and were treated with 3 different methods of brimonidine administration (eye drops, subconjunctival or intravitreal injections). Four different concentrations of brimonidine were tested for intravitreal injection (2 μg/μL, 4 μg/μL, 20 μg/μL, 40 μg/μL). All treatments continued for a period of 21 days. Tonometry, retinoscopy, and A-scan ultrasonography were used to monitor intraocular pressure (IOP), refractive error and axial length (AL), respectively. On day 21, guinea pigs were sacrificed for RNA sequencing (RNA-seq) to screen for associated transcriptomic changes. Results The myopia model was successfully established in FD animals (control eye vs. FD eye, respectively: refraction at day 20, 0.97 ± 0.18 D vs. − 0.13 ± 0.38 D, F = 6.921, P = 0.02; AL difference between day 0 and day 21, 0.29 ± 0.04 mm vs. 0.45 ± 0.03 mm, F = 11.655, P = 0.004). Among the 3 different brimonidine administration methods, intravitreal injection was the most effective in slowing myopia progression, and 4 μg/μL was the most effective among the four different concentrations of brimonidine intravitreal injection tested. The AL and the refraction of the brimonidine intravitreal injection group was significantly shorter or more hyperopic than those of other 2 groups. Four μg/μL produced the smallest difference in AL and spherical equivalent difference values. FD treatment significantly increased the IOP. IOP was significantly lower at 1 day after intravitreal injections which was the lowest in FD eye of intravitreal injection of brimonidine. At day 21, gene expression analyses using RNA-seq showed upregulation of Col1a1 and Mmp2 expression levels by intravitreal brimonidine. Conclusions Among the 3 different administration methods, intravitreal injection of brimonidine was the most effective in slowing myopia progression in the FD guinea pig model. Intravitreal brimonidine at 4 μg/μL significantly reduced the development of FD myopia in guinea pigs. Expression levels of the Col1a1 and Mmp2 genes were significantly increased in the retinal tissues of the FD-Inj-Br group.


Sign in / Sign up

Export Citation Format

Share Document