Actionable Genetic Features of Primary Testicular and Primary Central Nervous System Lymphomas

Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 74-74
Author(s):  
Bjoern Chapuy ◽  
Margaretha GM Roemer ◽  
Yuxiang Tan ◽  
Chip Stewart ◽  
Liye Zhang ◽  
...  

Abstract Introduction. Primary testicular lymphoma (PTL) and primary central nervous system lymphoma (PCNSL) are large B-cell lymphomas (LBCL) that occur in immune privileged (IP) sites and share certain clinical and molecular features. To date, the treatment of these IP lymphomas is largely empiric and more effective targeted therapies are needed. Methods. To define actionable genetic features of IP lymphomas, we performed comprehensive genomic analyses of 21 PCNSLs and 7 PTLs and validated specific alterations in an independent cohort of 43 additional PTLs. Recurrent copy number alterations (CNAs) were detected using high-density single nucleotide polymorphism (SNP) arrays and the GISTIC algorithm and integrated with transcriptional profiles to identify candidate driver genes. Recurrent somatic mutations were identified using a combination of whole exome sequencing (WES) of paired tumor/normal samples and whole transcriptome sequencing (RNA-Seq) of the additional tumors without paired normal samples. Results. In systemic diffuse large B-cell lymphomas (DLBCLs), multiple low-frequency CNAs and associated target genes decrease p53 activity and perturb cell cycle regulation; infrequent somatic mutations of TP53 also deregulate these pathways (Cancer Cell, 2012; 22:359-372). In contrast, PCNSLs and PTLs primarily exhibit bi-allelic deletion of the upstream regulator of p53 activity and cell cycle, CDKN2A (~70% PCNSLs and ~80% of PTLs) and rarely have copy loss or somatic mutations of TP53 or CNAs of additional pathway components. The most commonly mutated genes in PCNSL and PTL, CD79B and MYD88, are also perturbed in a subset of systemic DLBCLs. However, mutations of these two genes are much more frequent in IP lymphomas (70% MYD88 and 61% CD79B of analyzed PCNSLs and PTLs) and these alterations are commonly found in the same cases (57% of cases in this series). These data indicate that concurrent oncogenic activation of the B-cell receptor (BCR) and the Toll-like receptor (TLR) signaling pathways is a characteristic feature of IP lymphomas with implications for targeted therapies. Among the IP lymphomas, ~20% of PCNSLs and ~40% PTLs exhibit 3q12.3/NFKBIZ copy gain and increased expression of the NFKBIZ protein product, IκB-ζ, an atypical IκB family member induced by TLR signaling. In our PTL series, MYD88 wild-type tumors had the highest 3q12.3/NFKBIZ copy gains, and ~90% of all analyzed PTLs had structural bases for NFκB activation via the TLR pathway. Lentiviral-mediated IκB-ζ knockdown decreased expression of the IκB-ζ target genes, IκB-α and BCL-xL, and induced apoptosis of LBCL cell lines with MYD88 L265P mutations, NFKBIZ gain or both alterations. In addition, enforced expression of NFKBIZ enhanced the growth of LBCLs with normal NFKBIZ copy numbers. Taken together, these data suggest that many IP lymphomas depend upon oncogenic MYD88/NFKBIZ signaling. Although the majority of CNAs and somatic mutations were shared by PCNSLs and PTLs, certain alterations were primarily observed in PTL. In both the initial and independent validation series, > 40% of PTLs exhibited copy gain of chromosome 9p24.1/CD274 (PD-L1) / PDCD1LG2 (PD-L2) and associated overexpression of the PD-1 ligands. These observations were of particular interest because 9p24.1 copy gain is a characteristic abnormality in two additional lymphoid malignancies, primary mediastinal LBCL and classical Hodgkin lymphoma, PD-1 signaling promotes tumor immune evasion and the PD-1 pathway is targetable. We also identified one PTL in which a novel translocation juxtaposed the regulatory elements of TBL1XR1 (chromosome 3) to the start codon-bearing exon 2 of PDCD1LG2 (PD-L2) (chromosome 9). This translocation, which was detected by RNA-Seq and confirmed by 5’ RACE and a newly developed split-apart FISH assay, resulted in dramatic overexpression of the PD-L2 protein. These data suggest that PTLs utilize several genetic mechanisms to deregulate the PD-1 ligands and limit anti-tumor immunity. Conclusions. Integrative and comparative genomic studies define PCNSL and PTL as related but unique lymphoid malignancies with targetable genetic alterations, and associated p53 deficiency and cell cycle deregulation, concurrent oncogenic BCR and TLR signaling and PD-1 dependent immune evasion that warrant further clinical investigation. Note: B.C. and M.G.M.R have made equal contributions to this research. Disclosures Feuerhake: Roche Pharma Research and Early Development (pRED) from 2008-2012: Employment. Freeman:Merck: on the PD-1 pathway Patents & Royalties; EMD-Serrono: on the PD-1 pathway Patents & Royalties; Boehringer-Ingelheim: on the PD-1 pathway Patents & Royalties; Amplimmune: on the PD-1 pathway Patents & Royalties; Roche: on the PD-1 pathway Patents & Royalties; Bristol-Myers-Squibb: on the PD-1 pathway Patents & Royalties; Novatis: on the PD-1 pathway, on the PD-1 pathway Patents & Royalties. Shipp:Sanofi: Research Funding; Bayer: Membership on an entity's Board of Directors or advisory committees; Gilead: Consultancy, Membership on an entity's Board of Directors or advisory committees; Bristol-Myers-Squibb: Membership on an entity's Board of Directors or advisory committees, Research Funding; Pharmacyclics: Membership on an entity's Board of Directors or advisory committees; Merck: Membership on an entity's Board of Directors or advisory committees; Janssen R&D: Membership on an entity's Board of Directors or advisory committees.

Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 874-874
Author(s):  
Jonathan C Poe ◽  
Dadong Zhang ◽  
Jichun Xie ◽  
Rachel A. DiCioccio ◽  
Xiaodi Qin ◽  
...  

While B cells are known to contribute to the pathogenesis of chronic graft-versus-host disease (cGVHD) in mice, it has been challenging to elucidate intrinsic mechanisms of tolerance loss in patients. To identify distinct and potentially targetable B-cell subsets in cGVHD, we employed single-cell RNA-Seq along with an unsupervised hierarchical clustering analysis, targeting 10,000 single B cells from each of eight patients who were >12 months post-allogeneic hematopoietic stem cell transplantation (HCT) and either had active cGVHD manifestations (n=4) or never developed cGVHD (n=4). Bioinformatics analysis of pooled cell data (using R with Seurat extension package) identified 6 major B cell clusters common to all patients (Figure 1A). "Intra-cluster" gene comparison (using R package DESeq2, false-discovery rate 0.05) revealed numerous differentially expressed genes between patient groups. The greatest number of differentially-expressed genes occurred in a cluster referred to herein as 'Cluster 6' (Figure 1A, in yellow with asterisk). Within Cluster 6, B cells from active cGVHD patients expressed significantly increased ITGAX (CD11c, Padj =0.007), TNFRSF13B (TACI, a receptor for BAFF, Padj =0.003), IGHG1 (IgG1, Padj =9.3e-06) and IGHG3 (IgG3, Padj =1.7e-12), along with 44 additional genes (to be discussed). Thus, Cluster 6 in cGVHD patients may represent a CD11cpos, BAFF-responsive B cell subset primed to undergo isotype switching in response to alloantigen. Flow cytometry analysis on PBMCs from an independent HCT patient cohort (n=10) confirmed that CD11cpos B cells were indeed significantly expanded in cGVHD (P < 0.01, Figure 1B), and revealed these B cells were also TACIpos, CD19high, forward scatter high (FSChigh) blast-like cells (Figure 1C). We found that these CD11cpos B cells had mixed expression of CD21, CD27, IgD and CD24 (Figure 1C). Remarkably, other recent studies on bulk patient B cells have suggested that similar CD11cposCD21negCD19highT-BETpos cells are critical drivers of humoral autoimmunity in diseases including systemic lupus erythematosus (SLE; Scharer et al. 2019; Rubtsova et al. 2017; Rubtsov et al. 2011). This subset now identified by single-cell RNA-Seq is consistent with a population of TACIhigh B cells that produced IgG in response to BAFF treatment ex vivo (Sarantopoulos 2009). Data suggest we have identified functionally distinct and potentially targetable B cell subpopulations. We are employing functional assays to determine whether the additional molecular pathways now elucidated account for our previous work showing greater ex vivo B cell survival rates and hyper-responsiveness to surrogate antigen (Allen et al. 2012, 2014), certain TLR agonists (Suthers et al. 2017), and NOTCH ligand (Poe et al. 2017). In addition to more deeply characterizing B-cell subsets in cGVHD, our single-cell RNA-Seq analyses identified several genes significantly altered across multiple B cell clusters in the cGVHD group, implicating more broad alterations of some genes in this disease. Among these is CKS2, a critical cell cycle regulator, which was significantly increased in cGVHD B cells (Padj 1.0e-10 to 0.018, depending on the cluster evaluated). Increased CKS2 expression was validated by qPCR analysis on B cells from a separate HCT patient cohort with or without cGVHD (P < 0.001, Figure 1D), suggesting that the majority of cGVHD B cells are primed to enter the cell cycle at multiple stages of differentiation when exposed to the proper stimuli. In summary, we used an unbiased approach to identify and further characterize an extrafollicular CD11cposTACIposCD19high B cell population in cGVHD patients that appears to be activated and undergoing active IgG isotype switching. This plasmablast-like B cell population is potentially amenable to therapeutic intervention to prevent pathogenic antibody production. Importantly, we also identify gene alterations across the cGVHD peripheral B cell compartment that potentially underpin promotion of hyperactivated B cells in this disease. Therapeutic strategies to target these pathways will also be discussed. This work was supported by a National Institutes of Health grant, R01HL129061. Disclosures Ho: Omeros Corporation: Membership on an entity's Board of Directors or advisory committees; Jazz Pharmaceuticals: Research Funding; Jazz Pharmaceuticals: Consultancy. Horwitz:Abbvie Inc: Membership on an entity's Board of Directors or advisory committees. Rizzieri:Celgene, Gilead, Seattle Genetics, Stemline: Other: Speaker; AbbVie, Agios, AROG, Bayer, Celgene, Gilead, Jazz, Novartis, Pfizer, Sanofi, Seattle Genetics, Stemline, Teva: Other: Advisory Board; AROG, Bayer, Celgene, Celltron, Mustang, Pfizer, Seattle Genetics, Stemline: Consultancy; Stemline: Research Funding.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 1527-1527
Author(s):  
Sara Rodríguez ◽  
Cirino Botta ◽  
Jon Celay ◽  
Ibai Goicoechea ◽  
Maria J Garcia-Barchino ◽  
...  

Background: Although MYD88 L265P is highly frequent in WM, by itself is insufficient to explain disease progression since most cases with IgM MGUS also have mutated MYD88. In fact, the percentage of MYD88 L265P in CD19+ cells isolated from WM patients is typically <100%, which questions if this mutation initiates the formation of B-cell clones. Furthermore, a few WM patients have detectable MYD88 L265P in total bone marrow (BM) cells and not in CD19+ selected B cells, raising the possibility that other hematopoietic cells carry the MYD88 mutation. However, no one has investigated if the pathogenesis of WM is related to somatic mutations occurring at the hematopoietic stem cell level, similarly to what has been shown in CLL or hairy cell leukemia. Aim: Define the cellular origin of WM by comparing the genetic landscape of WM cells to that of CD34 progenitors, B cell precursors and residual normal B cells. Methods: We used multidimensional FACSorting to isolate a total of 43 cell subsets from BM aspirates of 8 WM patients: CD34+ progenitors, B cell precursors, residual normal B cells (if detectable), WM B cells, plasma cells (PCs) and T cells (germline control). Whole-exome sequencing (WES, mean depth 74x) was performed with the 10XGenomics Exome Solution for low DNA-input due to very low numbers of some cell types. We also performed single-cell RNA and B-cell receptor sequencing (scRNA/BCRseq) in total BM B cells and PCs (n=32,720) from 3 IgM MGUS and 2 WM patients. Accordingly, the clonotypic BCR detected in WM cells was unbiasedly investigated in all B cell maturation stages defined according to their molecular phenotype. In parallel, MYD88p.L252P (orthologous position of the human L265P mutation) transgenic mice were crossed with conditional Sca1Cre, Mb1Cre, and Cγ1Cre mice to selectively induce in vivo expression of MYD88 mutation in CD34 progenitors, B cell precursors and germinal center B cells, respectively. Upon immunization, mice from each cohort were necropsied at 5, 10 and 15 months of age and screened for the presence of hematological disease. Results: All 8 WM patients showed MYD88 L265P and 3 had mutated CXCR4. Notably, we found MYD88 L265P in B cell precursors from 1/8 cases and in residual normal B cells from 3/8 patients, which were confirmed by ASO-PCR. In addition, CXCR4 was simultaneously mutated in B cell precursors and WM B cells from one patient. Overall, CD34+ progenitors, B-cell precursors and residual normal B cells shared a median of 1 (range, 0-4; mean VAF, 0.16), 2 (range, 1-5; mean VAF, 0.14), and 4 (range, 1-13; mean VAF, 0.26) non-synonymous mutations with WM B cells. Some mutations were found all the way from CD34+ progenitors to WM B cells and PCs. Interestingly, concordance between the mutational landscape of WM B cells and PCs was <100% (median of 85%, range: 25%-100%), suggesting that not all WB B cells differentiate into PCs. A median of 7 (range, 2-19; mean VAF, 0.39) mutations were unique to WM B cells. Accordingly, many clonal mutations in WM B cells were undetectable in normal cells. Thus, the few somatic mutations observed in patients' lymphopoiesis could not result from contamination during FACSorting since in such cases, all clonal mutations would be detectable in normal cells. Of note, while somatic mutations were systematically detected in normal cells from all patients, no copy number alterations (CNA) present in WM cells were detectable in normal cells. scRNA/BCRseq unveiled that clonotypic cells were confined mostly within mature B cell and PC clusters in IgM MGUS, whereas a fraction of clonotypic cells from WM patients showed a transcriptional profile overlapping with that of B cell precursors. In mice, induced expression of mutated MYD88 led to a moderate increase in the number of B220+CD138+ plasmablasts and B220-CD138+ PCs in lymphoid tissues and BM, but no signs of clonality or hematological disease. Interestingly, such increment was more evident in mice with activation of mutated MYD88 in CD34+ progenitors and B-cell precursors vs mice with MYD88 L252P induced in germinal center B cells. Conclusions: We show for the first time that WM patients have somatic mutations, including MYD88 L265P and in CXCR4, at the B cell progenitor level. Taken together, this study suggests that in some patients, WM could develop from B cell clones carrying MYD88 L265P rather than it being the initiating event, and that other mutations or CNA are required for the expansion of B cells and PCs with the WM phenotype. Disclosures Roccaro: Janssen: Membership on an entity's Board of Directors or advisory committees; Celgene: Membership on an entity's Board of Directors or advisory committees; Celgene: Membership on an entity's Board of Directors or advisory committees; Transcan2-ERANET: Research Funding; AstraZeneca: Research Funding; European Hematology Association: Research Funding; Transcan2-ERANET: Research Funding; Associazione Italiana per al Ricerca sul Cancro (AIRC): Research Funding; Associazione Italiana per al Ricerca sul Cancro (AIRC): Research Funding; European Hematology Association: Research Funding; Janssen: Membership on an entity's Board of Directors or advisory committees; AstraZeneca: Research Funding; Amgen: Membership on an entity's Board of Directors or advisory committees. San-Miguel:Amgen, Bristol-Myers Squibb, Celgene, Janssen, MSD, Novartis, Roche, Sanofi, and Takeda: Consultancy, Honoraria. Paiva:Amgen, Bristol-Myers Squibb, Celgene, Janssen, Merck, Novartis, Roche, and Sanofi; unrestricted grants from Celgene, EngMab, Sanofi, and Takeda; and consultancy for Celgene, Janssen, and Sanofi: Consultancy, Honoraria, Research Funding, Speakers Bureau.


Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 3580-3580
Author(s):  
Ilaria Iacobucci ◽  
Heike Pfifer ◽  
Annalisa Lonetti ◽  
Cristina Papayannidis ◽  
Anna Ferrari ◽  
...  

Abstract Abstract 3580 Introduction: Although treatment with tyrosine kinase inhibitors (TKIs) has revolutionized the management of adult patients with BCR-ABL1 -positive acute lymphoblastic leukemia (ALL) and significantly improved response rates, relapse is still an expected and early event in the majority of them. It is usually attributed to the emergence of resistant clones with mutations in BCR-ABL1 kinase domain or to BCR-ABL1 -independent pathways but many questions remain unresolved about the genetic abnormalities responsible for relapse after TKI and chemotherapy-based regimens. Patients and methods: In an attempt to better understand the genetic mechanisms responsible for this phenomenon, we have analyzed matched diagnosis-relapse samples from 30 adult BCR-ABL1 -positive ALL patients using high resolution Affymetrix single nucleotide polymorphism (SNP) arrays (GeneChip® Human Mapping 250K NspI, n=15 pairs and Genome-Wide Human SNP 6.0, n= 15 pairs). Genetic differences were analyzed in terms of copy number changes and loss of heterozygosity (LOH) events. 20 patients were enrolled in clinical trials of GIMEMA AL Working Party and treated with imatinib alone or in combination with conventional chemotherapy (40%) or dasatinib as frontline therapy (60%). The median age at diagnosis was 54 years (range 23–74) and the median blast cell count was 97% (range 60–99). The median time to relapse was 27 months (range, 9–104). 10 patients were treated according to the GMALL trials, a high-dose chemotherapy based protocol in combination with imatinib. The median age at diagnosis was 65 years (range 19–79) and the median leucocyte count was 37300/μl (range 5000 – 220000/μl). The median time to relapse was 9.8 months (range, 3 – 25). Results: First, we compared diagnosis and relapse samples for the presence of macroscopic (> 1.5 MB) copy number alterations (CNA). Novel acquired macroscopic CNAs were detected in 7/20 (35%) TKI relapse cases and included losses of 3p12-p14, 5q34, 9q34, 10q24 and 12p13-p12 and gains of 1q, 9q34-q33 and 22q and in 4/10 (40%) chemotherapy-relapse cases and included losses of 9p21 and 12q21–22 and gains of all chromosome 8 or part of it in 2 patients. Since no common patterns of acquired alterations were observed, it is likely that relapse may be due to a more generalized genetic instability rather than to specific mechanisms. Moreover, chemotherapy did not select resistant clones with higher number of alterations. 8/20 (40%) TKI resistant cases and 4/10 chemotherapy resistant patients harbored the same CNAs present in the matched diagnosis sample (losses of 9p21 in 7 cases, 7p and 22q11 in single cases and gains of chromosomes 1q, 4, 8q, 17q and 21), indicating a common clonal origin. In contrast, in 5/20 (25%) TKI resistant cases and 4/10 (40%) chemotherapy resistant patients macroscopic CNAs present at diagnosis were lost at relapse (losses of chromosomes 7, 11q, 14q, 15q, 16q and 19p and gains of 5q, 8q, 9q34 and 22q11). Thereafter, we compared diagnosis and relapse samples for microscopic CNAs (< 1.5 MB). The alteration most frequently acquired at relapse was loss of the tumor suppressor CDKN2A (53% vs 33 % of diagnosis). Other common acquired CNAs at relapse included gains of ABC transporter genes, such as ABCC1, ABCC6 (1q41) and BCL8 (15q11); losses affected EBF1 (5q33) and IGLL3 (22q11) genes involved in B-cell development, BTG1 (12q21) involved in cell cycle regulation and CHEK2 (22q12) involved in DNA repair. Interestingly, for all relapse cases analysis of IKZF1 deletions, identified in 80% of patients, demonstrated a clonal relationship between diagnostic and relapse samples, suggesting that this alteration is not acquired with relapse but it is maintained with fidelity from diagnosis working as a marker of disease. The majority (92%) of relapse samples harbored at least some of the CNAs present in the matched diagnosis sample, indicating a common clonal origin. Conclusions: Genomic copy number changes evolving from diagnosis to relapse have been identified demonstrating that a diversity of alterations contributes to relapse and with the most common alterations targeting key regulators of tumor suppression, cell cycle control, and lymphoid/B cell development. Supported by European LeukemiaNet, AIL, AIRC, Fondazione Del Monte di Bologna e Ravenna, FIRB 2006, PRIN 2009, Ateneo RFO grants, PIO program, Programma di Ricerca Regione – Università 2007 – 2009. Disclosures: Soverini: Novartis: Consultancy; ARIAD: Consultancy; Bristol-Myers Squibb: Consultancy. Baccarani:Pfizer Oncology: Consultancy; Novartis: Consultancy; BMS: Consultancy; Ariad: Consultancy; Novartis: Research Funding; Pfizer Oncology: Honoraria; Novartis: Honoraria; BMS: Honoraria; Ariad: Honoraria; Novartis: Membership on an entity's Board of Directors or advisory committees; BMS: Membership on an entity's Board of Directors or advisory committees; Ariad: Membership on an entity's Board of Directors or advisory committees. Ottmann:Novartis Corporation: Consultancy, Honoraria, Research Funding, Speakers Bureau. Martinelli:Novartis: Consultancy, Honoraria; BMS: Consultancy, Honoraria; Pfizer: Consultancy.


Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 18-19
Author(s):  
Morten P Oksvold ◽  
Ulrika Warpman Berglund ◽  
Helge Gad ◽  
Baoyan Bai ◽  
Trond Stokke ◽  
...  

Although chemo-immunotherapy has improved survival in B-cell lymphoma patients, refractory and relapsed disease still represents a major challenge, urging for development of new therapeutics. A new approach is to target nucleotide metabolism. Karonudib (TH1579), was developed to inhibit MutT-homologue-1/Nudix hydrolase 1 (MTH1/NUDT1), an enzyme that prevents oxidized nucleotides to be incorporated into DNA. Under normal conditions with low reactive oxygen species (ROS) burden, MTH1 is not essential for cell survival. This contrasts cancer cells which frequently upregulate MTH1 to compensate for increased ROS with corresponding higher oxidized nucleotide levels, and therefore become more susceptible for MTH1 inhibition. Here, our aim was to perform preclinical testing of karonudib in B-cell lymphoma. Using two different gene expression datasets, we demonstrate that NUDT1, the gene encoding MTH1, was highly upregulated in tumor biopsies from patients with diffuse large B-cell lymphoma (DLBCL) and Burkitt's lymphoma as compared to follicular lymphoma and peripheral blood B cells from healthy donors, hence demonstrating a rationale for targeting MTH1 in aggressive B-cell lymphoma. We tested the efficacy of karonudib (0.06-1 µM) in vitro in a range of B-cell lymphoma cell lines using CellTiterGlo and by flow cytometry detection of active caspase-3 and TUNEL to identify apoptotic cells. Karonudib strongly reduced viability in all B-cell lymphoma cell lines tested (n = 12) and induced apoptosis at concentrations well tolerated by peripheral blood B cells from healthy donors. Cell cycle analysis and microscopy revealed that most cells arrested in prometaphase in the presence of karonudib. Failed spindle assembly led to mitotic arrest and subsequent apoptosis. Prometaphase arrest was seen in TP53 mutant as well as in TP53 wild type cell lines, confirming that karonudib induced apoptosis independent of TP53 mutational status. To test the efficacy of karonudib in vivo, we utilized two different lymphoma xenograft models, including an ABC DLBCL patient-derived model. Mice were treated with karonudib (90 mg/kg) or vehicle b.i.d, three times a week and tumor growth was monitored by in vivo imaging system or MR. In both models, karonudib as single agent completely controlled tumor growth, and significantly prolonged survival (p&lt;0.0001, as compared to control mice). The specificity of MTH1 inhibitors has been debated and TH588, the first generation MTH1 inhibitor, was recently shown to bind b-tubulin and induce mitotic arrest in MTH1 knock out cell lines (Patterson et al, Cell Syst 2019). To elucidate the mechanism of karonudib in B-cell lymphoma, we generated MTH1 knock out cells using CRISPR/Cas9, and compared the functional effects of karonudib in these clones with the original lymphoma cells. We demonstrated on-target effect of the inhibitor as the MTH1 knock out clones were less sensitive towards karonudib. However, MTH1 knock out clones exhibited a similar cell cycle arrest as the wild type cells after karonudib treatment. This clearly indicates that karonudib can induce cell cycle arrest independent of MTH1, and hence has a dual mechanism of action. Our preclinical data suggest that karonudib is a promising drug with potential therapeutic use in B-cell lymphoma, and may be particular effective in aggressive lymphoma types. Disclosures Warpman Berglund: Oxcia AB: Other: shareholders; non profit Thomas Helleday Foundation for Medical Research: Membership on an entity's Board of Directors or advisory committees, Patents & Royalties. Gad:Oxcia: Other: shareholder; non profit Thomas Helleday Foundation for Medical Research: Membership on an entity's Board of Directors or advisory committees, Patents & Royalties. Pham:Oxcia AB: Other: Shareholder. Sanjiv:Oxcia AB: Other: Shareholder; non profit Thomas Helleday Foundation for Medical Research: Membership on an entity's Board of Directors or advisory committees, Patents & Royalties. Helleday:Oxcia AB: Other: Shareholder; non profit Thomas Helleday Foundation for Medical Research: Membership on an entity's Board of Directors or advisory committees, Patents & Royalties.


Blood ◽  
2015 ◽  
Vol 126 (23) ◽  
pp. 3922-3922
Author(s):  
Bjoern Chapuy ◽  
Andrew J Dunford ◽  
Chip Stewart ◽  
Atanas Kamburov ◽  
Jaegil Kim ◽  
...  

Abstract Diffuse large B-cell lymphoma (DLBCL) is a genetically heterogeneous disease characterized by multiple low-frequency alterations including somatic mutations, copy number alterations (CNAs) and chromosomal rearrangements. We sought to identify previously unrecognized low-frequency genetic events, integrate recurrent alterations into comprehensive signatures and associate these signatures with clinical parameters. For these reasons, our multi-institutional international group assembled a cohort of 304 primary DLBCLs from newly diagnosed patients, 87% of whom were uniformly treated with state-of-the-art therapy (rituximab-containing CHOP regimen) and had long term followup. Tumors were subjected to whole exome sequencing with an extended bait set that included custom probes designed to capture recurrent chromosomal rearrangements. In this cohort, 47% of samples had available transcriptional profiling and assignment to associated disease subtypes. Analytical pipelines developed at the Broad Institute were used to detect mutations (MuTect), CNAs (Recapseq+Allelic Capseq) and chromosomal rearrangements (dRanger+Breakpointer) and assess clonality (Absolute). To analyze formalin-fixed paraffin-embedded tumors without paired normals we developed a method which utilized 8334 unrelated normal samples to stringently filter recurrent germline events and artifacts. Significant mutational drivers were identified using the MutSig2CV algorithm and recurrent CNAs were assessed with GISTIC2.0. In addition, we utilized a recently developed algorithm, CLUMPS2, to prioritize somatic mutations which cluster in 3-dimensional protein structure. With this approach, we identified > 90 recurrently mutated genes, 34 focal amplifications and 41 focal deletions, 20 arm-level events and > 200 chromosomal rearrangements in the DLBCL series. Of note, 33% of the mutational drivers were also perturbed by chromosomal rearrangements or CNAs, underscoring the importance of a comprehensive genetic analysis. In the large DLBCL series, we identified several previously unrecognized but potentially targetable alterations including mutations in NOTCH2 (8%) and TET2 (5%). The majority of identified chromosomal rearrangements involved translocations of potent regulatory regions to intact gene coding sequences. The most frequently rearrangements involved Ig regulatory elements which were translocated to BCL2, MYC, BCL6 and several additional genes with known roles in germinal center B-cell biology. After identifying recurrent somatic mutations, CNAs and chromosomal rearrangements, we performed hierarchical clustering and identified subsets of DLBCLs with comprehensive signatures comprised of specific alterations. A large subset of tumors shared recurrent alterations previously associated with follicular lymphoma including mutations of chromatin modifiers such as CREBBP, MLL2, and EZH2 in association with alterations of TNFRSF14 and GNA13 and translocations of BCL2. This cluster was enriched in GCB-type DLBCLs and contained a subset with select genetic alterations associated with an unfavorable outcome. An additional cohort of tumors was characterized by alterations perturbing B-cell differentiation including recurrent BCL6 translocations or alterations of PRDM1. A subset of these DLBCLs had alterations of NOTCH2 and additional pathway components or mutations of MYD88 in association with TNFAIP3, CD70 and EBF1, a master regulator of B-cell differentiation. An additional group of DLBCLs exhibited frequent MYD88 mutations in association with alterations of CD79B, PIM1, TBL1XR1 and ETV6 and BCL2 copy gain; these tumors were highly enriched for ABC-type DLBCLs. This coordinate signature and additional alterations of p53 pathway components were associated with outcome. We explored bases for the identified genetic alterations in DLBCL by performing an in silico mutational signature analysis. The most frequent mutational signatures were those of spontaneous deamination (aging) and AID with rare cases of microsatellite instability. We also assessed the clonality of identified genetic features to define cancer cell fraction and establish the timing of specific genetic events. The comprehensive genetic signatures of clinically annotated DLBCLs provide new insights regarding approaches to targeted therapy. Disclosures Link: Kite Pharma: Research Funding; Genentech: Consultancy, Research Funding. Rodig:Perkin Elmer: Membership on an entity's Board of Directors or advisory committees; BMS: Research Funding. Pfreundschuh:Boehringer Ingelheim, Celegene, Roche, Spectrum: Other: Advisory board; Roche: Honoraria; Amgen, Roche, Spectrum: Research Funding. Shipp:Gilead: Consultancy; Sanofi: Research Funding; BMS: Membership on an entity's Board of Directors or advisory committees, Research Funding; Merck: Membership on an entity's Board of Directors or advisory committees; Bayer: Membership on an entity's Board of Directors or advisory committees, Research Funding.


Blood ◽  
2015 ◽  
Vol 126 (23) ◽  
pp. 306-306 ◽  
Author(s):  
Francine E. Garrett-Bakelman ◽  
Sheng Li ◽  
Stephen S. Chung ◽  
Todd Hricik ◽  
Rapaport Franck ◽  
...  

Abstract Acute Myeloid Leukemia (AML) remains a clinical challenge, with most patients dying of relapsed disease. The complete biological basis of relapse remains unclear. Genetic lesions and heterogeneity have been proposed as key drivers of clinical outcome, yet do not fully explain leukemia relapse. Epigenomic dysregulation is a hallmark of newly diagnosed AML. Plasticity is a core property of the epigenome, enabling cells to adapt to stressful conditions, independent of genetic alterations. Hence we asked whether epigenomic plasticity might contribute to AML progression, have functional consequences and be independent of genetic influences in AML (a question that has not been addressed for any tumor type). Methods. We formed an international consortium to collect and profile paired diagnosis and relapse AML specimens. We extracted DNA and RNA from 138 clinically annotated AML patient samples. We obtained matched germline DNA as genetic controls, and fourteen normal CD34+ specimens as DNA methylation and transcriptome controls. We performed methylome sequencing (ERRBS), genomic sequencing (exomes and targeted resequencing) and transcriptomic (RNA-seq) profiling. For a single patient, more intensive multi-layer profiling (whole genome sequencing, ERRBS, RNA-seq and single cell RNA-seq) was performed at five serial time points. We quantified epigenetic allelic heterogeneity (epialleles) using a novel approach that employs entropy equations (MethClone), and validated epiallele composition using orthogonal methods. Some of the major conclusions are: 1) Epigenetic allelic diversity is an independent variable linked to clinical outcome. Statistically significant epiallele shift (ΔS <-90) was detected at thousands of genomic loci (eloci) at diagnosis. High eloci burden correlated (Wilcoxon test) with a shorter relapse free probability in the entire cohort (p = 0.043) and in intermediate-risk patients based on the Medical Research Council (p= 0.016) and European Leukemia Net (p=0.057) criteria. Multivariate analysis using Cox proportional hazards regression model revealed that the epiallele burden was an independent variable correlated with relapse free survival (p = 0.021). 2) Promoter epialleles are linked to hypervariable transcriptional regulation. We observed substantial change in epiallele burden at relapse versus diagnosis. A subset of the eloci localized to gene promoters. High promoter epiallele variance was significantly associated with high transcriptional variance (p<0.001) based on RNA-seq, including genes that were significantly differentially expressed at relapse. Deconvolution of leukemia blast populations using Single Cell RNA-seq confirmed that the presence of promoter epialleles was linked to hypervariable transcriptional states (p<0.001). 3) AML patients can be classified according to epigenetic allele progression at relapse. K-means clustering based on epiallele shift at diagnosis versus relapse distributed patients into three classes: those with reduced, increasing or stable epiallele burden. Strikingly, there was no correlation between epiallele changes and the patterns of genomic evolution. Furthermore, there was no correlation between epiallele patterns acquired with mutations in epigenetic modifiers or other recurrently mutated genes in AML. 4) Epigenetic heterogeneity upon disease relapse is divergent from the genetic landscape. Integrating whole genome sequencing and methylome analysis we observed that a) significant increases in epigenetic heterogeneity precede significant changes in the abundance of somatic mutations; b) whereas a high number of somatic mutations were shared across all time points, epialleles exhibited dominance of distinct and unique eloci at each time point; and c) the variant epiallele frequency decreased earlier in progression than somatic mutation variant allele frequency, suggesting that epigenetic clonal diversification can precede genetic clonal evolution. Summary. Based on our results we propose that epigenetic allele diversity allows populations of leukemia cells to sample transcriptional states more freely thus creating the potential for greater evolutionary fitness. This provides an additional independent mechanism of plasticity that can explain the resilient nature of AML to adapt and survive exposure to chemotherapy drugs, independent of genetic heterogeneity. Disclosures Perl: Actinium Pharmaceuticals: Consultancy; Asana Biosciences: Consultancy; Arog Pharmaceuticals: Consultancy; Ambit/Daichi Sankyo: Consultancy; Astellas US Pharma Inc.: Consultancy. Becker:Millenium: Research Funding. Lewis:Roche: Honoraria, Other: Travel; Amgen: Other: Travel. Levine:Loxo Oncology: Membership on an entity's Board of Directors or advisory committees; CTI BioPharma: Membership on an entity's Board of Directors or advisory committees; Foundation Medicine: Consultancy.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 2839-2839 ◽  
Author(s):  
Anagha Deshpande ◽  
Benson Chen ◽  
Parham Ramezani-Rad ◽  
Alessandro Pastore ◽  
Luyi Zhao ◽  
...  

Abstract Aberrant activation of the MYC proto-oncogene is a recurrent feature in human B-cell lymphomas of diverse sub-types, correlating with adverse prognosis and therapy resistance. Direct pharmacological MYC-targeting has proved difficult, but recent studies have shown that targeting chromatin regulators critical for MYC-driven oncogenesis may provide alternative avenues for therapeutic intervention. Recently, it has been demonstrated that MYC-driven oncogenesis in certain solid tumors is dependent on the histone 3 lysine 79 (H3K79) methyltransferase DOT1L. We hypothesized that B-cell lymphomas with hyperactive MYC-signaling might be responsive to DOT1L inhibition. In order to test this hypothesis, we tested the effect of the DOT1L inhibitor Pinometostat (EPZ-5676) on a panel of human B-cell lymphoma cell lines featuring elevated MYC. Pinometostat treatment reduced global H3K79 methylation levels, accompanied by a time and dose-dependent decrease in proliferation of several Burkitt's lymphoma cell lines including P493-6, Daudi and Raji. We observed that key MYC-target genes including CDK4, PPAT and NPM1 were significantly downregulated upon Pinometostat treatment, suggesting that DOT1L is required for the transcriptional activation of MYC-target genes in these cells. Pinometostat-treated B-lymphoma cells showed a significant decrease of cells in S-phase compared to controls as assessed by BrdU-labeling assays. Similar results were also obtained in a panel of B-cell lymphoma cell lines with MYC-rearrangements including mantle cell lymphoma (MCL) cell lines Jeko-1, JVM2, Mino-1 and Maver-1 and the diffused large B-cell lymphoma (DLBCL) cell line Karpas 422. Next, we sought to investigate whether the DOT1L-dependence of MYC-driven B-cell lymphoma could be reproduced in a well-defined model of MYC-driven B-cell lymphoma. Towards this end, we utilized a mouse model in which expression of the Cre recombinase from a B cell specific promoter leads to ectopic expression of a transgenic human MYC allele and concomitant deletion of the tumor suppressor Pten in B cells. Similar to our in vitro studies, Pinometostat treatment led to a significant reduction in proliferation of B-cell lymphoma cells from these mice with an IC50 of 0.5 µM. Furthermore, we sought to ascertain whether these findings reflected on-target effects related to DOT1L inhibition. Therefore, we deleted DOT1L using CRISPR/Cas9 in B-cell lymphoma cell lines and assessed the effect on proliferation using competitive-proliferation assays. We observed that DOT1L-deletion progressively diminished the relative growth of anti-DOT1L sgRNA-expressing P493-6 and Jeko1 cells compared to non-targeted cells invitro. In order to test the requirement for DOT1L in lymphoma propagation in vivo, we performed intravenous injections of equal number of Jeko-1 cells with either anti-DOT1L or anti-Renilla control sgRNAs into sub-lethally irradiated non-obese diabetic/severe combined immunodeficiency mice (NOD/SCID) mice. Mice injected with control anti-Renilla sgRNAs succumbed to disease with a median latency of 34 days while the latency of disease in the anti-DOT1L sgRNA cohort was 45 days. In summary, DOT1L depletion significantly delayed disease latency in this invivo disseminated model of B-cell lymphoma (P=0.02). We then performed transcriptomic analyses of Pinometostat-treated B-cell lymphoma cell lines compared to DMSO-treated counterparts using RNA-seq. Gene-set enrichment analysis (GSEA) of RNA-seq data from three different B-cell lymphoma cell lines demonstrated that Pinometostat treatment significantly decreased the expression of MYC-target genes. In order to investigate the intriguing role of DOT1L in regulating MYC-target gene expression, we used ChIP-seq to assess the genome-wide occupancy of MYC following DOT1L inhibitor treatment. Strikingly, our studies demonstrated that DOT1L inhibition significantly reduced the chromatin occupancy of MYC. Taken together, our experiments demonstrate the role of DOT1L in MYC-driven B-cell lymphoma pathogenesis invitro and invivo. Furthermore, our genome-wide studies demonstrate the importance of DOT1L for genomic MYC occupancy. Based on these findings, we propose that therapeutic DOT1L targeting may be a viable strategy in MYC-driven B-cell lymphoma. Disclosures Weigert: Roche: Research Funding; Novartis: Research Funding. Rickert:Pfizer: Employment. Ren:Elli Lilly: Consultancy, Membership on an entity's Board of Directors or advisory committees; Arima Genomics: Equity Ownership, Membership on an entity's Board of Directors or advisory committees. Deshpande:Salgomed Therapeutics: Membership on an entity's Board of Directors or advisory committees; A2A Pharma: Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2013 ◽  
Vol 122 (21) ◽  
pp. 3070-3070
Author(s):  
Eugenio Gaudio ◽  
Ivo Kwee ◽  
Andrea Rinaldi ◽  
Michela Boi ◽  
Elena Bernasconi ◽  
...  

Abstract Epigenome deregulation in cancer cells affects transcription of oncogenes and tumor suppressor genes. BET Bromodomain proteins recognize chromatin modifications and act as epigenetic readers contributing to gene transcription. BET Bromodomain inhibitors showed promising pre-clinical activity in hematological and solid tumors and are currently in phase I studies. The mechanism of action and relevant affected genes are not fully characterized and there are no established response predictors. We have shown activity of BET Bromodomain OTX015 in lymphoma cell lines (ASH 2012; ICML 2013). This study aimed at elucidating pathways and genes affecting response/resistance to BET Bromodomain inhibitors in lymphomas. Methods Baseline gene expression profiles (GEP) were obtained in 38 cell lines [22 diffuse large B-cell lymphoma (DLBCL), 8 anaplastic large T-cell lymphoma, 4 mantle cell lymphoma, 3 splenic marginal zone lymphoma, 1 chronic lymphocytic leukemia] with Illumina HumanHT-12 v4 Expression BeadChip. Genetic and biologic information were collected from literature. GEP/IC50 correlation (ASH 2012; ICML 2013) was assessed by Pearson correlation. Associations in two-way tables were tested for statistical significance using either chi-square or Fisher exact test, as appropriate. Differential expression analysis was performed using LIMMA, followed by multiple test correction using the BH method. Enrichment of functionally-related genes was evaluated by GSEA. For combination studies, 3 germinal center B-cell (GCB) and 2 activated B-cell (ABC) DLBCL were exposed to increasing doses of OTX015 alone or in combination with increasing doses of targeted agents for 72 hours, followed by MTT assay. Synergy was assessed by Chou-Talalay combination index (CI) with Synergy R package. Results Transcripts associated with resistance to OTX015 were significantly enriched of genes involved in cell cycle regulation, DNA repair, chromatin structure, early B-cell development, E2F/E2F2 target genes, IL6-dependent genes, and mRNA processing. Conversely, transcripts associated with OTX015 sensitivity were enriched of hypoxia-regulated genes, interferon target genes, STAT3 targets, and involved in glucose metabolism. Genes associated with OTX015 sensitivity included LDHA, PGK1 (glucose metabolism) and VEGFA (hypoxia), while BCL2L1/BCLXL, BIRC5/survivin (anti-apoptosis), ERCC1 (DNA repair), TAF1A and BRD7 (transcription regulation) were correlated with reduced sensitivity. GEP identified 50 transcripts differentially expressed, including IL6, HCK, SGK1, MARCH1 and TRAFD1, between cells undergoing or not apoptosis after OTX015 exposure. GSEA showed significant enrichment of genes involved in IL-10 signaling pathway. While there was no association between response to OTX015<500nM and presence of translocated MYC, analysis of genetic and biologic features identified the ABC phenotype (P=.008) and presence of concomitant somatic mutations in MYD88 and CD79B or CARD11 genes and wild type TP53 (P=.027) as associated with apoptosis. Based on these observations and since mutated MYD88 interacts with BTK and MYD88/CD79B mutations have been associated with clinical responses with the BTK inhibitor ibrutinib, we evaluated OTX015 combination with this compound. Synergy was observed in particular in ABC-DLBCL with a median CI of .04 (range .02-.1). The demonstrated down-regulation of the MYD88/JAK/STAT pathway after OTX015 treatment, as shown by additional GEP, highlighted the importance of this pathway for OTX015 activity. Other targeted agents (everolimus, lenalidomide, rituximab, decitabine, vorinostat) appeared to synergize with OTX015 (Fig 1). The mTOR inhibitor everolimus presented a very strong synergism with a median CI of .11 (.1-.2), in accordance with the association between OTX015 sensitivity to high glucose metabolism and high levels of SGK1 in cells undergoing apoptosis. Conclusions Our study identified genetic mechanisms contributing to the response to BET Bromodomain inhibitors and promising combination schemes, such as OTX015/everolimus, to be further investigated. Disclosures: Stathis: Oncoethix: NCT01713582 PI Other. Herait:Oncoethix: Membership on an entity’s Board of Directors or advisory committees. Noel:Oncoethix: Membership on an entity’s Board of Directors or advisory committees. Inghirami:Oncoethix: Research Funding. Bertoni:Oncoethix: Research Funding.


Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 705-705 ◽  
Author(s):  
Valeria Spina ◽  
Hossein Khiabanian ◽  
Alessio Bruscaggin ◽  
Monica Messina ◽  
Sara Monti ◽  
...  

Abstract Background. Nodal marginal zone lymphoma (NMZL) is one of the few B-cell tumors still remaining orphan of cancer gene lesions. By combining whole exome sequencing (WES), deep sequencing of tumor-related genes, high resolution SNP array and RNAseq, here we aim at characterizing the coding genome of NMZL and at disclosing the pathways that are molecularly deregulated in this lymphoma. Methods. The study was based on 35 NMZL (tumor representation >70%) with a diagnosis confirmed by: i) pathological revision of lymph node histology; and ii) lack of clinico-radiological evidence of extranodal or splenic disease either at diagnosis or during follow-up. Consistent with NMZL, the cases investigated: i) lacked CD5, CD10 and cyclin D1 expression, 7q deletion, t(11;14), t(14;18), t(11;18) and t(1;14) translocations; and ii) recurrently harbored +3 (14%), +12 (14%) and preferential usage of the IGHV4-34 gene (17%). WES (HiSeq 2500, Illumina; mean coverage per sample: 38x-114x) and high density SNP array (Cytoscan HD, Affymetrix) of tumor/normal DNA pairs from 18 discovery NMZL identified 557 non-synonymous somatic mutations (average: 30.9/case) affecting 504 genes and 61 copy number abnormalities (CNA) (average 3.4/case). To further characterize mutation recurrence, the 504 discovered genes were investigated in an independent validation panel of 17 NMZL by targeted sequencing of tumor/normal DNA pairs (MiSeq; target region: 1.6 Mb; mean coverage per sample: 171x-386x). The 17 validation NMZL were also assessed for CNA by high density SNP arrays. RNAseq of 11 discovery NMZL did not identify any recurrent gene fusion. Results. By compiling the results of WES and high resolution SNP array, 39 genes were recurrently affected in >3/35 (9%) NMZL by mutations (n=30 genes) or focal CNA (n=9 genes). Among these, MLL2 (34%), PTPRD (20%) and NOTCH2 (20%) were most frequently mutated. Overall, recurrently mutated genes pointed to the molecular deregulation of specific programs in NMZL, namely JAK/STAT, NOTCH, NF-κB and toll-like receptor (TLR) signaling, cell cycle, chromatin remodeling/transcriptional regulation and immune escape (Fig. 1A). JAK/STAT signaling was targeted by mutually exclusive lesions in 43% of NMZL, and the protein tyrosine phosphatase receptor delta (PTPRD) tumor suppressor was the most frequently affected gene of this system in 20% of NMZL (Fig. 1B-E). PTPRD inhibits JAK/STAT signaling through the dephosphorylation of active p-STAT3. PTPRD lesions in NMZL were represented by somatic mutations that truncated or modified the tyrosine phosphatase domain, as well as deletions of the entire gene locus, including focal and biallelic losses (Fig. 1B-C). Interrogation of institutional and public genomic datasets revealed that PTPRD mutations are specific for NMZL, being rare or absent in other mature B-cell tumors, including splenic marginal zone lymphoma (Fig. 1D). Other JAK/STAT signaling genes affected in NMZL were JAK2, CXCR4 (6%), PTPN2, JAK3, STAT2, SH2B3 and CUL3 (3%) (Fig. 1E). NF-kB signaling was altered in 54% of NMZL by lesions of TNFAIP3 (14%), BCL10, REL (11%), CARD11 (9%), TRAF3 and BIRC3 (6%). NOTCH signaling was targeted in 40% of NMZL by mutations that alternatively involved NOTCH2 (20%), SPEN (11%), RBPJL (6%), FBXW7, DTX1, ITCH and MAML2 (3%). TLR signaling was targeted in 17% of NMZL, including mutations of MYD88 (9%), IRAK1BP1, PELI2 and SEMP6 (3%). Several cell cycle genes were molecularly deregulated in 43% of NMZL, including CDKN2A, PARK2, PARKG (9%), CDC16, CDCA2 (6%), CCNA1, CCNT2, CDK5, CDK13, CDK20, BTG2, HECA and PLK2 (3%). Most (71%) NMZL harbored genetic lesions affecting epigenetic modifiers (MLL2: 34%; CREBBP: 9%; EP300: 6%; TRRAP: 6%), histones (20%) or transcriptional co-repressors (TBL1XR1: 14%; ARID1A: 14%; RCOR1: 11%; NCOR2: 9%, ARID1B: 9%). Finally, the TNFRSF14 and FAS genes, involved in T cell-mediated tumor surveillance, were disrupted by mutations and/or deletions in 17% and 14% NMZL, respectively. Conclusions. A number of actionable cellular programs are molecularly deregulated in NMZL, including JAK/STAT, NOTCH, NF-κB and TLR signaling, cell cycle and chromatin remodeling. PTPRD lesions are among the most recurrent alterations in NMZL and appear to be specific for this lymphoma type across mature B-cell tumors. Figure 1 Figure 1. Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 3547-3547
Author(s):  
Jiao Ma ◽  
Kui Nie ◽  
David Redmond ◽  
Yifang Liu ◽  
Daniel M Knowles ◽  
...  

Abstract PRDM1/Blimp1, a master regulator of B-cell terminal differentiation, has been identified as a tumor suppressor gene in the pathogenesis of diffuse large B-cell lymphoma (DLBCL). In DLBCL, PRDM1 is inactivated by mutations and deletions; however, there is also evidence that PRDM1 is down-regulated by microRNAs (miRNAs) in DLBCL and Hodgkin/Reed-Sternberg cells of classical Hodgkin lymphoma (cHL). A decrease in PRDM1 activity contributes to the pathogenesis of DLBCL and cHL by inhibiting plasma cell differentiation triggered by signal transduction pathways such as the NF-kB pathway. Since malignant EBV-positive B-cell lymphoproliferations are often associated with increased NF-kB activity, it is conceivable that abnormal PRDM1 down-regulation may play a role in their pathogenesis. EBV-positive B-cell lymphomas are postulated to originate from EBV-infected B-cells with latency III growth program of EBV gene expression. Thus, EBV-immmortalized lymphoblastoid cell lines (LCLs), which are of latency III type, serve as a good model to study EBV lymphomagenesis. We observed discordance in PRDM1 mRNA and protein levels in LCLs. By quantitative real-time reverse transcriptase PCR, PRDM1 mRNA levels in LCLs varied from 14.6% to 1259.7% relative to the multiple myeloma cell line U266, which expresses high levels of PRDM1. However, PRDM1 protein was discordantly low in LCLs compared to U266 based on immunohistochemistry and Western blotting assays, consistent with post-transcriptional regulation. EBV encodes 25 viral miRNAs, and we postulate that one of more of them may function to dampen PRDM1 expression. Indeed, a miRNA binding site containing seed match to bases 2-7 of EBV miR-BHRF1-2 was identified in positions 1565 to 1589 of PRDM1 3’ untranslated region. MiR-BHRF1-2 functionally targeted this specific binding site and repressed luciferase reporter activity. Mutation in the seed region of this site relieved the repression in comparison to the wild type control. MiR-BHRF1-2 was highly expressed in LCLs, while it was barely detectable in the EBV-positive Burkitt lymphoma cell line MUTU I, which has latency type I. Importantly, immunoblotting assay demonstrated an up-regulation of PRDM1 protein level in CCL156 and CCL159 LCL cells transfected with miR-BHRF1-2 inhibitor relative to those transfected with miRNA Inhibitor negative control, supporting a role of miR-BHRF1-2 in PRDM1 down-regulation in vivo. To examine the biological consequences of increased PRDM1 expression in LCL cells, PRDM1 was over-expressed in JY25 and CCL159 LCL cell lines. Enforced expression of PRDM1 induced apoptosis in both cell lines. Furthermore, bromodeoxyuridine (Brdu) incorporation study demonstrated that overexpression of PRDM1 reduced the percentage of S phase from 43.4% to 27.6% in CCL159 cells, and 39.5% to 27.9% in JY25 cells, respectively. Whole transcriptome sequencing (RNA-seq) identified a set of potential PRDM1 direct target genes whose expressions decreased in both LCL cell lines upon PRDM1 over-expression. These genes have broad functions including cell proliferation and survival, transcription and translation, mitochondrial functions, and cytoskeleton. Although no significant changes in cell cycle and apoptosis were observed upon transfection of miR-BHRF1-2 inhibitor, RNA-seq analysis of CLL159 cells transfected with miR-BHRF1-2 inhibitor revealed a small subset of repressed genes which overlapped with those identified by PRDM1 over-expression. This finding suggests that the increase in PRDM1 expression upon miR-BHRF1-2 inhibition, albeit small, is capable of repressing a subset of PRDM1 target genes with potential biological effects. In summary, our findings demonstrate that PRDM1 is a target of EBV miR-BHRF1-2. MiR-BHRF1-2 mediated PRDM1 down-regulation may contribute to the pathogenesis of EBV-associated B-cell lymphomas by inhibiting the transcription repression program of PRDM1 and limiting PRDM1-mediated cellular changes detrimental to tumor growth, including cell cycle arrest and apoptosis. Further characterization of the target genes whose expression is up-regulated by miR-BHRF1-2-mediated PRDM1 down-regulation may provide important clues to the pathogenetic function of miR-BHRF1-2 and EBV oncogenesis in general. Disclosures No relevant conflicts of interest to declare.


Sign in / Sign up

Export Citation Format

Share Document