scholarly journals Single-Cell Multi-Omics Reveals the Genetic, Cellular and Molecular Landscape of TP53 Mutated Leukemic Transformation in MPN

Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 3-3
Author(s):  
Alba Rodriguez-Meira ◽  
Haseeb Rahman ◽  
Ruggiero Norfo ◽  
Wei Wen ◽  
Agathe Chédeville ◽  
...  

Abstract In myeloid malignancies, presence of 'multi-hit' TP53 mutations is associated with lack of response to conventional therapy and dismal outcomes, particularly when found in combination with a complex karyotype. Therefore, it is crucial to understand the biological basis of TP53-mutant driven clonal evolution, suppression of antecedent clones and eventual disease transformation to inform the development of more effective therapies. Myeloproliferative neoplasms (MPN) represent an ideal tractable disease model to study this process, as progression to secondary acute myeloid leukemia (sAML) frequently occurs through the acquisition of TP53 missense mutations. To characterize tumor phylogenies, cellular hierarchies and molecular features of TP53-driven transformation, we performed single-cell multi-omic TARGET-seq analysis (PMID: 33377019 & 30765193) of 22116 hematopoietic stem and progenitor cells (HSPCs) from 35 donors and 40 timepoints (controls, MPN in chronic phase, pre-AML and TP53-mutated sAML; Figure1a). TARGET-seq uniquely enables single-cell mutation analysis with allelic resolution with parallel transcriptomic and cell-surface proteomic readouts. We invariably identified convergent clonal evolution leading to complete loss of TP53 wild-type alleles upon transformation, including parallel evolution of separate TP53 "multi-hit" subclones in the same patient (n=4/14) and JAK2-negative progression (n=2/14). Complex clonal evolution driven by chromosomal abnormalities (CAs) was present in all patients and TP53 multi-hit HSPCs without CAs were rarely observed. Subclones with recurrent CA such as monosomy 7 showed upregulation of RAS-associated transcription and preferentially expanded in xenograft models. Together, these data indicate that TP53 missense mutation, loss of TP53 wild-type allele and cytogenetic evolution are collectively required for leukemic stem cell (LSC) expansion. Integrated transcriptomic analysis of sAML samples (Figure1b) revealed three major populations: (1) a TP53-mutant cluster (Figure1c) characterized by an erythroid signature (e.g. KLF1, GATA1, GYPA; an unexpected finding as no cases showed diagnostic features of erythroid leukemia), (2) an LSC TP53-mutant cluster (Figure1d) and (3) a TP53-WT preleukemic cluster (Figure1e). The LSC cluster showed dysregulation of key stem cell regulators, from which we derived a novel 48-gene LSC score with prognostic impact in an independent AML cohort (HR=3.13; Figure1f). Importantly, this score was predictive of outcome irrespective of TP53 status for both de novo and sAML, demonstrating its broader potential clinical utility. TARGET-seq analysis uniquely allowed us to characterize rare TP53-WT preleukemic cells (preLSCs), which were almost exclusively confined to the immunophenotypic lineage-CD34+CD38-CD90+CD45RA- HSC compartment. PreLSC from sAML samples presented increased stemness, increased quiescence, aberrant inflammatory signaling and differentiation defects (Figure1g) as compared to HSCs from normal or MPN donors, both at the transcriptional and functional levels through in vitro long-term and short-term cultures. This indicates cell-extrinsic suppression of residual TP53-WT hematopoiesis. Longitudinal analysis of TP53-heterozygous mutant HSPCs at different stages of disease evolution (Figure1a) revealed that aberrant inflammatory signalling (e.g. BST2, IFITM1, IFITM3) in the genetic ancestors of TP53 "multi-hit" LSCs, but not the presence of TP53-mutations alone, was predictive of subsequent transformation. In a mouse model system, TP53-mutant cells challenged with sustained inflammatory stimuli acquired a mean 3-fold competitive advantage in WT: TP53 R172H/+chimeras. This indicates that pro-inflammatory cues from the tumour microenvironment promote fitness advantage of TP53-mutant cells whilst supressing antecedent clones. In summary, we present a comprehensive single-cell multi-omic analysis of the genetic, cellular and molecular landscape of TP53-mediated transformation, providing unique insights into the evolution of chronic hematological malignancies towards an aggressive acute leukemia (Figure1h). Since TP53 is the most commonly mutated gene in human cancer, we anticipate these findings will be of broader relevance to many other cancer types. Figure 1 Figure 1. Disclosures Kretzschmar: Vanadis Diagnostics, a PerkinElmer company.: Current Employment. Drummond: BMS: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; CTI: Membership on an entity's Board of Directors or advisory committees; Novartis: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau. Harrison: Geron: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; BMS: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Galacteo: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Keros: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Sierra Oncology: Honoraria; Constellation Pharmaceuticals: Research Funding; Abbvie: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; AOP Orphan Pharmaceuticals: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Incyte Corporation: Speakers Bureau; Promedior: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Janssen: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Roche: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Shire: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Gilead Sciences: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; CTI BioPharma: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Celgene: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Novartis: Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau. Mead: Abbvie: Consultancy, Honoraria; Celgene/BMS: Consultancy, Honoraria, Research Funding; Novartis: Consultancy, Honoraria, Speakers Bureau.

Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 27-28
Author(s):  
Yuki Nishida ◽  
Rafael Heinz Montoya ◽  
Kiyomi Morita ◽  
Tomoyuki Tanaka ◽  
Feng Wang ◽  
...  

MDM2 inhibition by small molecules as a means of restoring p53 function has shown clinical activity against acute myeloid leukemia (AML) (Andreeff, Clin Cancer Res 2015). However, we and others have found increased variant allele frequencies (VAFs) of TP53 mutations in AML cells after treatment with MDM2 inhibitors, either as monotherapy or in combination with other agents (Daver ASH 2019), which suggests that MDM2 inhibition selects preexisting clones or generates de novo clones with TP53 mutations. We performed a long-term culture of AML cells (MOLM-13, an AML cell line with wild-type p53 and FLT3-ITD) treated with increasing concentrations of an MDM2 inhibitor idasanutlin (up to 320 nM, less than 10% concentration of Cmax) (Selleck). We obtained MDM2 inhibitor-resistant (R) AML cells after 96 days of the drug exposure and found that the resistant cells harbor hotspot TP53 p.R248W (R248W) mutation. We next isolated single cell clones from MOLM-13 R cells by limiting dilution, and obtained twelve subclones (subclones #1-12 in order of development). All clones carried the same R248W mutation. To determine clonal patterns of these cells, we performed single cell DNA sequencing (scDNAseq) of MOLM-13 parental, R and subclone #1 and #2 (SC1 and SC2) cells using the MissionBio Tapestry system covering 125 amplicons of 19 genes frequently mutated in AML. scDNAseq identified FLT3-ITD mutations in all cells analyzed, as expected. In the parental cells we identified only 0.02% cells (1/ 5,240) with the identical R248W mutation found in MOLM-13 R cells. MOLM-13 R cells had only 0.6% of wild-type TP53 cells, 51% carrying R248W only, and 43% R248W/R213* mutation (R248W/R213*). SC1 and SC2 cells had 1% and 99% of R248W and R248W/R213* clones, respectively (Fig.1). Seven other mutations were detected by scDNAseq. Results suggest that MDM2 inhibition can accelerate the selection of TP53-mutant AML cells in vitro. Of note, the parental cells had remained mostly p53 wild-type, where the subclone with mutant R248W did not have a growth advantage over other cells. Next we analyzed patient samples enrolled in the phase 1 clinical trial (NCT02319369) for the MDM2 inhibitor milademetan (DS-3032b; Daiichi-Sankyo) in relapsed/refractory AML or high-risk MDS patients. Fifty seven patients were treated with single agent milademetan in the study. All but one patient had wild-type TP53 as determined by NGS at baseline. One patient (1.8%) had a TP53 p.R213* mutation at baseline with VAF of 91%. Four patients (7%) developed different TP53 mutations (R248Q, R248W, P250fs, V122fs and V274L), with increasing VAFs that were not detected at baseline to 19% average, ranging from 11% to 28% post treatment, One patient (anonymized ID 1001-1005) developed both, R248Q and R248W mutations, detected at cycle 2 day1 (C2D1, day 29). The pre-existing R213* mutation detected in one patient persisted with increased VAF after treatment (91% to 100%). To detect p53 mutations with higher sensitivity than NGS, we performed droplet digital PCR (ddPCR) for R248W/Q and R273H in samples from two patients. ddPCR detected 0.46% and 0.62% of R248Q and R248W mutations, respectively, at baseline, which were not detected by NGS, with increased VAFs of 18.2 and 27.6% at C2D1, respectively. ddPCR detected additional R273H mutations with VAFs of 0.08% and 0.18% at baseline, and 2.2% and 2.6% in these patients at C2D1, respectively. Conclusion: Data suggest that MDM2 inhibition selects rare AML subpopulations with TP53 mutations and careful monitoring of patients treated with MDM2 inhibitors with sensitive methods to detect TP53 mutant clones is warranted. This finding also points to the need to develop strategies to prevent/suppress these clones early on. Disclosures Kumar: Daiichi-Sankyo Inc.: Current Employment. Patel:Daiichi-Sankyo Inc.: Current Employment. Dos Santos:Daiichi Sankyo, Inc.: Current Employment. DiNardo:Daiichi Sankyo: Consultancy, Honoraria, Research Funding; Novartis: Consultancy; ImmuneOnc: Honoraria; Calithera: Research Funding; Celgene: Consultancy, Honoraria, Research Funding; Notable Labs: Membership on an entity's Board of Directors or advisory committees; AbbVie: Consultancy, Honoraria, Research Funding; MedImmune: Honoraria; Jazz: Honoraria; Agios: Consultancy, Honoraria, Research Funding; Takeda: Honoraria; Syros: Honoraria. Andreeff:Daiichi-Sankyo; Jazz Pharmaceuticals; Celgene; Amgen; AstraZeneca; 6 Dimensions Capital: Consultancy; Daiichi-Sankyo; Breast Cancer Research Foundation; CPRIT; NIH/NCI; Amgen; AstraZeneca: Research Funding; Centre for Drug Research & Development; Cancer UK; NCI-CTEP; German Research Council; Leukemia Lymphoma Foundation (LLS); NCI-RDCRN (Rare Disease Clin Network); CLL Founcdation; BioLineRx; SentiBio; Aptose Biosciences, Inc: Membership on an entity's Board of Directors or advisory committees; Amgen: Research Funding.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 516-516
Author(s):  
Pavan Bachireddy ◽  
Christina Ennis ◽  
Vinhkhang N Nguyen ◽  
Nikolas Barkas ◽  
Sachet Shukla ◽  
...  

The factors mediating GvL resistance following allogeneic stem cell transplant (SCT) in lymphoid malignancies remain incompletely characterized. Because cell-intrinsic features shape chemotherapeutic relapse, we hypothesized that they also shape GvL outcomes by influencing evolutionary trajectories of CLL relapse after reduced intensity conditioning SCT (RIC). We identified 9 heavily pre-treated patients (pts) (range: 1-5 therapies, median: 3) with various times to CLL relapse after RIC (range: 83-1825 days), of which 8 had at least partial responses before relapse. To define evolutionary trajectories, we generated paired whole-exome and RNA sequencing data from purified CLL cells pre/post-RIC, using MuTect2 and ABSOLUTE algorithms to identify somatic alterations (SAs) and corresponding cancer cell fractions (CCFs). 5 pts had clonal SAs in TP53 and/or SF3B1 pre-SCT, and no single SA was specific to post-RIC. Furthermore, we found no SAs nor altered expression of HLA class I/II or b2M in either baseline or post-RIC samples. However, we found 6 relapse pairs to exhibit complex branched evolution involving CCF shifts of at least 0.2 in subclonal and clonal SAs whereas 3 pairs showed genomic stability. Clonal evolution was associated with longer time to relapse (Wilcoxon, p=0.02; median 798 versus 304 days) as well as complete response (p=0.05), suggesting that GvL immune escape may be facilitated by clonal evolution. To determine the phenotypic consequences of clonal evolution, we examined single cell transcriptomes using scRNAseq from paired pre/post-RIC CLL cells from 2 pts with early (304, 442 days; "ERs") and 2 pts with late (1801, 1825 days; "LRs") relapses after RIC. Using the inDrop platform, we profiled a median of 3560 CLL cells/pt (range: 2254-5278). Clustering using Seurat revealed marked transcriptional stability after RIC in ERs whereas dramatic shifts in gene expression programs were observed in LRs. Single cell trajectory analysis using Monocle identified ordered biological processes through which LRs, but not ERs, progressed. Branched expression analysis revealed multiple patient specific pathways defining LRs, including within chromatin regulators (EBF1, BANK1), oncogenic pathways (AFF3, DENND4A) and ribosomal biosynthesis (EEF1G, NACA). Thus, genetic evolution in LRs results in distinct phenotypic consequences. To directly link SAs with transcriptional outcomes, we interrogated scRNAseq data for known SAs identified by WES. In one LR, loss of a CLL cancer driver (RPS15mut) was observed in two of three post-RIC transcriptional clusters, either through deletion of chr.19p (where RPS15 resides) or reversion to the wildtype allele (implying loss of heterozygosity). In addition, genomic and transcriptional loss of HLA genes were detectable in pre-RIC clusters that failed to expand at relapse in both LRs, suggesting that pre-existing HLA loss does not provide a selective advantage for CLL relapse after RIC, consistent with our bulk analyses. These data highlight how scRNAseq can delineate genetic selection pressures within subpopulations of a single patient. To investigate whether epigenetic dysregulation underlies these genetic changes, we measured locally disordered methylation (LDM), a known epigenetic mechanism of CLL genetic variability. Genome-wide methylome profiles revealed increases in LDM in LRs compared to ERs for various genomic regions (Kruskal-Wallis (KW), p<0.05 for promoters, genes, distal regulatory modules); no increases in LDM were observed in an independent cohort of late CLL relapse after chemotherapy alone (n=7; time between samples: 496-1511 days). Moreover, we controlled for time between samples by calculating the rate of change in LDM and still found significant differences only during LR after RIC (versus ER or late relapse after chemotherapy; KW, p<10-13). Finally, genes with increased LDM were enriched for multiple stem cell gene sets (q<0.01), implicating a common stem-like state in LRs. Altogether, these data highlight important features of GvL resistance in CLL: 1) GvL selective pressure, shown by LRs, can shape evolutionary trajectories through genotypic alterations that directly exert phenotypic consequences; 2) alterations in HLA genes have less influence in CLL than in myeloid malignancies; and 3) GvL immune editing may select for epigenetic variability that facilitates evasion through stem-like states. Disclosures Brown: Octapharma: Consultancy; Novartis: Consultancy; Loxo: Consultancy, Research Funding; Kite: Consultancy, Research Funding; Janssen: Honoraria; Invectys: Other: other; Gilead: Consultancy, Research Funding; Genentech/Roche: Consultancy; Dynamo Therapeutics: Consultancy; Catapult Therapeutics: Consultancy; BeiGene: Consultancy; AstraZeneca: Consultancy; Acerta Pharma: Consultancy; Morphosys: Other: Data safety monitoring boards ; Sun Pharmaceuticals, Inc: Research Funding; Sun: Research Funding; Verastem: Consultancy, Research Funding; TG Therapeutics: Consultancy; Teva: Honoraria; Sunesis: Consultancy; Pharmacyclics: Consultancy; Pfizer: Consultancy. Getz:MuTect, ABSOLTUE, MutSig and POLYSOLVER: Patents & Royalties: MuTect, ABSOLTUE, MutSig and POLYSOLVER; IBM: Research Funding; Pharmacyclics: Research Funding. Ho:Jazz Pharmaceuticals: Consultancy; Jazz Pharmaceuticals: Research Funding; Omeros Corporation: Membership on an entity's Board of Directors or advisory committees. Neuberg:Celgene: Research Funding; Pharmacyclics: Research Funding; Madrigal Pharmaceuticals: Equity Ownership. Soiffer:Gilead, Mana therapeutic, Cugene, Jazz: Consultancy; Jazz: Consultancy; Kiadis: Other: supervisory board; Mana therapeutic: Consultancy; Cugene: Consultancy; Juno, kiadis: Membership on an entity's Board of Directors or advisory committees, Other: DSMB. Ritz:TScan Therapeutics: Consultancy; Equillium: Research Funding; Merck: Research Funding; Kite Pharma: Research Funding; Aleta Biotherapeutics: Consultancy; Celgene: Consultancy; Avrobio: Consultancy; LifeVault Bio: Consultancy; Draper Labs: Consultancy; Talaris Therapeutics: Consultancy. Wu:Neon Therapeutics: Other: Member, Advisory Board; Pharmacyclics: Research Funding.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 797-797
Author(s):  
Talha Badar ◽  
Mark R. Litzow ◽  
Rory M. Shallis ◽  
Jan Philipp Bewersdorf ◽  
Antoine Saliba ◽  
...  

Abstract Background: TP53 mutations occur in 10-20% of patients with AML, constitute high-risk disease as per ELN criteria, and confer poorer prognosis. Venetoclax combination therapies and CPX-351 were recently approved for AML treatment and lead to improved outcomes in subsets of high-risk AML, however the most effective approach for treatment of TP53-mutated (m) AML remains unclear. In this study we explored the clinical outcome of TP53m AML patients treated over the last 8 years as novel therapies have been introduced to our therapeutic armamentarium. Methods: We conducted a multicenter observational study in collaboration with 4 U.S. academic centers and analyzed clinical characteristics and outcome of 174 TP53m AML patients diagnosed between March 2013 and February 2021. Mutation analysis was performed on bone marrow specimens using 42, 49, 199, or 400 gene targeted next generation sequencing (NGS) panels. Patients with an initial diagnosis of AML were divided into 4 groups (GP) based on the progressive use of novel therapies in clinical trials and their approvals as AML induction therapy during different time periods: 2013-2017 (GP1, n= 37), 2018-2019 (GP2, n= 53), 2019-2020 (GP3, n= 48) and 2020-2021 (GP4, n= 36) to analyze difference in outcome. Results: Baseline characteristics were not significantly different across different GP, as shown in Table 1. Median age of patients was 68 (range [R], 18-83), 65 (R, 29-88), 69 (R, 37-90) and 70 (R, 51-97) years in GP1-4, respectively (p=0.40). The percentage of patients with de novo AML/secondary AML/therapy-related AML in GP1-4 was 40/40/20, 36/29/24, 37.5/37.5/25 and 28/52/20, respectively (p=0.82). The proportion of patients with complex cytogenetics (CG) was 92%, 89%, 96% and 94% in GP1-4, respectively (p=0.54). The median TP53m variant allele frequency (VAF) was 48% (range [R], 5-94), 42% (R, 5-91), 45% (R, 10-94) and 60% (R, 8-82) in GP1-4, respectively (p=0.38). Four (11%), 13 (24.5%), 10 (21%) and 9 (25%) patients had multiple TP53 mutations in GP1-4, respectively (p=0.33). The proportion of patients who received 3+7 (30%, 16%, 6% & 8%; p=0.01), HMA only (11%, 18%, 2% & 8%; p=0.06), venetoclax-based (2.5%, 12%, 48%, & 61%; p <0.01) and CPX-351 induction (16%, 40%, 28% & 5%; p<0.001) were varied in GP1-4, respectively. The rate of CR/CRi was 22%, 26%, 28% and 18% in GP1-4, respectively (p=0.63). Treatment related mortality during induction was observed in 3%, 7%, 10% and 17% of patients in GP1-4, respectively (p=0.18). Overall, 28 (16%) patients received allogeneic hematopoietic stem cell transplantation (alloHCT) after induction/consolidation: 22%, 15%, 17% and 11% in GP1-4, respectively (p=0.67). In subset analysis, there was no difference in the rate of CR/CRi with venetoclax-based regimens vs. others (39% vs 61%, p=0.18) or with CPX-351 vs. others (25% vs 75%, p=0.84). The median progression-free survival was 7.7, 7.0, 5.1 and 6.6 months in GP1-4, respectively (p=0.60, Fig 1A). The median overall survival (OS) was 9.4, 6.1, 4.0 and 8.0 months in GP1-4, respectively (p=0.29, Fig 1B). In univariate analysis for OS, achievement of CR/CRi (p<0.001) and alloHCT in CR1 (p<0.001) associated with favorable outcome, whereas complex CG (p=0.01) and primary refractory disease (p<0.001) associated with poor outcome. Multiple TP53 mutations (p=0.73), concurrent ASXL1m (p=0.86), extra-medullary disease (p=0.92), ≥ 3 non-TP53m mutations (p=0.72), TP53m VAF ≥ 40% vs. < 40% (p=0.25), induction with CPX-351 vs. others (p=0.59) or venetoclax-based regimen vs. others (p=0.14) did not show significance for favorable or poor OS in univariate analysis. In multivariable analysis, alloHCT in CR1 (hazard ratio [HR]=0.28, 95% CI: 0.15-0.53; p=0.001) retained an association with favorable OS and complex CG (HR 4.23, 95%CI: 1.79-10.0; p=0.001) retained an association with dismal OS. Conclusion: We present the largest experience with TP53m AML patients analyzed by NGS. Although outcomes were almost universally dismal, alloHCT appears to improve the long-term survival in a subset of these patients. Effective therapies are warranted to successfully bridge patients to alloHCT and to prolong survival for transplant ineligible patients. Figure 1 Figure 1. Disclosures Badar: Pfizer Hematology-Oncology: Membership on an entity's Board of Directors or advisory committees. Litzow: Omeros: Other: Advisory Board; Pluristem: Research Funding; Actinium: Research Funding; Amgen: Research Funding; Jazz: Other: Advisory Board; AbbVie: Research Funding; Astellas: Research Funding; Biosight: Other: Data monitoring committee. Shallis: Curis: Divested equity in a private or publicly-traded company in the past 24 months. Goldberg: Celularity: Research Funding; Astellas: Consultancy, Membership on an entity's Board of Directors or advisory committees; Aprea: Research Funding; Arog: Research Funding; DAVA Oncology: Honoraria; Genentech: Consultancy, Membership on an entity's Board of Directors or advisory committees; Pfizer: Research Funding; Prelude Therapeutics: Research Funding; Aptose: Consultancy, Research Funding; AbbVie: Consultancy, Membership on an entity's Board of Directors or advisory committees, Research Funding. Atallah: BMS: Honoraria, Speakers Bureau; Takeda: Consultancy, Research Funding; Amgen: Consultancy; Abbvie: Consultancy, Speakers Bureau; Novartis: Consultancy, Honoraria, Research Funding; Pfizer: Consultancy, Research Funding. Foran: revolution medicine: Honoraria; gamida: Honoraria; bms: Honoraria; pfizer: Honoraria; novartis: Honoraria; takeda: Research Funding; kura: Research Funding; h3bioscience: Research Funding; OncLive: Honoraria; servier: Honoraria; aptose: Research Funding; actinium: Research Funding; abbvie: Research Funding; trillium: Research Funding; sanofi aventis: Honoraria; certara: Honoraria; syros: Honoraria; taiho: Honoraria; boehringer ingelheim: Research Funding; aprea: Research Funding; sellas: Research Funding; stemline: Research Funding.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 2920-2920
Author(s):  
Razan Mohty ◽  
Abdul Hamid Bazarbachi ◽  
Myriam Labopin ◽  
Jordi Esteve ◽  
Nicolaus Kröger ◽  
...  

Abstract Isocitrate dehydrogenase (IDH) 1 and 2 mutations occur in 20% of acute myeloid leukemia (AML). Patients with AML carrying IDH1-2 mutations have a similar prognosis compared to patients without these mutations (DiNardo et al, AM J H, 2015). However, the impact of IDH1-2 mutations on patients with AML undergoing allogeneic hematopoietic stem cell transplantation (allo-HCT) is not well known. In this study, we investigate the prognostic impact of IDH1-2 mutational status on AML patients undergoing alloHCT in complete remission (CR). In this retrospective registry-based analysis, we identified 710 consecutive adult patients (46.2% female; median age: 58.5 years [range, 18-78]) with AML undergoing allo-HCT in CR between 2015 and 2019 at 85 EBMT participating centers. Cord blood, ex-vivo graft manipulated transplants, and patients with favorable cytogenetics were excluded. Median follow-up was 15 months [95% CI 13.4-16.6]. Patients were categorized based on IDH1-2 mutational status, with 300 (42%) mutated and 410 (58%) wild type. Six hundred and fifty-two (92%) and 58 (8%) patients had de novo and secondary AML, respectively, and 141 (20%) patients had poor-risk cytogenetics. IDH1-2 mutation was positively correlated with NPM1 mutation (40% in IDH1-2 mutated vs 27% in wild type, p=0.0001) and more frequently encountered in middle-aged patients (p=0.01). No correlation was noted between IDH1-2 and FLT3 mutation or other patient characteristics. Minimal residual disease (MRD) data was available for 344 patients, 53% of which were MRD-negative at transplant in both groups. Six hundred and twenty-three (88%) and 87 (12%) patients were in first and second CR at time of transplant, respectively. Patients received grafts from a matched sibling (24%), unrelated (62%), or haploidentical (14%) donor, and myeloablative conditioning (MAC) was used in 42%. Ninety-three percent of the patients received peripheral blood as the stem cell source. At day 180, the cumulative incidence of grade II-IV acute graft-versus-host disease (aGVHD) was significantly lower in IDH1-2 mutated compared to wild-type patients (22% vs 33%, p=0.002). No differences in chronic GVHD rates were noted between the 2 groups (39% vs 40%, p=0.87). The 2-year cumulative relapse incidence (RI) was significantly lower and the GVHD-free, relapse-free survival (GRFS) was also improved in IDH1-2 mutated compared to wild-type patients (14.4% vs 27%, p=0.001 and 47% vs 39%, p=0.006, respectively). This led to an improved leukemia-free survival (LFS) in IDH1-2 patients (69% vs 59%, p=0.01), however, it did not translate into an overall survival (OS) difference. No significant difference was noted in non-relapse mortality (NRM) between the 2 groups (17% vs 14.2%, p=0.26). These findings were confirmed in multivariate analysis. In fact, IDH1-2 mutation was associated with significant improvement in RI (hazard ratio [HR]=0.4 [95%CI 0.25-0.64], p=0.0001), LFS (HR=0.7 [95%CI 0.51-0.95], p=0.022), aGVHD II-IV (HR=0.63 [95%CI 0.45-0.87], p=0.005) and GRFS (HR=0.69 [95%CI 0.54-0.89], p=0.004). Conversely, the presence of adverse cytogenetics and undergoing allo-SCT in second CR increased the RI (HR=2.29 [95%CI 1.46-3.61], p=0.0003 and HR=2.84 [95%CI 1.64-4.91], p=0.0002, respectively) and were associated with a shorter LFS (HR=1.67 [95%CI 1.18-2.36], p=0.004 and HR=1.61 [95%CI 1.06-2.44], p=0.025) while reduced intensity (RIC) conditioning had a worse impact on OS compared to MAC (HR=1.56 [95%CI 1.07-2.29], p=0.022). Additionally, in the subgroup of patients with available MRD data, MRD positivity at transplant significantly increased RI (HR=2.15 [95%CI 1.09-4.23], p=0.027) with no impact on survival. In conclusion, our data suggest that the presence of IDH1-2 mutations acts as an independent prognostic factor and is associated with improved outcome in patients with AML in CR undergoing allo-HCT. Indeed, patients with IDH1-2 mutations had significantly lower rates of RI and aGVHD, which translated into improved LFS and GRFS. Nevertheless, patients with MRD positivity at time of transplant had significantly increased RI. Further studies investigating allo-HCT outcomes in IDH1-2 mutated patients with AML in the era of IDH inhibitors (both in the pre- and post-transplant setting) would help to further define the impact of these mutations in this setting and thus optimize an individualized treatment approach. Disclosures Labopin: Jazz Pharmaceuticals: Honoraria. Esteve: Abbvie: Consultancy; Pfizer: Consultancy; Astellas: Consultancy; Novartis: Research Funding; Novartis: Consultancy, Research Funding; Bristol Myers Squibb/Celgene: Consultancy; Jazz: Consultancy. Kröger: Novartis: Research Funding; Riemser: Honoraria, Research Funding; Sanofi: Honoraria; Neovii: Honoraria, Research Funding; Jazz: Honoraria, Research Funding; Gilead/Kite: Honoraria; Celgene: Honoraria, Research Funding; AOP Pharma: Honoraria. Blaise: Jazz Pharmaceuticals: Honoraria. Socie: Alexion: Research Funding. Ganser: Jazz Pharmaceuticals: Honoraria; Novartis: Honoraria; Celgene: Honoraria. Yakoub-Agha: Jazz Pharmaceuticals: Honoraria. Bazarbachi: Novartis: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Roche: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Amgen: Honoraria, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Abbvie: Honoraria, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Hikma: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Janssen: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Pfizer: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Takeda: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Astra Zeneca: Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 15-16
Author(s):  
Ella R Thompson ◽  
Tamia Nguyen ◽  
Yamuna Kankanige ◽  
Mary Ann Anderson ◽  
Sasanka M. Handunnetti ◽  
...  

Progression of chronic lymphocytic leukemia (CLL) on venetoclax (VEN) and BTK inhibitors (BTKi) is associated with acquired genomic variants in BCL2/MCL1/BCL2L1 and BTK/PLCG2, respectively, in some patients. We aimed to assess the clonal structure and evolution of resistance in patients (pts) with progressive disease treated with single agent VEN or BTKi (or both as sequential monotherapies) using next generation sequencing (NGS) and single cell sequencing. Seven pts with CLL and 1 with mantle cell lymphoma (MCL) with disease progression on VEN, ibrutinib (IBR) or zanubrutinib (ZANU) were identified from patients treated at our institutions. Pts were selected on the basis of multiple known resistance mechanisms from previous analysis of mutations (muts) and copy number changes detected using clinical bulk NGS targeting genes of interest including BCL2, MCL1, BCL2L1, BAX, BAK1, BTK, PLCG2, CXCR4, as well as TP53 and SF3B1. Of the 8 pts selected for single cell analysis, all had disease that was relapsed/refractory to chemotherapy prior to receiving either VEN (3 pts), BTKi (2 pts) or sequential VEN-BTKi (3 pts). 6,520-16,378 individual cells from 9 samples (8 pts) were analyzed (total 103,388 cells) using a custom panel targeting pt-specific muts on the Tapestri platform (Mission Bio). A summary of genomic abnormalities detected across the cohort is presented in Figure 1. We first evaluated the relationship between genomic resistance mechanisms within the context of single agent (VEN or BTKi) as well as sequential VEN-BTKi treatment. In CLL pts treated with a single agent, all BCL2 muts in VEN pts and BTK muts in IBR or ZANU pts were identified in different subclones consistent with an oligoclonal pattern of disease progression with independent clonal acquisition of resistance mechanisms. Both pts who received ZANU (either as a single agent or sequentially) harbored the BTK L528W mut (previously described as enriched in ZANU progressors; Handunnetti ASH 2019) in independent clones from BTK C481 muts. In pts who received sequential VEN-BTKi treatment, clones were observed that harbored established or novel dual genomic resistance mechanisms within the same cell (BTK mut/MCL1 amp in CLL, BTK/BAX muts in MCL). However, this was not observed in all clones or for all pts, suggesting the presence of further undetected resistance mechanisms (genetic or other). Given the unique ability of single cell sequencing to resolve mut context within a clonal hierarchy, we next assessed this phenomenon within our cohort utilizing other muts known to be present in these tumors. Analysis of TP53 muts exemplified the diversity of clonal patterns observed, with resistance muts being detected subclonally to parental TP53 muts in some pts and independently of TP53 muts in others. In addition, further evolution of resistant clones was observed through the development of TP53 muts within clones harboring acquired resistance muts, consistent with continued clonal evolution within the resistant disease compartment. In one pt, post-resistance clonal evolution was identified through the clonal acquisition of a CXCR4 mut within a BTK mutated population. Finally, to understand the contribution of BTK zygosity and gender to BTKi resistance (given its location on the X-chromosome), we performed single cell analysis on a disease specimen from a female pt with progressive MCL harboring multiple BTK mutations following treatment with sequential VEN-BTKi. Analysis revealed four clonally independent heterozygous BTK muts inferring the sufficiency of a single mutant allele to drive resistance in this context. Interestingly, this pt also harbored a BCL2 mut and a BAX mut, the latter co-occurring with a BTK mut (BCL2 not assessable). This pt therefore represents the first description of BCL2 or BAX muts occurring in a pt with progressive MCL on VEN and the first of a BTK L528W mut in MCL progressing on ZANU. In summary, these data highlight the significant clonal complexity of CLL progression on VEN and BTKi. Our data show that disease progression in this context is consistently oligoclonal with separate clones harboring distinct identifiable resistance mechanisms. These data have pt-specific implications for the potential utility of cycling back to previously efficacious targeted therapies as well as providing a strong rationale for the early use of disease-appropriate combination targeted therapies. Disclosures Anderson: Walter and Eliza Hall Institute: Patents & Royalties: milestone and royalty payments related to venetoclax.. Handunnetti:AbbVie: Other: Travel expenses; Roche: Honoraria; Gilead: Honoraria. Yeh:Novartis: Honoraria; Gilead: Research Funding. Tam:BeiGene: Honoraria; Janssen: Honoraria, Research Funding; AbbVie: Honoraria, Research Funding. Seymour:Morphosys: Consultancy, Honoraria; Mei Pharma: Consultancy, Honoraria; Gilead: Consultancy; AstraZeneca: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees; Janssen: Consultancy, Honoraria, Research Funding; F. Hoffmann-La Roche: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Celgene: Consultancy, Honoraria, Research Funding; AbbVie: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Nurix: Honoraria. Roberts:Janssen: Research Funding; Servier: Research Funding; AbbVie: Research Funding; Genentech: Patents & Royalties: for venetoclax to one of my employers (Walter & Eliza Hall Institute); I receive a share of these royalties. Blombery:Amgen: Consultancy; Novartis: Consultancy; Invivoscribe: Honoraria; Janssen: Honoraria.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 2220-2220
Author(s):  
Robert L. Bowman ◽  
Tanmay Mishra ◽  
Shira E. Eisman ◽  
Louise Cai ◽  
Wenbin Xiao ◽  
...  

Abstract Genomic studies in acute myeloid leukemia (AML) have generated a near complete catalogue of genes mutated at varying frequencies both across patients and in individual leukemias. The high variability of mutation burden within a given leukemia is suggestive of a stepwise evolutionary process composed of early, clonal, mutations and subsequent subclonal events. The receptor tyrosine kinase, FLT3, is the most commonly mutated gene in AML, with mutations frequently manifesting as internal tandem duplications (ITDs) in the juxtamembrane domain leading to constitutive kinase activation. Although FLT3 is commonly a subclonal mutational event, FLT3 ITD mutations portend a poor prognosis particularly when combined with DNMT3A and NPM1, earlier mutations that drive clonal expansion. Notwithstanding its role as a subclonal driver, previous preclinical FLT3 models have utilized retroviral overexpression or germline mutant expression at the endogenous locus precluding accurate temporal modeling of disease. These efforts have prohibited evaluation of FLT3 mutational acquisition in the context observed in AML patients. Here, we report the development of an endogenously targeted, Flp inducible, Flt3 ITD mouse allele which can be somatically activated subsequent to cooperating disease alleles. When activated with a tamoxifen inducible FlpoER, Flt3 mutant mice developed rapid leukocytosis peaking at 4-6 weeks post activation and resolving by 8-10 weeks, a finding not previously observed in constitutive models. This leukocytosis was disproportionately monocytic and accompanied by pronounced anemia and thrombocytopenia. Long term, these mice develop a myeloproliferative disease , reminiscent of previously reported constitutive alleles. In competitive transplantation studies, Flt3 mutant cells initiated disease and outcompeted wild-type cells. Despite this competitive advantage, disease was incapable of transplanting into secondary recipients. We further observed a non-cell autonomous depletion of SLAM+ LSKs suggesting the Flt3 mutant cells cannot propagate disease in self-renewing stem cells. To evaluate how this allele influenced leukemic evolution we crossed this Flt3 ITD allele to a Flp inducible Npm1 c mouse where a pulse of tamoxifen simultaneously activated both alleles. The combination of mutant Npm1 and Flt3 resulted in progressive leukocytosis which did not resolve. Within 6 weeks of mutational activation, these mice developed a lethal AML with robust anemia, thrombocytopenia, leukocytosis and expanded cKIT+ blasts in the blood. RNA-sequencing and immunophenotyping by CyTOF revealed distinct patterns of differentiation, gene-expression and downstream signaling.In an effort to model sequential mutational acquisition, we crossed the Flp Flt3 ITD allele to a Cre-inducible Dnmt3a R878H. Cre mRNA was electroporated into lineage negative bone marrow cells to activate the Dnmt3a R878H allele and transplanted into lethally irradiated recipients. Four weeks post engraftment, Flt3 ITD was activated with a pulse of tamoxifen. In contrast to the Flt3-Npm1 model, we observed an increase and subsequent decrease in WBC similar to the kinetics observed in Flt3 ITD only mice. However, by 20 weeks we observed a robust and consistent increase in WBC accompanied by an emergence of cKIT+ cells in the blood. Histopathology indicated that >50% of mice expressing both alleles in sequence developed AML marked by increased blasts in the marrow, with moderate anemia and thrombocytopenia compared to the Flt3-Npm1 models. Critically, in contrast to Flt3 ITD only mice, acquisition of the Flt3 ITD in Npm1 or Dnmt3a mutant HPSCs induced fully transplantable AML with immunophenotypic characteristics seen in human AML with these same genotypes. Collectively these results demonstrate that different co-occurring mutations are capable of transforming Flt3 ITD mutant cells, albeit with distinct latencies and mechanisms of cooperativity. In summary, our studies utilizing novel multi-recombinase models of leukemogenesis reveal new insights into the early phase of oncogene activation, and how cooperating alleles influence this response. This inducible Flt3 ITD allele represents a significant advance in modeling clonal evolution in myeloid malignancies and provides a critical isogenic platform for preclinical development of novel leukemia therapeutic regimens. Figure 1 Figure 1. Disclosures Bowman: Mission Bio: Honoraria, Speakers Bureau. Xiao: Stemline Therapeutics: Research Funding. Miles: Mission Bio: Honoraria, Speakers Bureau. Trowbridge: Fate Therapeutics: Patents & Royalties; H3 Biomedicine: Research Funding. Levine: Amgen: Honoraria; Lilly: Honoraria; Mission Bio: Membership on an entity's Board of Directors or advisory committees; Imago: Membership on an entity's Board of Directors or advisory committees; Celgene: Research Funding; Ajax: Membership on an entity's Board of Directors or advisory committees; QIAGEN: Membership on an entity's Board of Directors or advisory committees; Gilead: Honoraria; Zentalis: Membership on an entity's Board of Directors or advisory committees; Isoplexis: Membership on an entity's Board of Directors or advisory committees; Roche: Honoraria, Research Funding; Janssen: Consultancy; Astellas: Consultancy; Morphosys: Consultancy; Incyte: Consultancy; Auron: Membership on an entity's Board of Directors or advisory committees; Prelude: Membership on an entity's Board of Directors or advisory committees; C4 Therapeutics: Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 1271-1271 ◽  
Author(s):  
Bing Z Carter ◽  
Po Yee Mak ◽  
Steven M. Kornblau ◽  
Wenjing Tao ◽  
Yuki Nishida ◽  
...  

In spite of recent progress in AML therapy, the outcomes of TP53 mutated AML remain extremely poor. The FDA-approved combination of Bcl-2 inhibition by venetoclax (VEN) with hypomethylating agents is resulting in CR/CRi rates of 70-95% and good tolerability in elderly AML patients. However, patients with TP53 mutations achieve lower response rates (CR/CRi 47%) (DiNardo CD et al., Lancet Oncol 2018; DiNardo CD et al., Blood 2019) and invariably relapse. Combined Bcl-2 inhibition and p53 activation is synthetic lethal in TP53 wild-type AML in part through targeting Mcl-1 (Pan R et al., Cancer Cell 2016). In TP53 deficient/mutant AML, direct targeting of Mcl-1 may partially compensate for the TP53 defect. We therefore postulate that combined inhibition of Mcl-1 and Bcl-2 effectively induces apoptosis in TP53 deficient/mutant AML cells. Reverse phase protein array analysis of a large cohort of newly diagnosed AML patients (n=511) enabled us to stratify patients into various prognostic groups based on p53 pathway protein expression, which included 8 core proteins: TP53, TP53pS15, MDM2, MDM4, TRIM24, SFN, IRS1.pS1101, and YWHAZ. The group with p53 pathway dysfunction, defined by high p53 protein levels, is characterized by poor outcomes (Quintas-Cardama A et al., Leukemia 2017). This group, encompassing both TP53 wild-type and TP53 mutations, had significantly lower expression of Bax (p=0.0007), the chief executioner of intrinsic apoptosis. A survey of Bcl-2 family proteins in TP53 wild-type and mutant AML showed that only Bax was significantly lower in patients with TP53 mutations (p=0.0498). We next investigated the roles of p53 in response to BH3 mimetics in TP53 wild-type and knockdown (KD) OCI-AML3 cells and in TP53 wild-type and mutant Molm13 cells, generated by long-term exposure of TP53 wild-type Molm13 cells to RG7388 (idasanutlin). Western blot analysis showed that Bax protein was consistently decreased in both TP53 KD and mutant cell lines compared to their respective controls. p53 KD OCI-AML3 and TP53 mutant Molm13 cells exhibited decreased sensitivity not only to VEN, but also to Mcl-1 inhibitor AZD5991 compared to OCI-AML3 vector control and Molm13 parental cells, respectively. To investigate if combined inhibition of Bcl-2 and Mcl-1 could counter-balance the loss of TP53 activity, p53 KD and TP53 mutated AML cells were treated with VEN and AZD5991: like in their respective control cells, the combination synergistically induced cell death in these cells (Fig. 1). The EC50 levels of VEN required for the described synergism in OCI-AML3 cells can easily be reached/exceeded clinically. The combination was also synergistic in leukemia cell lines lacking wild-type TP53 such as KG-1 and U937. Importantly, the combined inhibition was more effective than each single agent in primary AML cells and stem/progenitor cells lacking wild-type TP53 due to deletion of chromosome 17 or mutations in TP53 gene. Conclusion: AML cells deficient in functional TP53 have decreased Bax protein expression, increased apoptotic threshold and are more resistant to individual BH3 mimetics. Combined inhibition of Bcl-2 and Mcl-1 is highly synergistic in p53 deficient/mutant AML. Disclosures Carter: Amgen: Research Funding; AstraZeneca: Research Funding; Ascentage: Research Funding. Cidado:AstraZeneca: Employment. Drew:AstraZeneca: Employment. Andreeff:BiolineRx: Membership on an entity's Board of Directors or advisory committees; CLL Foundation: Membership on an entity's Board of Directors or advisory committees; NCI-RDCRN (Rare Disease Cliln Network): Membership on an entity's Board of Directors or advisory committees; Leukemia Lymphoma Society: Membership on an entity's Board of Directors or advisory committees; German Research Council: Membership on an entity's Board of Directors or advisory committees; NCI-CTEP: Membership on an entity's Board of Directors or advisory committees; Cancer UK: Membership on an entity's Board of Directors or advisory committees; Center for Drug Research & Development: Membership on an entity's Board of Directors or advisory committees; NIH/NCI: Research Funding; CPRIT: Research Funding; Breast Cancer Research Foundation: Research Funding; Oncolyze: Equity Ownership; Oncoceutics: Equity Ownership; Senti Bio: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; Eutropics: Equity Ownership; Aptose: Equity Ownership; Reata: Equity Ownership; 6 Dimensions Capital: Consultancy; AstaZeneca: Consultancy; Daiichi Sankyo, Inc.: Consultancy, Patents & Royalties: Patents licensed, royalty bearing, Research Funding; Jazz Pharmaceuticals: Consultancy; Celgene: Consultancy; Amgen: Consultancy.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 2038-2038 ◽  
Author(s):  
Francesca Romana Mauro ◽  
Alessandra Tedeschi ◽  
Alfonso Piciocchi ◽  
Marina Motta ◽  
Emilia Iannella ◽  
...  

Abstract Introduction. Observational studies from patients treated outside controlled clinical trials offer real life information and are relevant to understand whether data derived from prospective trials are reproducible in the clinical practice. A retrospective observational study was carried out by the GIMEMA (Gruppo Italiano Malattie EMatologiche dell'Adulto) group in order to evaluate the clinical characteristics and outcome of patients with chronic lymphocytic leukemia (CLL) treated with ibrutinib in Italy within a Named Patient Program (NPP). The NPP was intended to offer free and early drug access to CLL patients until ibrutinib became available on the Italian market. Methods. Patients included in the NPP program had refractory or relapsed (R/R) disease with progression within 24 months after prior chemo-immunotherapy, and/or 17p deletion/TP53 mutations. Patients were also required to have an ECOG performance status ≤2; serum creatinine ≤2 times, liver enzymes ≤3 times and total bilirubin ≤1.5 times the upper limit of normal. Key exclusion criteria were: the need of a concomitant treatment with a strong CYP3A inhibitor or warfarin, an allogeneic stem cell transplantation within the past 6 months or an ongoing active infection. All patients included in the program received ibrutinib orally as a single agent at the standard dose of 420 mg daily. Clinical data of 110 patients included in the NPP program between January 2014 and November 2014 have so far been collected and analyzed using the Research Electronic Data Capture (REDCap) system. Patients were managed at 20 Italian centers and received at least one dose of ibrutinib. Clinical data were reported by the treating physicians. Results. The median age of patients was 69.9 years (range 49.8-83.3); 53% were in Rai stage III-IV, 32% in stage II and 15% in stage 0-I. Sixty-two percent of patients had relapsed disease, 38% were refractory to prior treatment. The presence of a 17p deletion and/or TP53 mutations was recorded in 51 R/R patients. Eighty-six percent of patients had an unmutated IGHV gene profile. The median number of prior treatments was 3 and included allogeneic stem cell transplantation in 4 cases. Two or more comorbidities were reported in 57 patients (52%) and included atrial fibrillation (AF) in 10 (9.1%) and hypertension in 40 (36.4%). After a median follow-up of 12.1 months (range, 1.6-24.6), 87 patients (79%) were still on ibrutinib. A response to ibrutinib was reported in 98/110 patients (89.1%). The best recorded response was a CR/CRi in 19 patients (17.3%), while a PR was reported in 79 patients (72%; PR-L 21.1%). Similar response rates were observed in patients with unmutated IGHV genes (91.9%) and in those with 17p deletion/TP53 mutations (90.3%). At 12 months, the progression-free survival (PFS) and overall survival (OS) were 92.9% (95%CI: 87.9-98.2) and 95.2% (95%CI: 91.1-99.4), respectively. PFS at 12 months of patients who achieved a response was 96.3%, 98.9% in unmutated IGHV patients, 90.7% in those with 17p deletion/TP53 mutations. Five patients (4.5%) died during the NPP program (1 patient each for sepsis, heart failure, ileus perforation, cancer, unknown cause). Adverse events (AE) were recorded in 75 patients (68.2%); in 47 (42.7%) they were grade ≥3. Any grade AEs recorded in ≥5% of patients were: infections (35%; grade ≥3, 22%), granulocytopenia (18.8%; grade ≥3, 17.2%), bleeding (15.5%; grade ≥3, 2.7%), fever of unknown origin or febrile neutropenia (12%; grade ≥3, 5.4%), AF (10.9%; grade ≥3, 4.5%), diarrhoea (8.3; grade ≥3, 2%), hypertension (7.2%; grade ≥3, 5.4%). A new event of AF occurred in 1/10 patients with a prior history of AF. Warfarin was required in 1 patient with AF and this was the reason for ibrutinib discontinuation. Conclusions. The results of the first interim analysis of this retrospective, real life study confirms that ibrutinib, as a single agent, is an effective treatment for patients with poor-prognosis CLL. Our data also suggest that ibrutinib given to unselected patients, in a compassionate-use program, shows a clinical activity and a safety profile comparable to those reported in prospective trials. Data collection is ongoing in order to complete the analysis of this large NPP cohort in Italy. Disclosures Marasca: Roche: Honoraria; Gilead: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees; Janssen: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Pfizer: Honoraria. Coscia:Karyopharm: Research Funding; ROCHE: Honoraria, Other: Advisory board; Janssen: Honoraria; Gilead: Honoraria; Mundipharma: Honoraria. Zinzani:Abbvie: Membership on an entity's Board of Directors or advisory committees; Roche: Membership on an entity's Board of Directors or advisory committees; Janssen: Membership on an entity's Board of Directors or advisory committees; MorphoSys: Membership on an entity's Board of Directors or advisory committees; Takeda: Membership on an entity's Board of Directors or advisory committees; Celegene: Membership on an entity's Board of Directors or advisory committees. Molica:Jansen: Membership on an entity's Board of Directors or advisory committees; Abbvie: Membership on an entity's Board of Directors or advisory committees; Roche Italy: Membership on an entity's Board of Directors or advisory committees; Gilead Sciences: Speakers Bureau. Orlandi:Ariad: Honoraria; BMS: Honoraria; Novartis: Honoraria. Ghia:Gilead: Consultancy, Honoraria, Research Funding, Speakers Bureau; Janssen: Consultancy, Honoraria, Speakers Bureau; Abbvie: Consultancy, Honoraria; Adaptive Biotechnology: Consultancy; Roche: Honoraria, Research Funding. Foà:Amgen: Consultancy, Speakers Bureau; Janssen: Consultancy, Speakers Bureau; Roche: Consultancy, Speakers Bureau; Gilead: Consultancy, Speakers Bureau; Celgene: Consultancy, Speakers Bureau; BMS: Consultancy; Genentech: Consultancy; Pfizer: Speakers Bureau; Ariad: Speakers Bureau.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 1956-1956
Author(s):  
Amy Wang ◽  
Justin Kline ◽  
Wendy Stock ◽  
Satyajit Kosuri ◽  
Andrew S. Artz ◽  
...  

Background:Treatment options are limited for patients (pts) with hematologic malignancies who relapse after allogeneic stem cell transplantation (allo-SCT). We hypothesized that checkpoint inhibitors may offer a novel approach for maintaining remission after allo-SCT. Data from pre-clinical studies have suggested a potential role for PD-1/PD-L1 inhibitors in acute myeloid leukemia (AML) (Zhang et al., Blood 2009), so it is possible that immunomodulation with checkpoint inhibitors could stimulate the donor anti-leukemia immune response and prevent disease relapse. However, the safety of checkpoint blockade early after allografting remains to be established. Methods:We conducted a pilot study to assess the tolerability and efficacy of Nivolumab, a PD-1 inhibitor, as maintenance therapy after allo-SCT (NCT02985554). Pts were eligible if they were post allo-SCT without evidence of relapse or active graft-vs-host disease (GVHD) or history of prior greater than stage I skin acute GVHD. Nivolumab was to be administered intravenously at 1mg/kg every 2 weeks for 4 doses followed by dosing every 12 weeks. Treatment started 4 weeks after routine immunosuppression was discontinued until 2 years after the transplant. The primary objective was to determine the tolerability of Nivolumab on this schedule. Secondary objectives were evaluation of adverse events, relapse, and overall survival. Results:Four pts were enrolled from December 2017 through November 2018. (Table 1)All pts experienced immune-related adverse events (irAE) from Nivolumab, and 2 (50%) pts experienced serious adverse events. (Table 2)One pt developed grade (G) 4 neutropenia soon after the first dose. (Figure 1)The absolute neutrophil count nadired at 20 cells/µL, at which point pegfilgrastim was administered. An interim bone marrow biopsy (BMBx) confirmed no evidence of relapsed disease. Full neutrophil recovery occurred approximately 3 months after the initial dose, and no subsequent toxicities occurred. Another pt developed G3 autoimmune encephalopathy concurrently with G2 transaminitis and G2 thrombocytopenia after one dose of Nivolumab. (Figure 2)Intravenous methylprednisolone (1mg/kg daily for 3 days) and immunoglobulin (2g/kg in 4 divided doses) were administered, followed by a 7-week steroid taper with full resolution of symptoms. Relapsed disease was ruled out by a BMBx. A third pt developed G2 skin rash approximately 10 days after the first dose of Nivolumab. Skin biopsy demonstrated drug hypersensitivity reaction vs GVHD, and the pt was treated with a 3-week prednisone course (starting at 1mg/kg followed by a taper). A mild flare recurred 2 weeks later, which was treated with topical steroids only. However, Nivolumab was not resumed. The fourth pt developed G2 elevated TSH approximately 2 months into therapy and after 4 doses of Nivolumab. Thyroid hormone replacement was initiated with subsequent symptom improvement and normalization of TSH over a 4-month period. As a result of these unexpected severe toxicities, the study was closed to further enrollment, and further Nivolumab administration ceased. Thus far, one pt (#1) relapsed after a total remission duration of 530 days; the remission duration after starting Nivolumab was 318 days. One pt has mild chronic skin GVHD. All 4 patients remain alive with a median overall survival of 2.3 years (range, 1.9-4.7). Conclusions:Even at low doses, the use of Nivolumab as maintenance therapy in the post allo-SCT setting was not tolerable at the current dosing and schedule due to an unexpected number of high grade irAEs. Additional studies of dose and timing after allo-SCT are needed to improve safety and tolerability, in conjunction with correlative studies to better understand the immunomodulatory processes in the post-transplant setting. Disclosures Kline: Merck: Honoraria; Merck: Research Funding. Stock:Kite, a Gilead Company: Membership on an entity's Board of Directors or advisory committees; Pfizer: Membership on an entity's Board of Directors or advisory committees; Daiichi: Membership on an entity's Board of Directors or advisory committees; Astellas: Membership on an entity's Board of Directors or advisory committees; Agios: Membership on an entity's Board of Directors or advisory committees; UpToDate: Honoraria; Research to Practice: Honoraria. Artz:Miltenyi: Research Funding. Larson:Agios: Consultancy; Novartis: Honoraria, Other: Contracts for clinical trials; Celgene: Consultancy. Riedell:Novartis: Research Funding; Verastem: Membership on an entity's Board of Directors or advisory committees; Celgene: Membership on an entity's Board of Directors or advisory committees, Research Funding; Bayer: Honoraria, Speakers Bureau; Kite/Gilead: Honoraria, Research Funding, Speakers Bureau. Bishop:CRISPR Therapeutics: Consultancy, Membership on an entity's Board of Directors or advisory committees; Novartis: Consultancy, Membership on an entity's Board of Directors or advisory committees, Research Funding; Kite: Consultancy, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Juno: Consultancy, Membership on an entity's Board of Directors or advisory committees; Celgene: Consultancy, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau. Liu:Arog: Other: PI of clinical trial; BMS: Research Funding; Agios: Honoraria; Novartis: Other: PI of clinical trial; Karyopharm: Research Funding. OffLabel Disclosure: Nivolumab used as maintenance therapy in the post-transplant setting


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 2149-2149
Author(s):  
Romil Patel ◽  
Neeraj Y Saini ◽  
Ankur Varma ◽  
Omar Hasan ◽  
Qaiser Bashir ◽  
...  

Abstract Introduction: The role of autologous hematopoietic stem cell transplantation (auto-HCT) in the management of patients with Waldenström Macroglobulinemia (WM), a rare, indolent lymphoma, has not been established. We had previously published our experience with auto-HCT in a small cohort of WM patients1. Here, we present an updated analysis of auto-HCT with a larger cohort of WM patients. Methods and study population: The study cohort was comprised of 29 patients who underwent high-dose chemotherapy and auto-HCT at MD Anderson Cancer Center (MDACC). The Kaplan-Meier method was used to create survival curves. Overall survival (OS) was defined as the duration from date of transplant to death or last date of follow-up in living patients. Progression-free survival (PFS) was defined as the duration from date of transplant to either progressive disease or death, whichever occurred first. Results: Median age at auto-HCT was 60 (range, 43-75 years). Eight patients (28%) had concurrent light chain amyloidosis (AL). Of the five patients who had MYD88 testing completed, 3 were positive for the MYD88 mutation. Additionally, of these 3 patients, 2 were also positive for CXCR4 mutation. Patients received a median of 2 lines (range 1-6) of therapy prior to auto-HCT; 3(10%) patients had primary refractory disease, 8(28%) were in first remission, and 18 (62%) had relapsed disease. Median time from transplant to last follow-up for the surviving patients was 5.3 years. Preparative regimens received by the patients were: Melphalan (n=20), BEAM-R (n=2), Busulfan/Melphalan (n=1), Cyclophosphomaide/Etoposide/total body irradiation (n=1), Thiotepa/Busulfan/Cyclophosphamide (n=1), and Carmustine/Thiotepa (n=1). Three patients further went on to receive allogeneic transplant either after relapse from auto-HCT or due to disease transformation to aggressive lymphoma. Twenty-eight patients achieved engraftment with a median time to neutrophil engraftment of 11 days (range, 10-15 days). One patient suffered primary graft failure due to progression of disease and died 84 days after transplant. Non-relapse mortality was 3.4% at 1 year. All patients were eligible for response evaluation. The median OS from diagnosis was 12.2 years. Overall response rate was 96%: complete response (n=8, 27.6%), very good partial response (n=5, 17.3%), partial response (n=15, 51.7%), and progressive disease (n=1, 3.4%). PFS and OS at 5 years were 43.3% and 62.9%, respectively. Median PFS and OS from auto-HCT were 4.1 and 7.3 years (Fig. 1A). The median OS from auto-HCT in first remission + primary refractory and relapsed disease was 8.2 years and 4.1 years, respectively.16 patients were alive at the time of censoring while 13 patients had died. Causes of death include relapsed disease (n=6), secondary malignancy (n=2), infection (n=1), chronic graft-versus-host disease (n=1), and unknown (n=3). 8 patients (28%) were positive for concurrent AL amyloidosis. The sites of amyloid involvement were kidneys (n=2), lungs (n=1), bone marrow (n=1), heart(n=1), lymph nodes(n=1), gastrointestinal tract (n=1) and subcutaneous fat aspirate(n=5). The median overall survival for patients with amyloid involvement (n=8) was 12 years. On univariate analyses, the number of chemotherapy regimens prior to transplant (≤ 2 vs >2 lines) was the strongest predictor of overall survival (p=0.03, HR 0.3, CI: 0.09-0.9, log-rank) and PFS (p=0.001, HR 0.24, CI: 0.07-0.85, log-rank). The median PFS in patients with ≤ 2 lines and > 2 lines of therapy was 71 months versus 19 months, respectively (Fig. 1B). Conclusion: Auto-HCT is safe and feasible in selected patients with WM, with a high response rate and durable remission even in patients with relapsed or refractory disease. References: Krina Patel et.al. Autologous Stem Cell Transplantation in Waldenstrom's Macroglobulinemia. Blood 2012 120:4533; Disclosures Thomas: Celgene: Research Funding; Bristol Myers Squibb Inc.: Research Funding; Acerta Pharma: Research Funding; Array Pharma: Research Funding; Amgen Inc: Research Funding. Lee:Celgene: Consultancy, Membership on an entity's Board of Directors or advisory committees; Adaptive Biotechnologies Corporation: Consultancy; Amgen: Consultancy, Membership on an entity's Board of Directors or advisory committees; Chugai Biopharmaceuticals: Consultancy; Takeda Oncology: Consultancy, Membership on an entity's Board of Directors or advisory committees; Kite Pharma: Consultancy, Membership on an entity's Board of Directors or advisory committees. Orlowski:Takeda: Consultancy; Celgene: Consultancy; Spectrum Pharma: Research Funding; Janssen: Consultancy; Kite Pharma: Consultancy; Sanofi-Aventis: Consultancy; BioTheryX: Research Funding; Amgen: Consultancy, Research Funding; Bristol-Myers Squibb: Consultancy. Champlin:Otsuka: Research Funding; Sanofi: Research Funding. Patel:Poseida Therapeutics, Inc.: Research Funding; Takeda: Research Funding; Abbvie: Research Funding; Celgene: Research Funding.


Sign in / Sign up

Export Citation Format

Share Document