The Transcriptional Signature of Kinases Inhibited by the Multi-Targeted Kinase Inhibitor AS703569 Is Associated with Clinical Outcome in Multiple Myeloma (MM): Anti-MM Activity of AS703569 in Preclinical Studies.

Blood ◽  
2009 ◽  
Vol 114 (22) ◽  
pp. 730-730
Author(s):  
Jake Delmore ◽  
David N. Cervi ◽  
Douglas McMillin ◽  
Efstathios Kastritis ◽  
Jana Jakubikova ◽  
...  

Abstract Abstract 730 Multi-targeted kinase inhibitors, when associated with manageable toxicity, offer the therapeutically desirable option of targeting, through a single chemical entity, several pathways that may contribute to the complexity and heterogeneity of molecular lesions harbored by neoplasias such as multiple myeloma (MM). However, intractable questions often emerge while prioritizing for preclinical studies different multi-targeted agents with extensive and/or only partially overlapping of sets of known targets. We have hypothesized that the potential therapeutic relevance of a multi-targeted inhibitor may be reflected on the prognostic relevance of its targets' transcriptional signature. We applied this concept in the case of the orally bioavailable multi-targeted kinase inhibitor AS703569, which targets (with IC50 in low nM range) all 3 Aurora kinase (AK) isoforms as well as various other kinases (e.g. cSRC, FGFR1, Flt3, Fyn, Lyn, Rsk1-3, Yes, Axl, et.c.) and evaluated the transcriptional signature of AS703569 kinase targets (with IC50 <10 nM) in MM cells of patients receiving Bortezomib as part of Phase II/III trials (specifically SUMMIT/APEX). We observed that patients with high transcriptional signature of AS703569 targets had inferior progression-free and overall survival (p=0.005 and p=0.012, log-rank test) and also validated that, in a study of tandem autologous transplant, a subset of patients with high levels of this AS703569 target transcriptional signature also have inferior overall survival (p=0.032, log-rank test) compared to cases with low levels of the signature. These observations supported the notion that the kinome space targeted by AS703569 is enriched for targets associated with adverse clinical outcome in MM. In preclinical assays, we observed that AS703569 decreased the viability of MM cell lines and primary CD138+ MM tumor cells in a time- and dose-dependent manner, with IC50 values <50 nM for the majority of cell lines tested; and without evidence of cross-resistance with established anti-MM agents. Combinations of AS703569 with dexamethasone, doxorubicin, or bortezomib did not exhibit antagonism, suggesting that AS703569 can be incorporated in regimens with these established anti-MM drug classes. Interestingly, in vitro compartment-specific bioluminescence imaging (CS-BLI) assays showed that against MM cells which respond to stromal cells with increased proliferation and survival, the anti-MM activity of AS703569 is more pronounced when these MM cells are co-cultured with bone marrow stromal cells than in conventional cultures in isolation. This indicated that AS703569 is capable of overcoming the protective effects that BMSCs confer to MM tumor cells and prompted in vivo validation studies in our orthotopic SCID/NOD model of diffuse MM bone lesions established by i.v. injection of MM-1S-GFP/Luc cells monitored by whole body bioluminescence imaging. AS703569 (50 mg/kg p.o. once weekly)-treated mice had longer overall survival than vehicle-treated mice (median 50.0 days, 95% C.I. 40.3-59.7 days vs. 39.0 days, 95% C.I., 35.4-42.6 days, p=0.019, log-rank test). An alternative schedule of AS703569 at 16.7 mg/kg 3 times/week also resulted in longer overall survival (median 54.0 days, 95% C.I. 33.2-74.8 days, p=0.023, log-rank test). These data indicate that AS703569 exhibits anti-MM activity in vitro and in orthotopic in vivo MM models, and suggests that this multi-targeted inhibitor merits considerations for further preclinical studies, as well as potential clinical studies in MM, especially given the otherwise adverse outcome associated with the inhibitor's target transcriptional signature. Disclosures: Laubach: Novartis: Consultancy, Honoraria. Rastelli:EMD Serono: Employment. Clark:EMD Serono: Employment. Sarno:EMD Serono: Employment. Richardson:Millenium: (Speakers' Bureau up to 7/1/09), Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau; Celgene: (Speakers' Bureau up to 7/1/09), Membership on an entity's Board of Directors or advisory committees, Speakers Bureau. Anderson:Millennium: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Novartis: Consultancy, Honoraria, Research Funding; Celgene: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding. Mitsiades:Millennium: Consultancy, Honoraria; Novartis : Consultancy, Honoraria; Bristol-Myers Squibb : Consultancy, Honoraria; Merck & Co: Consultancy, Honoraria; Kosan Pharmaceuticals: Consultancy, Honoraria; Pharmion: Consultancy, Honoraria; PharmaMar: Patents & Royalties; Amgen: Research Funding; AVEO Pharma: Research Funding; EMD Serono : Research Funding; Sunesis Pharmaceuticals: Research Funding.

Blood ◽  
2013 ◽  
Vol 122 (21) ◽  
pp. 1538-1538
Author(s):  
Aristoteles Giagounidis ◽  
Alan List ◽  
Eva Hellström-Lindberg ◽  
Mikkael A. Sekeres ◽  
Ghulam J. Mufti ◽  
...  

Abstract Introduction The proportion of aberrant metaphases is prognostic for overall survival (OS) in MDS patients with trisomy 8 (Mallo M, et al. Leuk Res. 2011;35:834-6). The impact of the proportion of metaphases with del(5q) on clinical outcomes, including OS, disease progression and response to therapy with LEN in MDS remains undefined. In two large multicenter studies of LEN (MDS-003 and MDS-004) in RBC transfusion-dependent patients with International Prognostic Scoring System (IPSS) Low- or Intermediate (Int)-1-risk del(5q) MDS, RBC transfusion independence (TI) ≥ 8 weeks was achieved in 51–67% of patients (List A, et al. N Engl J Med. 2006;355:1456-65; Fenaux P, et al. Blood. 2011;118:3765-76). This retrospective analysis evaluated response to treatment, progression to acute myeloid leukemia (AML) and OS by proportion of del(5q) metaphases in patients with isolated del(5q) from the MDS-003 and 004 studies. Methods In order to allow sufficient patient numbers for analysis, ≥ 16 metaphases were evaluated for del(5q) by standard karyotyping (MDS-003 and MDS-004) and 200 interphase nuclei were evaluated by fluorescence in situ hybridization (FISH; MDS-004 only) using a probe for the commonly deleted region 5q31 (LSI EGR1/D5S721, Abbott, Wiesbaden, Germany). Patients received LEN on days 1–21 of each 28-day cycle (10 mg) or continuously (5 mg or 10 mg), or placebo. In MDS-004, patients randomized to placebo could cross over to LEN 5 mg by week 16. RBC-TI ≥ 26 weeks, time to AML progression and OS were analyzed by the proportion of del(5q) metaphases or interphases (≤ 60% vs > 60%) using standard karyotyping and FISH, respectively. Results Of the 353 patients from MDS-003 and MDS-004, 194 had isolated del(5q) by standard karyotyping; median proportion of del(5q) metaphases was 96% (range 4–100). Baseline characteristics including age, time from diagnosis, RBC transfusion burden, hemoglobin level, platelet and absolute neutrophil counts were comparable among patients with ≤ 60% (n = 21) and > 60% (n = 173) del(5q) metaphases. Rates of RBC-TI ≥ 26 weeks were similar across patients in the ≤ 60% and > 60% groups (P = 0.6515). Time to AML progression was comparable for patients in the ≤ 60% group versus the > 60% group (log-rank test P = 0.9802); 2-year rates were 22.2% (95% confidence interval [CI]: 7.7–54.5%) and 14.6% (95% CI: 9.9–21.2%), respectively. Time to AML progression was similar when analyzed with death without AML as competing risk (Gray’s test P = 0.5514). OS was longer in the > 60% versus the ≤ 60% group (log-rank test P = 0.0436); median OS was 3.7 years (95% CI: 3.0–4.2) and 2.4 years (95% CI: 1.5–4.9), respectively. In MDS-004, the proportion of del(5q) interphases was analyzed using FISH in 106 patients, including 46 with ≤ 60% and 60 with > 60%. When analyzed by FISH, rates of RBC-TI ≥ 26 weeks were similar across patients in the ≤ 60% and > 60% groups (P = 1.000). Time to AML progression and OS were similar across these groups (log-rank test P = 0.7311 and P = 0.8639, respectively) when analyzed by FISH. In the ≤ 60% and > 60% groups respectively, 2-year AML progression rates were 14.8% (95% CI: 6.9–30.1%) and 18.6% (95% CI: 10.4–32.0%), and median OS was 3.1 years (95% CI: 2.3–4.8) and 2.9 years (95% CI: 2.3–4.4). Time to AML progression was similar when analyzed with death without AML as competing risk (Gray’s test P = 0.8631). Conclusions In IPSS Low- or Int-1-risk MDS patients with isolated del(5q) treated with LEN in MDS-003 and MDS-004 studies, baseline characteristics, RBC-TI ≥ 26 weeks and AML progression were comparable in patients with > 60% versus ≤ 60% del(5q) metaphases. Although similar across groups when analyzed by FISH in a subset of patients, surprisingly, OS was longer in patients with > 60% del(5q) metaphases than in those with ≤ 60% del(5q) metaphases by standard karyotyping. However, the number of patients with ≤ 60% del(5q) metaphases was limited and no adjustment was made for multiple testing. These findings suggest that the number of cells with the isolated del(5q) abnormality measured by FISH does not impact clinical outcome in this RBC transfusion-dependent study population, but this finding could not be confirmed for OS by standard karyotyping. Disclosures: Giagounidis: Celgene: Consultancy, Honoraria, Membership on an entity’s Board of Directors or advisory committees. List:Celgene: Serve on Celgene Data Safety & Monitoring Committee Other. Hellström-Lindberg:Celgene: Membership on an entity’s Board of Directors or advisory committees, Research Funding. Sekeres:Celgene: Membership on an entity’s Board of Directors or advisory committees; Amgen: Membership on an entity’s Board of Directors or advisory committees. Mufti:Celgene: Honoraria, Membership on an entity’s Board of Directors or advisory committees, Research Funding. Schlegelberger:Celgene: Consultancy. Morrill:Celgene: Employment, Equity Ownership. Wu:Celgene: Employment, Equity Ownership. Skikne:Celgene: Employment, Equity Ownership. Fenaux:Celgene: Honoraria.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 4264-4264
Author(s):  
Justyna Bartoszko ◽  
Tony Panzarella ◽  
Caroline Jane McNamara ◽  
Anthea Lau ◽  
Aaron D. Schimmer ◽  
...  

Abstract Introduction. Myelofibrosis is a disease characterized by aberrant bone marrow function with eventual fibrosis. Current widely-used disease prognostic indices, such as the Dynamic International Prognostic Scoring System (DIPSS) do not take into account comorbidities, which may have significant effects on patient survival as well as disease course. We sought to describe the comorbidity distribution in this patient population and assess the impact of comorbidities as scored by two different widely used scales in clinical practice, the Adult Comorbidity Evaluation 27 (ACE-27) and the Hematopoietic Cell Transplant Comorbidity Index (HCT-CI), on overall survival and leukemic transformation in myelofibrosis. A score of 3 on ACE-27 or ≥3 on HCT-CI generally indicates a high burden (severe) comorbidities. Methods. We conducted a retrospective study of 309 patients seen at the MPN program at the Princess Margaret Cancer Centre, with a confirmed diagnosis of myelofibrosis [primary myelofibrosis (PMF), post-essential thrombocytopenia (PET-MF) or post-polycythemia vera (PPV-MF)]. Patients were seen from 1999-2014 with a median follow-up time of 2 years. Time to death and leukemic transformation was examined from the date of first presentation to our centre. Our primary aim was to examine the impact of comorbidity scores, as assessed by ACE-27 and the HCT-CI, on overall survival. In a secondary analysis we examined the impact of comorbidity scores on leukemic transformation. Multivariable Cox proportional hazards models were constructed for the primary and secondary outcomes. A series of descriptive analyses were carried out examining the distribution of various comorbidities as captured by the two scales. Results. The most common comorbidities captured by ACE-27 were hypertension (n=92, 22.3%), diabetes mellitus (n=43, 10.4%), venous disease (n=26, 6.3%), solid tumour including melanoma (n=26, 6.3%), and angina/coronary artery disease (n=23, 5.6%). The most common comorbidities captured by HCT-CI were cardiac (n=49, 17.3%), diabetes (n=43, 15.2%), mild hepatic (n=28, 9.9%), cerebrovascular disease (n=25, 8.8%), prior solid tumour (n=24, 8.5%). The distribution of comorbidity scores as compared between scales is shown in Table 1. A total of 78 patients (25.2%) experienced the primary outcome of interest, which was all-cause death. For the primary outcome of overall survival, there were differences across groups of patients with different comorbidity scores using ACE-27 or HCT-CI, with the highest severity groups having worse outcomes (Figure 1). Progressively increasing DIPSS categories (Low, Intermediate-1, Intermediate-2, and High risk) were also associated with worse overall survival. On multivariable survival analysis, an ACE-27 score of 3 when compared to a lower score of 0-1 was associated with an almost two-fold increase in the risk of all-cause death [HR 1.95 (95% CI 1.06-3.58), p=0.03]. On multivariable analysis, an HCT-CI score of 3+ when compared with 0-1 was marginally significantly associated with an increased risk of all-cause death [HR 1.60 (95% CI 0.96-2.68), p=0.07]. Interaction terms were tested between the scores and age at presentation and no effect of age on survival across varying severities of comorbidities was found. In our secondary analysis, there was no impact of the ACE-27 or HCT-CI on leukemic transformation. Conclusions. ACE-27 picked up severe co-morbidities in 13% of patients in our cohort while HCT-CI picked up severe comorbidities in 23%. Although the incidence of severe co-morbidities was lower when assessed by ACE-27, the overall impact on survival of severe comorbidities as assessed by both scores is likely to be similar. The presence of severe comorbidities at the time of diagnosis conferred a significant survival disadvantage in patients with myelofibrosis, but had no impact on progression to leukemic transformation. Table Overall survival by ACE-27 comorbidity category, showing differences between categories of comorbidity severity (p=0.047, log rank test). Table. Overall survival by ACE-27 comorbidity category, showing differences between categories of comorbidity severity (p=0.047, log rank test). Figure Figure. Disclosures Panzarella: Cellgene: Consultancy. Schimmer:Novartis: Honoraria. Schuh:Amgen: Membership on an entity's Board of Directors or advisory committees. Yee:Novartis Canada: Membership on an entity's Board of Directors or advisory committees, Research Funding. Gupta:Novartis: Consultancy, Honoraria, Research Funding; Incyte: Consultancy, Research Funding.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 2175-2175 ◽  
Author(s):  
Michael H. Albert ◽  
Mary Slatter ◽  
Andrew Gennery ◽  
Tayfun Guengoer ◽  
Henric-Jan Blok ◽  
...  

Abstract PV and AL contributed equally Multiple studies from the EBMT registry and others have shown excellent survival rates after allogeneic haematopoietic stem cell transplantation (HSCT)for Wiskott-Aldrich syndrome (WAS) patients (Ozsahin et al, Blood 2008). The importance of myeloid engraftment for full disease correction has also been demonstrated (Moratto et al, Blood 2011). However, the vast majority of HSCTs in these studies were performed with (oral) busulfan/cyclophosphamide-based conditioning and in the early 2000 years or before. In 2005, the inborn errors working party (IEWP) of EBMT and ESID first recommended busulfan/fludarabine (BuFlu) or treosulfan/fludarabine (TreoFlu) based conditioning for primary immunodeficiencies such as WAS, with some centers deciding to add thiotepa (TT) to the conditioning. We performed a retrospective analysis via the EBMT and SCETIDE registries of WAS patients transplanted between 01/01/20006 and 12/31/2016 with these two regimens. The primary objective was to compare the overall (OS) and event-free survival (EFS) after HSCT with either BuFlu±TT or TreoFlu±TT conditioning. Secondary objectives included the influence of either conditioning regimen on acute and chronic GVHD, the degree of donor chimerism, incidence of secondary procedures after HSCT (2nd HSCT, stem cell boost, DLI, gene therapy or splenectomy) and rates of disease-specific complications after HSCT. At the time of this interim analysis, 174 patients were included, 92 (53%) with BuFlu±TT and 82 (47%) with TreoFlu±TT conditioning with a median age of 1.57 years (range 0.21-29.96) at HSCT and a median follow-up of 32.9 months (range 1.5-128.9). The donor was an HLA-matched sibling (MSD) in 30, a matched related donor (MRD) in 5, a matched unrelated donor (MUD, 9/10 or 10/10) in 105, a mismatched unrelated donor (MMUD, <9/10) in 9 and a mismatched family donor (MMFD) in 25 (18 with ex-vivo T-cell depletion). Stem cell source was bone marrow in 93 (53%), peripheral blood in 62 (36%) and cord blood in 18 (10%). Two year overall survival (OS) of the entire cohort was 88.6% (95% confidence interval 83.5%-93.6%). There was no significant difference in OS between patients treated with BuFlu±TT or TreoFlu±TT conditioning (2-year OS 88.1% vs. 89.5%; log-rank test p=0.7). Patients aged >5 years had a worse OS as compared to those 5 years or younger at HSCT (74.9% vs. 90.8%; log-rank test p=0.005). The type of donor had no influence on OS: 96.4% for MSD/MFD, 86.8% for MUD/MMUD and 87.7% for MMFD (log-rank test p=0.4). Whole blood chimerism was complete (>90% donor) in 60/75 evaluable patients (80%) at last follow-up or before secondary procedure (if a patient had one), 39/40 (98%) in the BuFlu±TT group and 21/35 (60%) in the TreoFlu±TT group. Twenty-six patients required a secondary procedure: stem cell boost in 4 patients, donor lymphocyte infusion in 9, 2nd HSCT in 15 and splenectomy in 1. Twenty-two of these 26 (84.6%) are alive and 14 of 16 with available chimerism data have a complete donor chimerism (>90%donor) at last follow-up. The 2-year cumulative incidence (CI) of secondary procedures was higher at 33.9% in the TreoFlu±TT versus 12.8% in the BuFlu±TT group (Gray's test p=0.017), and 2-year EFS (secondary procedure or death as event) was 61.4% in the TreoFlu±TT and 75.0% in the BuFlu±TT group (log-rank test p=0.2). Grade II-IV acute GVHD had the same incidence in both groups (100 day CI 24.4% vs. 26.3%; Gray's test p=0.849) and chronic GVHD of any grade was borderline more frequent in the TreoFlu±TT group (2 year CI 17.2% vs 6.7%; Gray's test p=0.054). The cumulative incidence of disease-specific complications occurring more than 6 months post HSCT (severe infections, bleeding, autoimmunity) was not different between the two groups (6.5% vs. 6.4%; Gray's test p=0.92). There was no malignancy reported after HSCT except for one EBV-post-transplant lymphoproliferative disorder (PTLD) 2.7 months after HSCT. In summary, HSCT with either BuFlu±TT or TreoFlu±TT conditioning reliably cures almost 90% of patients with WAS regardless of donor type. WAS-related complications are very rare events more than 6 months post HSCT. More patients required secondary procedures after treosulfan-based than busulfan-based conditioning. These data confirm the feasibility and efficacy of the regimens currently recommended by the IEWP. Disclosures Slatter: Medac: Other: Travel assistance. Chiesa:Gilead: Consultancy; Bluebird Bio: Consultancy. Kalwak:Sanofi: Other: travel grants; medac: Other: travel grants. Locatelli:Novartis: Consultancy, Membership on an entity's Board of Directors or advisory committees; bluebird bio: Consultancy; Bellicum: Consultancy, Membership on an entity's Board of Directors or advisory committees; Amgen: Honoraria, Membership on an entity's Board of Directors or advisory committees; Miltenyi: Honoraria. Sykora:Aventis-Behring: Research Funding; medac: Research Funding. Zecca:Chimerix: Honoraria. Veys:Pfizer: Honoraria; Servier: Research Funding; Novartis: Honoraria.


Blood ◽  
2010 ◽  
Vol 116 (21) ◽  
pp. 923-923 ◽  
Author(s):  
Apostolia Maria Tsimberidou ◽  
William G. Wierda ◽  
Sijin Wen ◽  
William Plunkett ◽  
Susan O'Brien ◽  
...  

Abstract Abstract 923 Background: To enhance the response rate with a decrease in myelosuppression that were observed with oxaliplatin, fluradabine, Ara-C, and rituximab (OFAR1) (Tsimberidou et al, J Clin Oncol, 2008;26:196), the daily dose of oxaliplatin was increased from 25 to 30mg, the daily dose of Ara-C was decreased from 1 g/m2 to 0.5 g/m2 and the optimal number of days of fluradabine and Ara-C administration was explored (OFAR2). Methods: OFAR2 consisted of oxaliplatin 30mg/m2 D1-4; fludarabine 30mg/m2; Ara-C 0.5g/m2; rituximab 375mg/m2 D3; and pelfigrastim 6mg D6. Fludarabine and Ara-C were given on D2-3 (level 1) D2-4 (level 2) or D2-5 (level 3) every 4 weeks. Tumor lysis, DNA virus, and PCP prophylaxis was administered. A “3+3” design was used (Phase I) and and the planned number of patients in the Phase II was 90 (CLL, 60; RS, 30). Results: Overall 102 patients (rel. CLL 67, RS 35) were treated. Twelve patients were treated in the Phase I portion of the study. Dose-limiting toxicities were noted in 2/3 patients at level 3 (G4 diarrhea and G4 sepsis). Level 2 was the maximum tolerated dose. Ninety patients (CLL, 60; RS, 30) were treated in Phase II portion of the study (age > 60 years 67%, 17p del 37.5%, 11q del 15%, 13q del 18%, +12, 17%; neg. 12.5%; unmutated IgVH 81.5%, ZAP70-positive 77%, and CD38 30%, 63%). Response in 80 of 90 patients (Phase II) is shown in Table (too early, n=10). The overall response rates in patients (Phase II) with 17p deletion and 11q deletion were 29% and 41%, respectively. Twenty-nine patients underwent SCT after OFAR2 (response status to OFAR2 at the time of SCT: CR, n=3; nPR, n=2; 15; no response, n=9). With a median follow-up of 20.8 months, the median survival was 19 months (95% CI, 13–37+) and the median FFS was 6 months (95% CI, 3.4 – 8.2). Overall, 238 cycles were administered. G3-4 neutropenia, thrombocytopenia, and anemia were noted in 67%, 74%, and 44% of patients (51%, 64%, and 25% of cycles); and G3-4 infections in 19% of patients. Clinical outcomes of OFAR2 were compared with those of OFAR1. In patients with RS, the overall response rate was 41% (11/27) with OFAR2 and 50% (10/20) with OFAR1 (p = 0.57, Fisher's test); the median survival with OFAR2 and OFAR1 was 8.3 months and 18+ months, respectively (p = 0.92, log-rank test); and the respective median FFS was 3.0 months and 4.1 months (p = 0.40, log-rank test). In patients with CLL, the overall response rate was 55% (29/53) with OFAR2 and 33% (10/30) with OFAR1 (p = 0.36, Fisher's test); the median survival with OFAR2 was 21.4 months and 13.8 months with OFAR1 (p = 0.19, log-rank test); and the respective median FFS was 6.6 months and 4.9 months (p = 0.69, log-rank test). Conclusion: OFAR2 induced response in 41% of patients with RS and 55% of patients with relapsed/refractory CLL in the phase II study. Antileukemic activity was also noted in patients with 17p deletion. Although the numbers of patients are small, OFAR1 was associated with a trend towards superior clinical outcomes in patients with RS compared to OFAR2; and OFAR2 was associated with a trend towards superior clinical outcomes compared to OFAR1 in patients with relapsed/refractory CLL. Disclosures: Tsimberidou: Sanofi: Membership on an entity's Board of Directors or advisory committees, Research Funding; ASCO: Career Development Award, Research Funding. Off Label Use: Drug: Oxaliplatin. Oxaliplatin combined with fludarabine, cytarabine, and rituximab has antileukemic activity in patients with relapsed/refractory Chronic Lymphocytic Leukemia and Richter Syndrome. Wierda:Genentech: Consultancy, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Celgene: Consultancy, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Micromet: Consultancy, Membership on an entity's Board of Directors or advisory committees; GlaxoSmithKline: Research Funding; Abbott Laboratories: Research Funding. O'Brien:Biogen Idec: Consultancy, Membership on an entity's Board of Directors or advisory committees, Research Funding; Genentech: Consultancy, Research Funding. Kipps:Sanofi Aventis: Research Funding. Jones:GlaxoSmithKline: Consultancy, Membership on an entity's Board of Directors or advisory committees; Abbott Laboratories: Research Funding.


Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 2745-2745 ◽  
Author(s):  
Deborah L. White ◽  
Liu Lu ◽  
Timothy P. Clackson ◽  
Verity A Saunders ◽  
Timothy P Hughes

Abstract Abstract 2745 Ponatinib is a potent pan-BCR-ABL tyrosine kinase inhibitor (TKI) currently in a pivotal phase 2 clinical trial. Ponatinib (PON) was specifically designed to target both native and all mutant forms of BCR-ABL, including T315I. The phase I study of oral ponatinib in patients with refractory CML/ALL or other hematologic malignancies recently reported that 66% and 53% of patients with CP-CML achieved MCyR and CCyR respectively (Cortes et al., ASH 2011 abstract #210). While extensive modelling experiments in BaF3 cells have been performed characterising in vitro response to ponatinib, little is known about the interactions of this drug and drug transporters that impact the response of other tyrosine kinase inhibitors (TKIs). To explore this we have examined both the degree of in vitro kinase inhibition mediated by ponatinib in BCR-ABL+ cell lines, and the intracellular uptake and retention (IUR) of ponatinib achieved. The IC50 was determined by assessing the reduction in %p-Crkl in response to increasing concentrations of ponatinib in vitro. The IUR assay was performed as previously using [14-C]-ponatinib. To determine the role of ABCB1 and ABCG2, both previously implicated in the transport of other TKIs, IC50 analysis was performed on K562 cells, and variants; ABCB1 overexpressing K562-DOX and ABCG2 overexpressing K562-ABCG2. As shown in Table 1, in contrast to the results previously observed with imatinib (IM), nilotinib (NIL) and dasatinib (DAS) there was no significant difference in the IC50ponatinib between these three cell lines, suggesting neither ABCB1 nor ABCG2 play a major role in ponatinib transport. Furthermore, the addition of either the ABCB1 and ABCG2 inhibitor pantoprazole, or the multidrug resistance (MDR) inhibitor cyclosporin did not result in a significant change in the IC50ponatinib in any of the cell lines tested. In contrast the addition of either pantoprazole or cyclosporin resulted in a significant reduction in IC50IM, IC50NIL. and IC50DAS of K562-DOX cells, supporting the notion that these TKIs interact with ABCB1.Table 1:The IC50 of ponatinib (compared to IM, NIL and DAS) in K562 cells and the over-expressing variants DOX and ABCG2 in the presence of the ABC inhibitors pantoprazole and cyclosporin. n=5. *p<0.05IC50% reduction in IC50+ pantoprazole+ cyclosporinPON (nM)IM (μM)NIL (nM)DAS (nM)PONIMNILDASPONIMNILDASK5627.793751111544*NA−107NA2DOX7.919*598*100*1018*63*1655*88*ABCG26.4730025*6NA To further examine the effect of ABC transporters on ponatinib efflux we have determined the IUR of [14-C]-ponatinib in K562, DOX and ABCG2 cell lines. We demonstrate no significant difference in the IUR between these cell lines at 37°C (n=6) (K562 vs DOX p=0.6; K562 vs ABCG2 p=0.37 and DOX vs ABCG2 p=0.667 at 2uM respectively). Temperature dependent IUR experiments reveal a significant reduction in the ponatinib IUR at 4°C compared to 37°C in K562 cells (n=6) (p=0.008), DOX cells (p=0.004) and ABCG2 cells (p=0.002) supporting the likely involvement of an ATP/temperature dependent, and yet to be determined, component of ponatinib influx. There was no significant difference in the IUR between these cell lines at 4°C (p=0.824, p=0.7 and p=0.803 respectively). Importantly, these data are consistent with the IC50ponatinib findings. If ATP dependent efflux pumps (ABCB1 and ABCG2) were actively transporting ponatinib, a significant decrease in IUR in DOX and ABCG2 at 37°C compared to K562 cells would be expected, but is not observed here. Analysis of ponatinib IUR in the prototypic ABCB1 over-expressing CEM-VBL100 cells, and their parental, ABCB1 null counterparts (CCRF-CEM) further confirmed these findings. The IUR in VBL100 cells was significantly higher than that observed in CEM's (p<0.001; n=5), providing further evidence that ponatinib was not being exported from the cell actively via ABCB1. These data suggest that the transport of ponatinib is, at least in part, temperature-dependent indicating a yet to be determined ATP transporter may be involved in the transport of ponatinib into leukaemic cells. Importantly, this data suggests that ponatinib is unlikely to be susceptible to resistance via the major ATP efflux transporters (ABCB1 or ABCG2) that have been previously demonstrated to significantly impact the transport of, and mediate resistance to other clinically available TKIs. Disclosures: White: BMS: Honoraria, Research Funding; Novartis Pharmaceuticals: Honoraria, Research Funding. Clackson:ARIAD: Employment. Hughes:Novartis: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; BMS: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; ARIAD: Honoraria, Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2012 ◽  
Vol 120 (21) ◽  
pp. 4602-4602
Author(s):  
Iwona Hus ◽  
Dariusz Jawniak ◽  
Magdalena Gorska-Kosicka ◽  
Aleksandra Butrym ◽  
Justyna Dzietczenia ◽  
...  

Abstract Abstract 4602 Purpose: In this study, we carried out a retrospective analysis of the efficacy and toxicity of bendamustine in patients with B-cell lymphoproliferative diseases. Methods: Bendamustine was administered both in monotherapy and combined protocols in 111 patients, including 81 patients with chronic lymphocytic leukaemia (CLL), 20 patients with indolent lymphoma, and 10 patients with aggressive lymphoma (8 mantle cell lymphoma and 2 diffuse large B-cell lymphoma). Median age of patients was 61 years (range: 44 – 87 years). Almost all patients, except 3 previously untreated patients with CLL had relapsed (78 patients) or refractory disease (30 patients). 60.4% of patients were treated with bendamustine plus rituximab, while 28.2% received bendamustine in monotherapy, and 31.4% received other combined regimens. Results: Overall response rate (ORR) was 65.8%, including 20.7% of complete response (CR) and 45.1% of partial response (PR). In CLL patients, ORR was 59.3% with 16% of CR and 43.2% of PR. In patients with indolent NHL, ORR was 80%, including 30% of CR and 50% of PR. In patients with aggressive lymphoma, OR and CR rates were respectively: 90% and 40%. In patients with CLL, a likelihood of response was significantly lower in patients with ZAP-70 expression (p=0.006) and 17p deletion (p=0.009). Median overall survival (OS) for all patients was 11.5 months (range 1–40). For CLL patients, median OS was 11 months (range 1–40), for patients with indolent lymphoma 15.5 months (range 5–29 ) and for patients with aggressive lymphoma - 10 months (6–38). Median OS was significantly longer in patients responding to therapy as compared to non-responders (15 months vs. 8 months; p=0.0001). Median progression-free survival (PFS) in all patients was 6 months (0–38), including 4 months (0–38) for CLL patients, and 8.5 months (0–33) for patients with both aggressive and indolent lymphoma. Among pre-treatment parameters, β2-microglobulin (RR=1.234; p=0.002), haemoglobin level (RR=0.803; p=0.03) and PLT count (RR=1.005; p=0.03) significantly influenced survival. Also, the higher number of bendamustine cycles was associated with longer overall survival (RR=0.715; p=0.003). In patients with CLL, 17p deletion was associated with reduced overall survival (p=0.021; log-rank test). OS was significantly longer in patients who received ≤ 2 lines of previous therapies as compared to > 3 lines (p=0.018; log-rank test), and who received ≥ 4 courses of therapy (p=0.02; log-rank test). In patients with NHL, both OS and PFS were significantly longer in patients with lactate dehydrogenase level <250 U/l (p= 0.01; p=0.02; respectively), and who received ≥ 4 cycles of treatment (P= 0.03). Toxicity was predominantly haematological, including grade III/IV neutropenia in 32%, thrombocytopenia in 15% and anaemia in 16% of patients. Conclusion: Bendamustine, both in monotherapy and combination regimens, is an effective therapy with a favourable toxicity profile even in heavily pretreated patients. Disclosures: Hus: Mundipharma: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees; Roche: Consultancy, Honoraria; Glaxosmithkline: Consultancy. Off Label Use: In Europe bendamustine is indicated for first-line treatment of chronic lymphocytic leukaemia (Binet stage B or C) in patients for whom fludarabine combination chemotherapy is not appropriate and indolent non-Hodgkin's lymphomas as monotherapy in patients who have progressed during or within 6 months following treatment with rituximab or a rituximab containing regimen. Wiktor-Jedrzejczak:Genzyme: Speakers Bureau; Celgene: Speakers Bureau; Genopharm: Speakers Bureau; Bayer: Consultancy; Pfizer: Consultancy; Novartis: Consultancy, Speakers Bureau; Amgen: Consultancy; Janssen-Cilag: Consultancy; Bristol-Myers Squibb: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau. Walewski:Roche: Honoraria, Speakers Bureau; GSK: Membership on an entity's Board of Directors or advisory committees; Celgene: Membership on an entity's Board of Directors or advisory committees; Janssen-Cilag: Membership on an entity's Board of Directors or advisory committees; Cephalon: Research Funding; Mundipharma: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding, Speakers Bureau. Dmoszynska:Millenium Pharmaceuticals: Consultancy; Celgene: Consultancy; Mundipharma: Consultancy, Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 4208-4208 ◽  
Author(s):  
Kristin A. Simar ◽  
Vishwanath Sathyanarayanan ◽  
Amir K Issa ◽  
Mohamed Amin Ahmed ◽  
Mansoor Noorani ◽  
...  

Abstract Background: Due to ~50% risk of relapse with standard therapy (rituximab, cyclophosphamide, doxorubicin, vincristine and prednisone, R-CHOP), an increasing number of patients with high risk diffuse large B-cell lymphomas (DLBCL) are being treated with dose-adjusted (DA) EPOCH-R (rituximab, etoposide, doxorubicin, vincristine, prednisone, cyclophosphamide). DA-EPOCH-R contains a 96-hour continuous infusion can be delivered either in the inpatient or outpatient setting, via use of ambulatory infusion pumps. Potential advantages of outpatient therapy include reduced inpatient burden for routine chemotherapy, less exposure to resistant bacterial infections, increased patient satisfaction, and reduced cost. The ability to administer outpatient DA-EPOCH-R is dependent on the ability of the healthcare facility to administer the regimen safely while maintaining dose adjustments and schedule. We hypothesize that patients who receive DA-EPOCH-R as an outpatient have similar outcomes and toxicity rates as patients who receive the regimen as an inpatient. We further hypothesize that there is a significant cost benefit for patients to receive DA-EPOCH-R as an outpatient. Methods: This was a retrospective database analysis of newly diagnosed consecutive DLBCL patients ≥ 18 years of age who received DA-EPOCH-R chemotherapy at MD Anderson Cancer Center between 2010 and 2014. Patients with double hit lymphoma defined as having a MYC and BCL2 or BCL6 translocation were excluded due to their aggressive nature. We descriptively analyzed demographic variables in this population including, age, gender, international prognostic index (IPI)) and outcome (overall response rates (ORR), complete response (CR), progression free survival (PFS), overall survival (OS), and hospital admission for neutropenic fever events). Additionally, we evaluated the number of outpatient cycles received in relation to survival outcomes and neutropenic fever events. Statistical analysis was done using Fisher's exact test or Chi-square test to evaluate the association between two categorical variables and Wilcoxon rank sum test was used to evaluate the difference in a continuous variable between patient groups. Kaplan-Meier method was used for time-to-event analysis including overall survival and progression free survival. The Log-rank test was used to evaluate the difference in time-to-event endpoints between patient groups. Results: A total of 196 patients had data available for analysis, with 138 patients (70.4%) receiving all cycles as an inpatient, while 58 patients (29.6%) received at least 1 outpatient cycle of DA-EPOCH-R (Table 1). Compared with patients who received no outpatient cycle, the patients who received any outpatient therapy were younger, had a lower proportion of high IPI, and experienced fewer episodes of febrile neutropenia. The median OS and PFS for the entire population has not been reached, with a median follow-up time for the censored observations of 2.78 years (range: 0.24 - 8.64 years). The difference in OS between the patients who had any outpatient therapy and no outpatient therapy was not significant by the log-rank test (p-value=0.11). The difference in PFS between the patients who had any outpatient therapy and no outpatient therapy was marginally significant for OS by the log-rank test (p-value=0.07). Our cost analysis for 6 cycles of inpatient DA-EPOCH-R is estimated to be ~$88K, or $14.6K/cycle. The cost savings incurred for chemotherapy only expenses for each outpatient cycle is at least $3.3K/cycle or $19.8K for a total of 6 cycles. Conclusion: DA-EPOCH-R is a highly effective regimen for treating aggressive DLBCL which can be administered in an outpatient setting safely, efficaciously, and in a cost-effective manner without any apparent effect on outcome or rate of admission for neutropenic fever. There can be savings of about of nearly $20K per patient for a 6-cycle course of therapy. In our series, patients who received outpatient therapy were younger and may have had greater social support, which could potentially confound results. This retrospective analysis supports the use of outpatient DA-EPOCH-R, but additional studies are warranted to evaluate which patients may benefit most. Disclosures Oki: Novartis: Research Funding. Nastoupil:Janssen: Other: Travel, Accommodations, Expenses, Research Funding; TG Therapeutics: Research Funding; Celgene: Honoraria; AbbVie: Research Funding. Fowler:Pharmacyclics, LLC, an AbbVie Company: Consultancy, Research Funding; Janssen: Consultancy, Research Funding. Wang:Asana BioSciences: Research Funding; Acerta: Consultancy, Research Funding; Dava Oncology: Honoraria; BeiGene: Research Funding; Kite Pharma: Research Funding; Juno Therapeutics: Research Funding; Janssen: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Asana biosciences, Beigene, Celgene, Juno, Kite, Onyx, Pharmacyclics: Research Funding. Fayad:Seattle Genetics: Consultancy, Research Funding. Westin:Celgene: Membership on an entity's Board of Directors or advisory committees; Chugai: Membership on an entity's Board of Directors or advisory committees; Spectrum: Membership on an entity's Board of Directors or advisory committees; ProNAi: Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 4035-4035
Author(s):  
Jing Ai ◽  
Danielle Boselli ◽  
Lauren M. Bohannon ◽  
Thomas G. Knight ◽  
Brittany Ragon ◽  
...  

Abstract Background: The hypomethylating agents (HMAs) azacitidine and decitabine have been increasingly used in the frontline setting for elderly and/or unfit patients with acute myeloid leukemia (AML). While these therapies are oftentimes used in the palliative setting, HMAs have also been used with curative intent in some patients, as a bridge to allogeneic hematopoietic cell transplantation (HCT). It is well known that 4-6 cycles of HMA therapy can be necessary to achieve a response; however, it is still common in practice for treating physicians to stop HMAs earlier when a rapid response is not observed. Few studies have investigated time to response in the setting of frontline HMA treatment for AML. Methods: We retrospectively evaluated all patients who were initiated on frontline HMAs for AML at our institution from September 2013 to April 2018. HMAs were administered without dose reduction or treatment delays. Responses were evaluated and categorized as hematologic response (HR; defined by neutrophils > 1000/µL, platelets > 100k/µL, transfusion independence, and no peripheral blasts), complete remission (CR), CR with incomplete platelet recovery (CRp), CR with incomplete hematologic recovery (CRi), and no response. Patient, cytogenetic, and treatment characteristics were summarized and described. We evaluated patients exhibiting a response before receiving 4 cycles of HMA therapy ("early" responders) and compared these patients to those achieving a response after receiving at least 4 cycles of HMA ("late" responders). The Kaplan-Meier method estimated overall survival (OS). Odds of early response were estimated with logistic regression models. Results: During the study period, 137 patients received frontline HMAs. 122 (89.1%) patients received azacitidine, and 15 (10.9%) were treated with decitabine. Mean age at HMA start was 70.1 years (range 39.8-94.3). Most patients (62.8%) were male, and most (77.4%) were Caucasian. 21 (15.3%) patients were NCCN favorable-risk, 52 (38.0%) were intermediate-risk, and 64 (46.7%) were poor-risk. At first testing, 21.4% of the patients were positive for the FLT3/ITD mutation, 0.8% the FLT3/TKD mutation, and 21.4% the NPM1 mutation. Some patients received concomitant therapies along with the HMA; these included hydroxyurea, lenalidomide, and sorafenib. Median survival in the entire population was 11.6 (95% CI 8.3, 15.2) months. Overall response rate (ORR) was 60.6%. Among responders, 80.7% achieved HR as their first response; 19.3% were first noted to have marrow responses (1.2% CR, 4.8% CRp, and 13.3% CRi). Most patients did not undergo bone marrow evaluation to assess response. Among responding patients, 60.2% responded "early", whereas 39.8% responded "late". Median overall survival was 15.2 (9.3, 17.7) months in early responders, and 22.2 (11.7, 38.9) months in late responders. There was no difference in survival between the groups (p=0.108, log-rank test; Figure 1), although there was a trend toward improved survival in late responders. Nineteen patients underwent allogeneic HCT. Time to response was not associated with odds of receiving HCT (p=0.812). Conclusions: HMAs have a high ORR as frontline therapy in elderly and unfit AML patients. Among AML patients receiving frontline HMAs, later responders have equivalent survival to earlier responders. Physicians should consider exercising patience when treating AML patients with HMAs. These findings warrant validation in a larger, prospective study. Figure. Figure. Disclosures Avalos: Juno: Membership on an entity's Board of Directors or advisory committees. Grunwald:Alexion: Consultancy, Membership on an entity's Board of Directors or advisory committees; Ariad: Consultancy, Membership on an entity's Board of Directors or advisory committees; Genentech: Research Funding; Incyte Corporation: Consultancy, Membership on an entity's Board of Directors or advisory committees, Research Funding; Celgene: Consultancy, Membership on an entity's Board of Directors or advisory committees; Merck: Consultancy, Membership on an entity's Board of Directors or advisory committees; Amgen: Consultancy, Membership on an entity's Board of Directors or advisory committees, Research Funding; Pfizer: Consultancy, Membership on an entity's Board of Directors or advisory committees; Agios: Consultancy, Membership on an entity's Board of Directors or advisory committees; Forma Therapeutics: Research Funding; Janssen: Research Funding; Cardinal Health: Consultancy, Membership on an entity's Board of Directors or advisory committees; Medtronic: Equity Ownership.


Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 5153-5153
Author(s):  
Steffen Koschmieder ◽  
Mirle Schemionek ◽  
Christian Elling ◽  
Utz Krug ◽  
Torsten Kessler ◽  
...  

Abstract Abstract 5153 Introduction: Pazopanib is a tyrosine kinase inhibitor with proven activity against metastatic renal cancer. Due to its target spectrum of kinases (PDGFR, KIT, and VEGFR), we sought to investigate its activity against myeloid malignancies. Methods: 32D cells transduced with FIP1L1-PDGFRA or several activating PDGFRA point mutations as well as AML cell lines were analyzed for the effects of pazopanib vs. imatinib on cell growth, MTS activity, 7-AAD positivity, and clonogenic growth. Pazopanib and imatinib were purchased from LC Laboratories Woburn, MA, USA. Results: Pazopanib was found to decrease the growth of FIP1L1-PDGFRA transduced cell lines at low nanomolar concentrations (10 nM). 1000 nM doses were equally effective as 1000 nM of the PDGFR tyrosine kinase inhibitor imatinib. MTS assays confirmed that at 10 nM, pazopanib reduced cell proliferation to 28% of that of vehicle-treated control cells (DMSO 0.05%), while 100 nM of pazopanib completely abolished cell growth and suppressed MTS activity. Analysis of 7-AAD positivity using flow cytometry showed that reduction in cell growth and MTS activity was due to induction of apoptosis (17+/−0.6%, 63+/−0.5%, and 87+/−0.5%, 86+/−0.1% for 10, 100, and 1000 nM of pazopanib and 1000 nM of imatinib, respectively). Interestingly, while two PDGFRA activating point mutations (H650Q and R748G) recently identified in patients with idiopathic hypereosinophilic syndrome-type myeloproliferative neoplasms (MPN) were equally sensitive against pazopanib and imatinib, two other such point mutations showed differential sensitivity, with Y849S being more sensitive to imatinib and N659S being more sensitive to pazopanib. When evaluating the effect of pazopanib on acute myeloid leukemia cell lines, we found that imatinib inhibited the cell growth and reduced colony forming units of Kasumi-1 and BCR-ABL positive K562 but not MV4-11 or NB-4 cells, while Kasumi-1, MV4-11, and NB-4 cells were sensitive towards pazopanib treatment. Moreover, while the clonogenic growth of bone marrow-derived cells from two control patients (lymphoma or AML in remission) were only reduced to 75% of control with pazopanib treatment in vitro (100 nM), 100 nM pazopanib decreased colony growth to 18% in another patient with AML at diagnosis. Conclusion: Our data suggest that pazopanib may be active against a variety of myeloid neoplasms and that clinical studies assessing its efficacy are warranted. Also, cells may show differential sensitivity against the PDGFR and KIT inhibitors pazopanib and imatinib. Disclosures: Koschmieder: GlaxoSmithKline: Honoraria, Membership on an entity's Board of Directors or advisory committees; Novartis: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding. Off Label Use: Data on in vitro activity of pazopanib and imatinib in MPN and AML discussed. Krug:MedA Pharma: Honoraria; Novartis: Honoraria; Alexion: Honoraria; Boehringer Ingelheim: Research Funding; Sunesis: Honoraria. Müller-Tidow:Novartis: Research Funding.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 2635-2635
Author(s):  
Weiguo Zhang ◽  
Guopan Yu ◽  
Hongying Zhang ◽  
Charlie Ly ◽  
Bin Yuan ◽  
...  

Abstract Fms-like tyrosine kinase 3 (FLT3)-targeted therapy represents an important paradigm in the management of patients with highly aggressive FLT3 mutated acute myeloid leukemia (AML). However, clinical efficacy is usually transient and followed by emergence of resistance to FLT3-inhibitors (Borthakur et al., 2011; Cortes et al., 2013; Zhang et al., 2008). Such resistance often results from acquired mutations of TKD, which are frequently identified in D835, Y842 and F691 residues (Smith et al., 2015; Smith et al., 2012; Zhang et al., 2014). It was reported that the FLT3-ITD-targeting drug sorafenib can induce autophagy in human myeloid dendritic cells (Lin et al., 2013). Induction of autophagy has also been reported to play a crucial role in resistance to BCR-ABL targeted imatinib therapy in CML (Hekmatshoar et al., 2018). Additionally, inhibition of autophagy can re-sensitize cancer cells to apoptosis induction (Fitzwalter et al., 2018; Piya et al., 2017), suggesting that inhibition of autophagy may represent a novel therapeutic strategy for overcoming resistance to FLT3-targeted therapy. In the present study, we assessed autophagy levels in leukemia cell lines bearing different FLT3 mutations and in AML patient samples obtained from sorafenib-resistant patients. All tested resistant cell lines bearing TKD or ITD+TKD mutations showed increased basal autophagy levels. Resistant AML patient samples also demonstrated greater autophagy compared to matched pre-treatment samples in FLT3-mutated, but not in FLT3-wild type samples. Upregulation of autophagy was also observed in the bone marrow (BM)-mimetic microenvironment (i.e., hypoxia and the presence of mesenchymal stem cells (MSCs) in vitro. Inhibition of autophagy with chloroquine (CQ) potentiated quizartinib-induced apoptosis and partially abrogated MSC-mediated protection in FLT3-ITD- and/or D835-mutated AML cells by suppressing c-Myc, mTOR/S6K signaling and activating transcription factor 4 (ATF4). We also observed upregulation of BTK activation accompanied by increased autophagy levels in hypoxic/MSC co-culture with leukemic cells and in resistant primary patient samples. Co-targeting BTK and FLT3 with ibrutinib (or BTK siRNA) and quizartinib enhanced leukemic cell killing and abrogates MSC-mediated protection of FLT3 mutated leukemia cells. We further investigated a novel, highly potent small molecule pan-FLT3/pan-BTK kinase inhibitor CG-806 (IC50s 0.8 and 5.0 nM against FLT3-ITD and BTK, respectively) (Aptose, San Diego, CA). CG'806 abolished MSC/hypoxia-mediated protection of AML cells and induced apoptosis in FLT3-mutated cells in vitro. Of note, CG'806, but not quizartinib, exerted profound pro-apoptotic effects in primary AML patient cells harboring ITD+D835 mutations ex vivo. Further evaluation in a PDX leukemia model inoculated with the ITD+D835 mutated primary AML cells showed that CG'806 significantly reduced leukemia cell burden and benefited for mouse survival. Taken together, autophagy is associated with AML resistance to FLT3-targeted therapy, which can be overcome by the pan-FLT3/pan-BTK kinase inhibitor CG-806 through concomitant blockade of FLT3 and BTK. Co-targeting FLT3 and BTK might provide a strategy for preventing/overcoming FLT3 inhibitor resistance in AML patients with FLT3 mutations. Phase I trials of CG'806 are in preparation. Disclosures Zhang: Aptose Biosciences, Inc: Employment. Battula:United Therapeutics Inc.: Patents & Royalties, Research Funding. Konopleva:Stemline Therapeutics: Research Funding. Rice:Aptose Biosciences, Inc: Equity Ownership. Andreeff:Eutropics: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; Reata: Equity Ownership; SentiBio: Equity Ownership; Oncolyze: Equity Ownership; Astra Zeneca: Research Funding; Jazz Pharma: Consultancy; Oncoceutics: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; United Therapeutics: Patents & Royalties: GD2 inhibition in breast cancer ; Aptose: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; Daiichi-Sankyo: Consultancy, Patents & Royalties: MDM2 inhibitor activity patent, Research Funding; Amgen: Consultancy, Research Funding; Celgene: Consultancy.


Sign in / Sign up

Export Citation Format

Share Document