flt3 inhibitors
Recently Published Documents


TOTAL DOCUMENTS

309
(FIVE YEARS 116)

H-INDEX

31
(FIVE YEARS 8)

Author(s):  
Yazan Numan ◽  
Zaid Abdel Rahman ◽  
Justin Grenet ◽  
Stephanie Boisclair ◽  
Jan Philipp Bewersdorf ◽  
...  

2021 ◽  
Author(s):  
Xinhui Zheng ◽  
Xiangjun Li ◽  
Er-Lie Jiang ◽  
Liwei Lv

Abstract The purpose of this paper is to systematically analyze the outcome of FLT3 inhibitors maintenance treatment following hematopoietic stem cell transplantation (HSCT) for patients suffering from FLT3-ITD-mutated acute myeloid leukemia (AML). Pubmed, Embase, and Cochrane Library databases were retrieved before November 2021. Fifteen studies were included eventually containing six without control and nine with control. Thirteen studies evaluated sorafenib, and two assessed quizartinib and midostaurin, separately. Via survival analysis, the main outcomes in the FLT3 inhibitors group were improved greatly with the hazard ratio(HR) of overall survival of 0.38 (95% confidence interval [CI], 0.29-0.49; P < 0.001), HR of leukemia-free survival of 0.35 (95%CI, 0.27-0.47; P < 0.001) and HR of cumulative incidence of relapse of 0.32 (95%CI, 0.23-0.45; P < 0.001). Moreover, the TKI use didn’t seem to increase the incidence of graft-versus-host disease (GVHD) and adverse effects in statistics. Through subgroup analysis, MRD-positive patients before and after HCT, and MRD-negative patients before HCT might benefit a lot from sorafenib maintenance.


2021 ◽  
Vol 23 (2) ◽  
Author(s):  
Yuta Kaito ◽  
Mitsuhito Hirano ◽  
Muneyoshi Futami ◽  
Masanori Nojima ◽  
Hideto Tamura ◽  
...  

Author(s):  
Mohammad Azhar ◽  
Zachary Kincaid ◽  
Meenu Kesarwani ◽  
Arhama Ahmed ◽  
Mark Wunderlich ◽  
...  

Kinase activating mutation in FLT3 is the most frequent genetic lesion associated with poor prognosis in acute myeloid leukemia (AML). Therapeutic response to FLT3 tyrosine kinase inhibitor (TKI) therapy is dismal, and many patients relapse even after allogenic stem cell transplantation. Despite the introduction of more selective FLT3 inhibitors, remissions are short-lived, and patients show progressive disease after an initial response. Acquisition of resistance-conferring genetic mutations and growth factor signaling are two principal mechanisms that drive relapse. FLT3 inhibitors targeting both escape mechanisms could lead to a more profound and lasting clinical responses. Here we show that the JAK2 inhibitor, momelotinib, is an equipotent type-1 FLT3 inhibitor. Momelotinib showed potent inhibitory activity on both mouse and human cells expressing FLT3-ITD, including clinically relevant resistant mutations within the activation loop at residues, D835, D839, and Y842. Additionally, momelotinib efficiently suppressed the resistance mediated by FLT3 ligand (FL) and hematopoietic cytokine activated JAK2 signaling. Interestingly, unlike gilteritinib, momelotinib inhibits the expression of MYC in leukemic cells. Consequently, concomitant inhibition of FLT3 and downregulation of MYC by momelotinib treatment showed better efficacy in suppressing the leukemia in a preclinical murine model of AML. Altogether, these data provide evidence that momelotinib is an effective type-1 dual JAK2/FLT3 inhibitor and may offer an alternative to gilteritinib. Its ability to impede the resistance conferred by growth factor signaling and activation loop mutants suggests that momelotinib treatment could provide a deeper and durable response; thus, warrants its clinical evaluation.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 2350-2350
Author(s):  
David Kipp ◽  
Sun Loo ◽  
Andrew Charles Perkins ◽  
Steven W Lane ◽  
Emily Blyth ◽  
...  

Abstract Introduction Despite the advent of targeted therapy for FLT3-mutated AML, unmet need still exists for patients unfit for intensive chemotherapy, with no evidence that overall survival (OS) can be improved by combining either venetoclax (Konopleva et al., ASH 2020) or gilteritinib (Astellas press release, December 2020) with azacitidine. Although gilteritinib has been shown to improve median OS from 5.5 to 9.8 months, the majority will relapse (Perl et al., 2019). Adaptive on-target gilteritinib resistance may be due to the FLT3-F691L gatekeeper mutation, whereas off-target resistance may be due to loss-of-function variants in CBL, which encodes an E3 ubiquitin-protein ligase that negatively regulates FLT3 (McMahon et al, 2019). Ponatinib is a type-1 FLT3 inhibitor that is active in vitro against FLT3 F691L (Smith et al., 2013) and had an overall response rate (ORR) of 43% in a small pilot phase-I study (Talpaz et al., 2011). Combination of a FLT3 inhibitor with azacitidine may antagonize the synergistic hypermethylation reported for FLT3-ITD in association with epigenetic mutations (Shih et al., 2015). CBL loss-of-function mutations may also enhance responsiveness to FLT3 inhibitors (Taylor et al, 2015). We thus hypothesize that the combination of ponatinib and azacitidine could mitigate the rapid evolution of drug resistance typical of more selective FLT3 inhibitors used as single agents. Methods A phase-Ib study was conducted with the primary objective safety and key secondary objective preliminary efficacy of azacitidine in combination with ponatinib in patients with FLT3-ITD AML failing prior therapy or unfit for intensive chemotherapy. Exploratory objectives included mechanisms of ponatinib resistance and responsiveness of CBL-mutant AML to FLT3 inhibition. At dose level 1 (DL1), patients received azacitidine 60 mg/m 2 on days 1-5 and 8-9 and ponatinib 30 mg daily on days 5-25 of each cycle. In patients not achieving CR or CRi after cycle 1, the ponatinib dose was increased to 45 mg during cycle 2. For dose level 2 (DL2), the dose of azacitidine was increased to 75 mg/m 2. Results Thirty-one patients were evaluable for response. Median age was 67 years (range, 26-87). Frequency of prior lines of therapy was 0 (15%), 1 (46%), 2 (23%) or 3 (8%). Four patients had a history of prior allogeneic hematopoietic cell transplant and one had previously received a FLT3 inhibitor. FLT3-ITD was present in 28 patients (median VAF 0.33; range, 0.009-17.95) and 3 had inactivating CBL mutations. A total of 20 patients were treated at DL1 and 12 patients at DL2. There were two grade-4 DLTs (raised AST/ALT [DL1] and tubulointerstitial nephritis [DL2]). Three grade-2 thromboembolic events were observed (two cannula-related DVTs and a distal lower-limb DVT). There were two grade-5 AEs (infection and cardiac failure), which were not considered drug related. The most common grade-3-4 AEs were febrile neutropenia (57%), neutropenia (47%), infections (47%), thrombocytopenia (40%) and anaemia (27%). Cardiac arrhythmias (atrial fibrillation/flutter, bradycardia, sinus tachycardia and ventricular tachycardia [1 patient]) were observed in 30% of patients. Of these, 80% were grade 1 or 2 and only one was considered by the investigator to be related to study treatment. Response was evaluable in 23 of 31 patients. Nine patients (39%) achieved CR or CRi, 3 (13%) achieved a PR and 8 (35%) achieved SD (ORR 52%). ORR at DL1 and DL2 was 43% and 66%, respectively. Median time to best response was 1.4 months (range 1.0-11.9). Median duration of best response was 12.9 months at both dose levels. Median OS for DL1 was 6.5 months and not reached for DL2. Despite shorter follow-up, DL2 patients experienced better OS than DL1 patients (p = 0.015). Responses were seen in 2 of 4 patients with post-allograft relapse. Two of three patients with a CBL mutation responded (1 CR and 1 CRi). Eradication of the CBL mutation was seen in one patient, who remains on therapy after 15 cycles. Molecular studies to investigate dynamic changes in molecular architecture are ongoing. Conclusions The recommended phase-II dose of ponatinib is 30 mg on days 5-25 and that of azacitidine is 75 mg/m 2 for seven doses each cycle. The ORR was 52% and durable disease control was observed, especially in patients receiving DL2. Preliminary efficacy was observed in CBL-mutated patients. Further clinical investigation of this regimen is warranted in patients with FLT3- or CBL-mutant AML. Figure 1 Figure 1. Disclosures Kipp: Novartis: Honoraria. Perkins: Celgene: Consultancy; Novartis: Consultancy, Honoraria, Speakers Bureau; Abbvie: Honoraria, Speakers Bureau. Lane: Novartis: Consultancy; Geron: Consultancy; BMS: Consultancy, Research Funding; Abbvie: Honoraria; Astellas: Membership on an entity's Board of Directors or advisory committees. Enjeti: Sanofi: Honoraria; Novartis: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; AbbVie: Honoraria; Roche: Speakers Bureau; Astra Zeneca: Honoraria. Bajel: Abbvie, Amgen, Novartis, Pfizer: Honoraria; Amgen: Speakers Bureau. Reynolds: Novartis AG: Current equity holder in publicly-traded company; Abbvie: Research Funding; Alcon: Current equity holder in publicly-traded company. Wei: Abbvie, Amgen, Astellas, AstraZeneca, Celgene/BMS, Genentech, Janssen, MacroGenics, Novartis, Pfizer, and Servier: Membership on an entity's Board of Directors or advisory committees; Novartis, Abbvie, Celgene/BMS: Speakers Bureau; Former employee of Walter and Eliza Hall Institute: Patents & Royalties: Prof. Andrew Wei is a former employee of the Walter and Eliza Hall Institute and is eligible for a fraction of the royalty stream related to Venetoclax; Abbvie, Amgen, Astellas, AstraZeneca, Celgene/BMS, Genentech, Janssen, MacroGenics, Novartis, Pfizer, and Servier: Honoraria; Servier: Consultancy; Abbvie, Amgen, AstraZeneca, Celgene/BMS, Novartis, Servier and F. Hoffmann-La Roche: Research Funding. OffLabel Disclosure: Ponatinib - used as an experimental therapy for AML in combination with azacitidine


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 3324-3324
Author(s):  
Seiji Fukuda ◽  
Nozomi Matsuda

Abstract RUNX1 generally functions as a tumor suppressor in the hematopoietic system. However, RUNX1 expression is significantly elevated in human AML cells with FLT3/ITD mutations, promotes leukemogenesis induced by FLT3/ITD (Behrens et al. JEM 2017) and enhances the resistance of FLT3/ITD + cells to type-II FLT3 inhibitor quizartinib (Hirade et al IJH 2016). We previously reported that RUNX1 expression is higher in CXCR4-low FLT3/ITD + cells compared to Cxcr4-high FLT3/ITD + cells, even though Cxcr4 expression is trans-activated by RUNX1. This difference in RUNX1 expression level was associated with divergent response to CXCL12 in FLT3/ITD + cells harboring different CXCR4 expression levels that were exposed to quizartinib (Fukuda S. et al. ASH 2019). Our data also demonstrated that RUNX1 expression is down-regulated following withdrawal of quizartinib in FLT3/ITD + cells that became refractory to quizartinib (Hirade et al. IJH 2016), suggesting that RUNX1 expression may be up-regulated by quizartinib in FLT3/ITD + cells. Since RUNX1 regulates proliferation of FLT3/ITD + AML cells, the present study investigated association between RUNX1 expression levels and proliferation of quizartinib resistant FLT3/ITD + cells that are exposed to quizartinib. In the sensitive FLT3/ITD + Ba/F3 cells, RUNX1 protein expression was transiently up-regulated but eventually down-regulated by 5 nM quizartinib, coincident with decline in the viable cells. In contrast, RUNX1 expression was up-regulated by quizartinib and remained elevated in the resistant FLT3/ITD + Ba/F3 cells. Since RUNX1 enhances proliferation of FLT3/ITD + cells, we next examined whether proliferation FLT3/ITD + cells that acquired resistance to quizartinib is facilitated by quizaritinib as a result from quizartinib-mediated up-regulation of RUNX1, using the Cxcr4-low and Cxcr4-high FLT3/ITD + cells that acquired resistance to quizartinib. Although CXCL12 barely enhanced the proliferation of refractory FLT3/ITD + Ba/F3 cells, 5 nM quizartinib significantly increased the proliferation of both Cxcr4-low and Cxcr4-high FLT3/ITD + Ba/F3 cells that acquired resistance to quizartinib compared to those without quizartinib. This increase in the proliferation of Cxcr4-low and Cxcr4-high FLT3/ITD + Ba/F3 cells coincided with the elevation in RUNX1 and CXCR4 protein expression. Moreover, the resistant Cxcr4-low FLT3/ITD + Ba/F3 cells proliferated significantly faster than Cxcr4-high FLT3/ITD + cells, with concomitant higher expression of RUNX1 in Cxcr4-low FLT3/ITD + cells than in Cxcr4-high FLT3/ITD + cells. Likewise, type-I FLT3 inhibitor gilteritinib significantly enhanced proliferation of Cxcr4-low and Cxcr4-high FLT3/ITD + Ba/F3 cells that acquired resistance to gilteritinib. Knocking down Runx1 using shRNAs significantly decreased the enhanced proliferation induced by quizartinib in refractory FLT3/ITD + Ba/F3 cells, coincident with reduction in CXCR4 expression. Since CXCR4 expression level was elevated by quizartinib in the FLT3/ITD + cells refractory to quizartinib, we next examined CXCL12-induced migration in quizartinib-resistant FLT3/ITD + cells following exposure to quzartinib. Pre-incubating the quizartinib resistant Cxcr4-low or Cxcr4-high FLT3/ITD + Ba/F3 cells with 5 nM quizartinib for 72 hours significantly enhanced their migration to 100 ng/ml of Cxcl12 compared to those without quizartinib, coincident with elevation in RUNX1 levels. Surprisingly, migration of CXCR4-low FLT3/ITD + cells to CXCL12 was significantly elevated compared to CXCR4-high cells, with concomitant higher expression of RUNX1 in Cxcr4-low FLT3/ITD + cells than in Cxcr4-high FLT3/ITD + cells. Silencing Runx1 using shRNAs significantly decreased migration to CXCL12 in refractory Cxcr4-low FLT3/ITD + Ba/F3 cells. These data indicate that the FLT3 inhibitor itself can facilitate the proliferation and migration to CXCL12 in FLT3/ITD + cells that are refractory to FLT3 inhibitors by up-regulating RUNX1. The results implicate that FLT3 inhibitors may worsen the disease progression in the patients that became refractory to FLT3 inhibitors by facilitating proliferation and migration to CXCL12 of the resistant FLT3/ITD + AML cells. In this regard, targeting RUNX1 may represent additional strategy to eradicate resistant FLT3/ITD + AML cells, in which their proliferation and migration are supported by FLT3 inhibitors. Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 3333-3333
Author(s):  
Han Zhong Pei ◽  
Xiaomei Zhuang ◽  
Ming Yang ◽  
Yao Guo ◽  
Zhiguang Chang ◽  
...  

Abstract Frequently mutated in Acute myeloid leukemia (AML), FLT3 is considered as one of the favorable targets for treatment. The FLT3 internal tandem duplication (ITD) mutation enhances kinase activity and causes hyperactivation of downstream signal transduction. Several small molecule FLT3 inhibitors have developed, but their clinical efficacy is limited due to generation of drug resistance. In this study, we define a new mechanism of drug resistance toward tyrosine kinase inhibitors (TKIs). Initially, we found a rapid decrease in the protein level of tumor suppressor p53 in FLT3-ITD-positive MV4-11 and MOLM13 cells and peripheral blood mononuclear cells (PBMCs) from FLT3-ITD AML patients upon treatment with TKIs including sorafenib, sunitinib and quizartinib. The decrease is not caused by changes in mRNA expression as revealed by qPCR analyses but rather by accelerated protease degradation because the p53 protein was stabilized by proteasome inhibitor MG132. Furthermore, treatment of cells with RG7388, a potent disruptor of p53 and MDM2 interaction, prevented the TKI-induced p53 loss. Since MDM2 is the most important E3 ligase responsible for ubiquitination of p53, the data suggest that TKIs may lead to the degradation of p53 by promoting ubiquitination. Indeed, ubiquitination assays verified that TKIs promoted K48 poly-ubiquitination of p53. Previous studies have demonstrated that activations of FLT3 downstream signaling components such as ERKs and Akt reduce p53 protein stability through ubiquitination by activating MDM2. It is somewhat unexpected that inhibition of FLT3-ITD and its downstream signaling pathways also resulted in decreased p53 stability due to increased ubiquitination. We treated FLT3-ITD-containing cells with specific ERK, AKT and STAT5 inhibitors. Interestingly, while inhibition of ERKs and AKT had no significant effect on the stability of p53, STAT5 inhibition resulted in a reduced level of p53 accompanied by increased K48 poly-ubiquitination. We further analyzed the interaction of p53 with MDM2 in AML cells by using immunoprecipitation. The results showed that the p53-MDM2 interaction was significantly enhanced after treatment with TKIs and STAT5 inhibitors, which was diminished in the presence of RG7388. Subcellular fractionation revealed the presence of p53 and STAT5 in both nucleus and cytoplasm. Treatment of cells with TKIs resulted in a decreased level of p53 and STAT5 in the nucleus, and immunoprecipitation of nuclear proteins with a p53 antibody revealed a reduced association of p53 with STAT5. Taken together, the data suggest that FLT3 inhibitors inhibited nuclear translocation of STAT5 and reduced its interaction of p53 thereby facilitating p53/MDM2 interaction and subsequent ubiquitination and degradation of p53. This study reveals a novel mechanism by which drug resistance to TKIs may occur and further support the use of MDM2/p53 interaction inhibitors in combination with TKIs for treatment of AML. Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 1163-1163
Author(s):  
Jonelle K. Lee ◽  
Mario Scarpa ◽  
Aditi Chatterjee ◽  
Moaath Mustafa Ali ◽  
Prerna Singh ◽  
...  

Abstract BACKGROUND Internal tandem duplication in the fms-like tyrosine kinase 3 receptor tyrosine kinase (FLT3-ITD) is present in acute myeloid leukemia (AML) in 30% of patients, associated with poor treatment outcomes due to rapid relapse. FLT3 inhibitors are used in the clinic, but with incomplete efficacy and development of resistance. Further treatment options are needed. The serine/threonine kinase proviral integration site for Moloney murine leukemia virus (Pim-1) is upregulated downstream of FLT3-ITD; it directly stimulates cell growth and inhibits apoptosis, and also phosphorylates and stabilizes FLT3 in a positive feedback loop in cells with FLT3-ITD. Dual targeting of Pim-1 and FLT3 is a promising treatment strategy. The c-Myc transcription factor contributes to dysregulation of cell growth and apoptosis in cancers, including AML. In addition to transcriptional regulation, c-Myc is regulated post-translationally by T58 phosphorylation by the serine/threonine kinase glycogen synthase kinase-3- β (GSK-3β). Here we show that concurrent treatment of cells with FLT3-ITD with Pim and FLT3 inhibitors activates GSK-3β, which phosphorylates and post-translationally downregulates c-Myc. METHODS Ba/F3-ITD and MV4-11 cells, with FLT3-ITD, and FLT3-ITD AML patient blasts were cultured with the pan-Pim inhibitor AZD1208 (1 μM) and/or the FLT3 inhibitors gilteritinib or quizartinib (15 nM, 1 nM), with and without the GSK-3β inhibitor TCG-24 (20 μM). c-Myc, p-GSK3-α/β (S21/9) and GSK3-α/β protein expression was measured by immunoblotting. c-Myc mRNA was measured by qPCR. Cells were also cultured with cycloheximide (100 μg/mL) with and without the proteasome inhibitor MG-132 (20 μM) to measure protein half-life and proteasomal degradation. To study the role of c-Myc overexpression and activation, Ba/F3-ITD cells were infected with retroviral estrogen receptor (ER)-Myc plasmid, causing c-Myc nuclear translocation when activated by 4-hydroxytamoxifen (4-OHT; 300 nM). To study the role of c-Myc phosphorylation at T58, Ba/F3-ITD cells were infected with MycT58A plasmid, preventing c-Myc phosphorylation at T58. Apoptosis was detected by Annexin V and propidium iodide staining, measured by flow cytometry. RESULTS Treatment with Pim inhibitor AZD1208 and FLT3 inhibitor gilteritinib or quizartinib combination rapidly downregulated c-Myc protein expression in Ba/F3-ITD and MV4-11 cells, with FLT3-ITD, and in primary FLT3-ITD AML patient blasts, compared to quizartinib or gilteritinib alone. Pim inhibitor and FLT3 inhibitor combination treatment did not decrease c-Myc mRNA levels, but markedly decreased c-Myc protein half-life, from 36 mins without drugs and 24 mins with gilteritinib to 18 mins with combination. Half-life did not decrease when cells were pre-treated with the proteasome inhibitor MG-132, consistent with post-translational downregulation through proteasomal degradation. Apoptosis induction by Pim inhibitor and FLT3 inhibitor combination decreased by more than 50% in Ba/F3-ITD cells infected with ER-Myc plasmid and treated with 4-OHT, demonstrating the major role of c-Myc downregulation in apoptosis induction by combination treatment. GSK-3b is inactivated by phosphorylation, and combination treatment rapidly decreased p-GSK-3b levels, while total GSK-3b levels were unchanged, indicating activation of GSK-3b. Treatment of cells with FLT3-ITD with the GSK-3b inhibitor TCG-24 in addition to Pim and FLT3 inhibitors abrogated c-Myc protein downregulation, demonstrating that Pim and FLT3 inhibitor combination downregulates c-Myc through activation of GSK-3b. Finally, Pim and FLT3 inhibitor combination treatment did not downregulate c-Myc in Ba/F3-ITD cells transfected with c-Myc T58A, preventing c-Myc phosphorylation at T58, showing that c-Myc phosphorylation at T58 is necessary for its downregulation by combination treatment. CONCLUSIONS Concurrent treatment of cells with FLT3-ITD with Pim kinase inhibitor enhances the efficacy of FLT3 inhibitors through activation of GSK-3β and GSK-3β-mediated phosphorylation of c-Myc at T58, with resulting c-Myc downregulation through increased proteasomal degradation. This work and previous work in our laboratory on PP2A activating drugs and FLT3 inhibitor combination (Mol Cancer Ther 20:676, 2021) support GSK-3β activation as a mechanism for enhancing efficacy of FLT3 inhibitors in AML with FLT3-ITD. Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 1225-1225
Author(s):  
Mahesh Swaminathan ◽  
Mai M M Aly ◽  
Katherine G. Akers ◽  
Seongho Kim ◽  
Harry Ramos ◽  
...  

Abstract Background: Mutations in FMS like tyrosine kinase gene 3 (FLT3) are the most common genetic aberrations in acute myeloid leukemia (AML). Internal tandem duplications (ITD) are the most frequent FLT3 mutations, occurring in 20-30% of patients (pts) with newly diagnosed AML, and a minority are tyrosine kinase domain mutations. FLT3 mutation confers adverse prognosis in pts treated with standard chemotherapy (chemo) alone, and relapse risk is also higher compared to FLT3-wild type AML (HR-1.75). Studies of FLT3 inhibitors as monotherapy or in combination with chemo in the treatment of pts with FLT3-ITD relapsed/refractory (R/R) AML are associated with mixed results. No evidence synthesis has yet focused on the effect of FLT3 inhibitor alone in the treatment of R/R AML and high-risk myelodysplastic syndrome (HR-MDS). Hence, we analyzed the efficacy of various FLT3 inhibitors from different clinical trials used as monotherapy to treat pts with R/R AML and HR-MDS. Methods: We conducted a systematic review and meta-analysis of clinical trials of FLT3 inhibitors for pts with R/R AML and HR-MDS. We searched EMBASE and PubMed for clinical trials published between 1/1/2000 and 5/26/2021 using a combination of keywords and subject headings related to FLT3 inhibitors and AML. Titles/abstracts and full texts were screened by two independent reviewers, with conflicts arbitrated by a third reviewer. Studies were included if they were (1) full-length published journal articles that (2) reported the results of single-arm or double-arm phase I/II/III clinical R/R AML or HR-MDS. Outcomes of interest were composite response rate (CRc=complete response + complete response with incomplete count recovery) and overall response rate (ORR). Results: Thirty studies were included after the application of inclusion criteria (Figure 1). Two were excluded due to inadequate representation of pts with R/R AML and outcomes reported for stem cell transplant following FLT3 inhibitor use, respectively. Hence, 28 studies were included in the meta-analysis, with a total of 1927 pts. Quizartinib and sorafenib were the most frequently studied inhibitors (6 and 5 studies, respectively, Table 1). Heterogeneity testing was performed using Cochran's Q test and I 2 values. The Cochran's Q test p-value was less than 0.10 (p&lt;0.01), and the I 2 value was more than 50% (I 2=87%), indicating the presence of heterogeneity. Thus, random-effects models were used. Asymmetry test was performed using Egger's linear regression test, which suggested the presence of publication bias (p=0.001; Figure 2). Thus, the trim-and-fill method was used to adjust for publication bias. Pooled ORR was 53% (95% CI, 43-63%) (Figure 2). Similarly, pooled CRc was 34% (95% CI, 26-44%, Figure 3). Pooled CRc and ORR were higher in studies involving quizartinib (40% and 61%, respectively) and gilteritinib (40% and 52%, respectively, Table 1). The CRc rates seen with FLT3 inhibitors are higher than the historical comparison of pts with FLT3-mutated-R/R AML treated with chemo only (CRc ~15%). This meta-analysis shows that FLT3 inhibitors as monotherapy leads to meaningful clinical responses in pts with FLT3-mutated-R/R AML. Conclusion: Most FLT3 inhibitors are effective as monotherapy for the treatment of pts with FLT3-mutated-R/R AML. Pooled response rates are notably higher in studies involving second-generation FLT3 inhibitors, particularly quizartinib and gilteritinib. Though not conclusive, the efficacy spectrum of various FLT3 inhibitors in the R/R setting could help design future studies and guide appropriate treatment selection; however, further validation is needed. Prospective clinical trials are required to compare the effectiveness of newer generation FLT3 inhibitors in pts with FLT3-mutated-R/R AML. Figure 1 Figure 1. Disclosures Maciejewski: Bristol Myers Squibb/Celgene: Consultancy; Novartis: Consultancy; Regeneron: Consultancy; Alexion: Consultancy. Balasubramanian: Servier Pharmaceuticals: Research Funding.


Sign in / Sign up

Export Citation Format

Share Document