scholarly journals JMML Fetal Identity Results Either from Retention of a Physiologic Signature or Aberrant Activation of Master Oncofetal Regulators

Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 4-5
Author(s):  
Marion Strullu ◽  
Aurélie Caye-Eude ◽  
Loïc Maillard ◽  
Chloé Arfeuille ◽  
Elodie Lainey ◽  
...  

Objectives: Juvenile myelomonocytic leukemia (JMML) is a rare but aggressive myeloproliferative/myelodysplastic neoplasm affecting infants and young children. The narrow age-window of onset suggests that a prenatal environment is needed for JMML oncogenesis. In search of a transcriptional reminiscence of embryo-fetal characteristics that would confirm this hypothesis, we investigated how the gene expression profile of JMML hematopoietic progenitors compared to their healthy counterpart isolated at different stages of ontogeny. Methods: Hematopoietic stem cell and progenitor cell (HSPC) fractions of JMML (n=16), bone marrow (BM) of healthy children (n=7), fetal liver (FL; n=3) and fetal BM (FBM; n=2) were phenotyped and sorted using signatures validated in the fetal and adult BM (Notta et al, Science 2011). RNAseq was performed using the TruSeq® Stranded Total RNASample preparation kit. Unsupervised hierarchical clustering analysis was done with the Bioconductor edgeR package. Differentially expressed genes were identified with the Bioconductor limma package. Results: To eliminate the impact of variations in the HSPC distribution, the JMML transcriptome was assessed on FACS-sorted common myeloid progenitor (CMP), granulocyte-monocyte progenitor (GMP) and megakaryocyte-erythroid progenitor (MEP) cell fractions from 16 JMML and compared to healthy counterparts at different stages of ontogeny (FL, FBM, age-matched children BM). Unsupervised hierarchical clustering separated the samples into 4 groups (C1-4), primarily according to ontogeny, with 14/15 embryo-fetal fractions in C1 and all healthy post-natal progenitors in C2 (CMP, MEP) or C3 (GMP). Most JMML fractions clustered either with the prenatal fractions (C1; 17/47 fractions from 8/16 patients) or in a separate group containing no healthy sample (C4; 23/47 samples from 10/16 patients). Two groups were defined accordingly: one with JMML resembling embryo-fetal samples ('Fetal-JMML'; n=6/16), and a JMML-specific group ('Onco-JMML'; 8/16). Patients with Onco-JMML tended to be older, with a more severe presentation and elevated fetal hemoglobin levels. All PTPN11-mutated JMML were in this group whereas 5/6 Fetal-JMML had NRAS or KRAS mutations. Analysis of differential gene expression between Fetal and Onco-JMML highlighted 344 up-regulated genes versus 19 up-regulated genes in Onco-JMML. Surprisingly, LIN28B and WT1, both known to activate fetal pathways, were the most up-regulated genes in Onco-JMML. These key transcription factors were deregulated as early as the hematopoietic stem cell (HSC) compartment. Gene Set Enrichment Analysis (GSEA) confirmed enrichment in LIN28B and WT1-related signatures and showed enrichment in an AML signature in Onco-JMML. On the other hand, Fetal-JMML showed striking overexpression of monocytic /dendritic cell markers and inflammasome and innate immunity components. GSEA confirmed the strong monocyte identity of Fetal-JMML progenitors compared to onco-JMML or healthy postnatal progenitors. Part of the monocytic markers 'aberrantly' expressed in JMML progenitors was expressed in healthy fetal progenitors. Analysis of the HSC and multipotent progenitor (MPP) fractions showed that up regulation of monocytic markers was limited to the JMML progeny compartments. As we were able to confirm the transcriptional and functional identity of the sorted progenitors, these data suggest an early monocytic priming in these JMML progenitors, reminiscent of the monocyte-biased myelopoiesis characterizing physiologic fetal hematopoiesis. Conclusion: Our findings give a striking example of how ontogeny-related features are involved in childhood oncogenesis. They highlight a strong but complex link beween JMML and development, with a fetal identity resulting either from retention of a physiologic fetal monocytic signature or from aberrant (re)activation of master oncofetal regulators. Intriguingly, although LIN28B is thought to reprogram hematopoietic progenitors into a fetal-like state, its activation does not lead to an overall fetal profile in JMML, suggesting a regulatory mechanism distinct from that of physiological development. These two ontogeny-based signatures are likely to uncover the biology underlying previous classifiers based on AML-like profile or DNA methylation and suggest that a subset of JMML patient may benefit from immunomodulating therapies. Disclosures Dalle: Bellicum: Consultancy, Honoraria; bluebird bio: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees; Sanofi-Genzyme: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding; Medac: Consultancy, Honoraria; Novartis: Honoraria, Membership on an entity's Board of Directors or advisory committees; Jazz Pharmaceuticals: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees; Gilead: Honoraria; AbbVie Pharmacyclics: Membership on an entity's Board of Directors or advisory committees; Orchard: Consultancy, Honoraria; Incyte: Consultancy, Membership on an entity's Board of Directors or advisory committees. Baruchel:Jazz Pharmaceuticals: Consultancy, Honoraria; Celgene Corporation: Consultancy, Honoraria; Astra Zeneca: Consultancy; Servier: Consultancy, Honoraria; Novartis: Consultancy, Honoraria; Shire: Research Funding; Bellicum: Consultancy.

Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 42-42 ◽  
Author(s):  
Stefan O. Ciurea ◽  
Ankur Varma ◽  
Piyanuch Kongtim ◽  
Samer Srour ◽  
Qaiser Bashir ◽  
...  

Introduction Allogeneic hematopoietic stem cell transplantation (AHSCT) is increasingly performed for older patients with AML; however, the optimal conditioning regimen for these patients remains unclear. Methods: We retrospectively evaluated outcomes of 404 patients with AML, ≥60 years receiving AHSCT at our institution between 01/2005-08/2018 who received 4 conditioning regimens: 1) fludarabine+melphalan 100mg/m2 (FM100, N=78), 2) fludarabine+melphalan 140mg/m2 (FM140, N=89), 3) fludarabine+IV busulfan x 4 days with Bu AUC≥5,000/day (equivalent dose 130mg/m2/day) (Bu≥5,000, N=131), 4) fludarabine+IV busulfan x 4 days with Bu AUC 4,000/day (equivalent dose 110mg/m2/day) (Bu4,000, N=106). To adjust for potential selection bias in choices of conditioning regimen, propensity score was calculated and used as a stratifying variable in a multivariable Cox regression model. Factors included in the propensity score calculation were age, secondary AML, ELN2017 genetic risk, remission status before transplant, induction failure, donor type, stem cell source and KPS. Results are presented for the FM100, FM140, Bu≥5,000 and Bu4000, respectively. Median follow-up survivors were 40, 74, 30 and 44 months, respectively (p=0.06). Donors are matched sibling, matched unrelated, haploidentical and mismatched unrelated donor in 126 (31%), 218 (54%), 40 (10%) and 20 (5%) patients, respectively. Patients in the FM100 group were significantly older and had lower KPS. The median age was 67, 64, 64 and 65 years, respectively (p=0.001), while 51%, 32%, 27% and 27% had KPS<90%, respectively (p<0.001). The HCT-CI of ≥3 was present in 57%, 62%, 56% and 70%, respectively (p=0.33), while 42%, 78%, 47% and 51% had high and very high-risk DRI, respectively (p<0.001), and 12%, 46%, 18% and 32% of the patients were transplanted in active disease (p<0.001). No significant differences were seen in both cytogenetic and ELN2017 genetic risk. More patients in FM100 group were treated using a standard of care protocol (73%, 64%, 25% and 31%, respectively, p<0.001). Grade 2-4 aGVHD at day 100 were 26% vs. 26%, 36% and 40% (p=0.04), and extensive cGVHD at 3 years 14% vs. 42%, 36% and 37%, respectively (p=0.07). The NRM at 3 years were 19%, 29%, 25% and 21% (p=0.06), and 3-year relapse rates were 32% vs. 32%, 30% and 55%, respectively (p=0.003). Among 4 groups, FM100 group had a significantly better PFS and GRFS with 5-year PFS for these 4 groups were 44%, 30%, 33% and 22% (p=0.02) and 5-year GRFS were 28%, 20%, 18% and 9% (p=0.006), respectively (Figure 1). For subgroup of patients with KPS <90%, 5-year PFS were 41%, 27%, 28%, 22%, respectively (p=0.007), while there was no significant difference between 4 conditioning groups in patients with high-risk AML defined as either secondary AML, induction failure or high-risk cytogenetics/high ELN2017 risk, suggesting that a more intense conditioning is not beneficial in this group of patients. The survival benefit of FM100 persisted after adjusted for baseline factors, transplant characteristics as well as propensity scores in a multivariable analysis (MVA). In MVA for PFS, HR was 0.57 (p=0.013) for FM100, 0.68 (p=0.056) for FM140 and 0.77 (p=0.137) for Bu> 5000 as compared with Bu 4,000 group (Figure 1). In the MVA for GRFS, HR for FM100, FM140 and Bu> 5000 was 0.53 (p=0.005), 0.78 (p=0.196), and 0.81 (p=0.178), respectively as compared with Bu 4,000 group. Other factors that independently predicted PFS were secondary AML (HR 1.68, p=0.001), remission status before transplant (HR 1.82, p=0.048 for CR with MRD positive, HR 1.87, p=0.043 for CR with unknown MRD status and HR 2.86, p=0.001 for active disease at transplant as compared with CR with MRD negative), KPS (HR 0.98, p=0.005) and use of a mismatched unrelated donor (HR 2.46, p=0.001 compared with matched related donor transplant). Conclusions: Older patients with AML benefit from a reduced-intensity conditioning with FM100 conditioning regimen, which was associated with better survival despite the fact that patients who could not receive more intense conditioning preferentially received this regimen. Higher intensity conditioning does not appear to improve survival in older patients. Alternative approaches to increase in conditioning intensity are needed to improve survival in patients with AML receiving allogeneic hematopoietic stem cell transplantation. Disclosures Ciurea: Kiadis Pharma: Membership on an entity's Board of Directors or advisory committees, Other: stock holder; Miltenyi: Research Funding; Spectrum: Membership on an entity's Board of Directors or advisory committees; MolMed: Membership on an entity's Board of Directors or advisory committees. Bashir:Imbrium: Membership on an entity's Board of Directors or advisory committees; Celgene: Research Funding; Takeda: Membership on an entity's Board of Directors or advisory committees, Research Funding; Kite: Membership on an entity's Board of Directors or advisory committees; Acrotech: Research Funding; StemLine: Research Funding; Spectrum: Membership on an entity's Board of Directors or advisory committees; Amgen: Membership on an entity's Board of Directors or advisory committees. Oran:Astex pharmaceuticals: Research Funding; AROG pharmaceuticals: Research Funding. Popat:Bayer: Research Funding; Incyte: Research Funding; Jazz: Consultancy. Konopleva:Stemline Therapeutics: Consultancy, Honoraria, Research Funding; Reata Pharmaceuticals: Equity Ownership, Patents & Royalties; Ablynx: Research Funding; Astra Zeneca: Research Funding; Agios: Research Funding; Ascentage: Research Funding; Calithera: Research Funding; Forty-Seven: Consultancy, Honoraria; Kisoji: Consultancy, Honoraria; Eli Lilly: Research Funding; AbbVie: Consultancy, Honoraria, Research Funding; Cellectis: Research Funding; Amgen: Consultancy, Honoraria; F. Hoffman La-Roche: Consultancy, Honoraria, Research Funding; Genentech: Honoraria, Research Funding.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 1916-1916
Author(s):  
Daniel Cancilla ◽  
Haresh Thakellapalli ◽  
Marvin J Meyers ◽  
Michael P. Rettig ◽  
Ezhilarasi Chendamarai ◽  
...  

Background: Hematopoietic stem cell (HSC) transplant is an essential treatment for a variety of blood disorders and malignancies. A key step in this procedure is the mobilization of donor stem cells. The most commonly used regimen for donor mobilization is a 5-day course of G-CSF. The length of this regimen coupled with the associated side effects emphasizes a need for superior alternatives. In recent years, there has been a growing understanding of mechanisms governing stem cell retention within the bone marrow niche. This has led to the development of new mobilization drugs that specifically target these processes. Two examples of previously described drugs that target mechanisms of stem cell retention are Plerixafor (a CXCR4 inhibitor already in clinical use), and truncated Gro-Beta (tGroβ; a CXCR2 agonist). Another potential target for inducing mobilization is disruption of the interaction between the VLA-4 integrin and its ligand VCAM-1. In this study, we evaluate the efficacy of novel VLA-4 inhibitors (VLA4i) alone and in combination with Plerixafor and/or tGroβ for the purposes of hematopoietic stem cell mobilization. Methods: We synthesized over 15 novel VLA-4 inhibitor molecules and tested their potency using soluble VCAM-1 binding assays. The 5 inhibitors determined to be most potent were then tested in vivo in DBA mice for their ability to mobilize HSCs alone and in combination with tGroβ and/or Plerixafor (n=5). HSC mobilization was measured in wild-type and splenectomized mice via flow cytometry to quantify the proportion of LSK (Lineage- Sca+ cKit+) cells as well as via Colony Forming Unit (CFU) assays. For competitive transplant, mobilized CD45.1+ BALB/c mouse blood (10 uL) was injected into lethally irradiated CD45.2+ BALB/c recipients alongside 2.5x105 CD45.2+ BALB/c bone marrow cells (n=10 / cohort). HSC engraftment was monitored monthly via flow cytometry for ratio of 45.1+ vs. 45.2+ cells in peripheral blood. Results: Firetagrast and BIO5192 are previously characterized VLA4i that have been administered to humans for indications unrelated to HSC mobilization. Our best VLA4i to date, LGB-2019, exhibited similar potency as BIO5192 in preventing the binding of sVCAM-1 to VLA-4 (IC50: 1.7nM) and was >200-fold more potent than firategrast. LGB-2019 showed increased aqueous solubility and mobilized 1.5-fold more murine LSK cells for a longer time period (peak HSC mobilization maintained for 4 hours) than BIO5192 when administered alone. Simultaneous injection of C57BL/6 mice with LGB-2019 (VLA4i), Plerixafor (CXCR4i) and tGro-β (CXCR2a) resulted in a synergistic increase in circulating CFUs (Fig. 1A; 9.8 x 103 CFUs/mL) and LSKs (Fig. 1B; 12.8 LSKs/uL) at 4 hours post-injection. In contrast, 5 days of G-CSF treatment mobilized approximately 3-fold and 8-fold less CFUs and LSKs, respectively (Fig. 1A-B). We saw no significant difference in mobilization for splenectomized vs. wildtype mice (23.4 x 103 CFUs/mL vs. 23.0 x 103 CFUs/mL) when mobilizing DBA/2 mice via VLA4i+CXCR4i+CXCR2a. Three months after competitive transplantation, blood obtained from BALB/c mice mobilized with the triple combination engrafted significantly better than blood obtained from mice treated with G-CSF or the dual combinations (Fig. 1C). Summary: New insights about the stem cell niche have allowed for the development of targeted drugs for the purposes of mobilization. Here, we show that a novel VLA-4 receptor inhibitor in combination with two other known mobilizers induces mobilization of hematopoietic stem and progenitor cells (CFU/LSK) at levels superior to the standard of care G-CSF and in a dramatically shortened time frame. Mouse transplant data also show superior engraftment in lethally irradiated recipients when using the triple cocktail regimen compared to the G-CSF mobilized graft. Secondary transplants are ongoing and will provide a more complete picture of primitive HSC mobilization and serial engraftment properties of the cells. Disclosures Rettig: WashU: Patents & Royalties: Patent Application 16/401,950. Karpova:WashU: Patents & Royalties: Patent Application 16/401,950. Ruminski:WahU: Patents & Royalties: Patent Application 16/401,950. Morrow:Magenta Therapeutics: Employment, Equity Ownership, Patents & Royalties. DiPersio:Cellworks Group, Inc.: Membership on an entity's Board of Directors or advisory committees; RiverVest Venture Partners Arch Oncology: Consultancy, Membership on an entity's Board of Directors or advisory committees; Magenta Therapeutics: Equity Ownership; Incyte: Consultancy, Research Funding; Bioline Rx: Research Funding, Speakers Bureau; Macrogenics: Research Funding, Speakers Bureau; Karyopharm Therapeutics: Consultancy; Celgene: Consultancy; Amphivena Therapeutics: Consultancy, Research Funding; WUGEN: Equity Ownership, Patents & Royalties, Research Funding; NeoImmune Tech: Research Funding.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 3472-3472 ◽  
Author(s):  
Erden Atilla ◽  
Pervin Topcuoglu ◽  
Erman Akkus ◽  
Pinar Ataca Atilla ◽  
Sinem Civriz Bozdag ◽  
...  

Abstract Introduction: Allogeneic hematopoietic stem cell transplantation (allo-HSCT) is widely used to treat malignant and non-malignant hematological diseases. The impact of ABO mismatch on outcome following transplantation remains controversial. In this study, our aim is to define effects of ABO mismatch on engraftment, graft vs host disease, relapse free survival (RFS) and overall survival (OS) in patients who underwent allo-HSCT. Patients and Methods: Between 1988 and 2016, we retrospectively identified 1016 patients who underwent allo-HSCT at Ankara University School of Medicine, Department of Hematology. Chi-square and Fisher's exact tests were used where appropriate in comparison. Cox regression model and Kaplan Meier curves were applied for survival analysis. P<0.05 was considered as statistically significant. Results: The median follow-up period was 34.7 months (range, 0.2-229). In our cohort, there were 420 (41.3%) ABO-mismatched transplants occurred including 167 (16.4%) major, 197 (19.4%) minor and 55 (5.4%) bidirectional mismatches. The pre-transplant characteristics of patients are summarized in table. Allo-HSCTs from unrelated donors and peripheral blood grafts were detected higher in ABO mismatched patients vs ABO matched patients (28% vs 11%, P<0.0001; 78% vs 67%, P<0.0001). The engraftment failure was higher in ABO mismatch group compared to ABO matched group (67 (16%) vs 70 (11%), P=0.05). Neutrophil and platelet engraftment rates were not statistically different in major, minor or bidirectional ABO mismatched vs matched donors. The acute graft vs host disease (GVHD) and chronic GVHD incidences did not alter in patients with ABO match and mismatch (44% vs 45%, P=0.78; 41% vs 39%,P=0.81). In ABO-mismatched group, hemolysis after infusion of graft occurred in 50 patients (12%) whereas during engraftment in 35 patients (8%). Although not statistically significant, hemolysis were occurred higher in major ABO mismatch. Plasma exchanges were performed in 18 patients in the major ABO mismatched group due to high anti-donor type isoagglutinin titers (≥1/128). Pure red cell aplasia was diagnosed in 5 (3%) major ABO mismatched patients. Major ABO mismatch (HR:1.46, 95% Cl:1.06-2.03;P=0.022) was found to be related with lower RFS and OS (HR:1.31, 95% Cl:1.06-1.62;P=0.013). 3-year OS and 1-year RFS were lower with major ABO mismatch (38% vs 47%, P=0.02; 15% vs 24%; P=0.02) (Figure). Conclusion: Engraftment failure was detected higher in patients with ABO mismatch as well as major ABO mismatch was related with lower RFS and OS although the cohort is heterogeneous. Close monitoring and early treatment strategies for expectable complications would reduce the number fatal events by ABO mismatched allo-HSCT. Disclosures Civriz Bozdag: NOVARTIS: Consultancy; MSD: Research Funding; TAKEDA: Consultancy. Özcan:Abbvie: Other: Travel payment; Bayer: Research Funding; MSD: Other: travel support, Research Funding; Roche: Honoraria, Research Funding; Janssen: Other: Travel Support, Research Funding; BMS: Honoraria; Novartis: Research Funding; Jazz: Other; Jazz: Other: Travel support; Archigen: Research Funding; Celgene: Other: Travel support, Research Funding; MSD: Research Funding; Takeda: Honoraria, Other: Travel payment, Research Funding. Beksac:Deva: Honoraria, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Janssen Cilag: Honoraria, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Takeda: 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; Bristol Myers Squibb: Honoraria, 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, Speakers Bureau. Ilhan:Roche: Speakers Bureau; BMS: Speakers Bureau; Celgene: Speakers Bureau; Alexion: Speakers Bureau.


Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 10-10
Author(s):  
Rajni Agarwal ◽  
Kenneth I. Weinberg ◽  
Hye-Sook Kwon ◽  
Anne Le ◽  
Janel R Long-Boyle ◽  
...  

Successful hematopoietic stem cell transplantation (HSCT) requires vacating recipient hematopoietic stem cell (HSC) niches in the bone marrow to permit donor HSC engraftment that can provide life-long hematopoietic and immune function. Currently, HSCT in SCID relies on DNA damaging chemotherapy to eliminate recipient HSC and achieve niche clearance. We have pursued a non-toxic approach to target and deplete HSC using a humanized monoclonal antibody, JSP191, that binds human CD117 (c-Kit). We previously showed the safety and successful HSC engraftment in a Phase 1 trial of the first 6 patients with severe combined immunodeficiency (SCID), who underwent a second transplant because of HSC engraftment failure and poor immunity after their first transplantation. In these re-transplant patients even a low level of stringently measured myeloid chimerism resulted in significant and sustained generation of naive T cells and clinical improvement. Based on these results, the study of JSP191 (NCT#02963064)has opened a cohort of newly diagnosed infants with SCID. Here we report data from the first patient in this cohort, a SCIDX1 patient who received a primary HSCT with haploidentical CD34+ cells after conditioning with JSP 191. The patient had a c.270-15A&gt;G variant in the IL2RG gene, which is predicted to cause a null phenotype. Besides a T- B+ NK- phenotype typical of SCIDX1 including dysfunctional B cells, the patient had anemia and intermittent neutropenia and thrombocytopenia. Despite evidence of maternal T cell engraftment, the patient had no clinical graft-versus-host disease (GVHD). The patient was initially enrolled in a trial of lentiviral gene therapy, but harvested bone marrow cells died in vitro during transduction and culture. The patient also mobilized poorly with G-CSF/Plerixafor. Further investigation revealed heterozygosity for loss-of-function mutations in two genes involved in DNA repair, BRCA1 and RAD51; Diepoxybutane (DEB) breakage study showed greater than normal pathologic chromosomal breaks, but less than that seen in Fanconi anemia. Because of concern for possible hypersensitivity to alkylating agent-based conditioning, the patient was referred for transplant with JSP191 conditioning. The patient received a CD34+ peripheral blood HSCT from his father after conditioning with 0.3 mg/kg of JSP 191 antibody intravenously over an hour on Day -8 and rATG (Thymoglobulin) on Day -5, -4, -3 and -2 (3.5 mg/kg total) to prevent rejection by the maternal T cells. The cryopreserved donor CD34+ cells were administered after sufficient clearance of the JSP191 serum level. The antibody infusion was well tolerated without toxicity, and the post-transplant course was uneventful without acute toxicities or GVHD. As a surrogate marker for HSC engraftment, CD15+ myeloid cells from peripheral blood were stringently sorted by flow cytometry and donor levels were quantified by short-tandem repeat (STR) analysis. Progressive levels of myeloid engraftment were observed beginning at Week 4. The level of donor chimerism at 12 weeks was 8% in the sorted CD15+ blood cells, and a marrow aspirate showed 25% donor CD34+ cells. By 3 months pre-existing abnormal CD19-CD20+ host B lymphocytes were significantly reduced, and CD19+ donor-derived B lymphocytes were emerging. At 2 months, CD4+ recent thymic emigrant and naïve T lymphocytes were observed, and by 3 months, overall T and NK lymphocyte numbers were 390/uL and 117/uL, respectively. Normal blastogenic responses to the T cell mitogen PHA were observed at 3 months. These first-in-class results provide proof of concept of the safety and efficacy of the use of JSP191 antibody to clear host marrow niche space to enable sufficient donor HSC engraftment and immune reconstitution as primary therapy of SCID. Non-genotoxic conditioning with JSP191 may replace conventional conditioning for newly diagnosed infants with SCID, thereby avoiding toxicities of chemotherapy. Disclosures Kohn: Allogene Therapeutics: Consultancy, Membership on an entity's Board of Directors or advisory committees; Orchard Therapeutics: Consultancy, Patents & Royalties, Research Funding. De Oliveira:Orchard Therapeutics: Research Funding; bluebird bio, Inc.: Research Funding. Czechowicz:Rocket Pharmaceuticals, Inc.: Research Funding. Brown:Merck: Membership on an entity's Board of Directors or advisory committees; Ansun: Membership on an entity's Board of Directors or advisory committees; Cidara: Membership on an entity's Board of Directors or advisory committees; Allogene: Membership on an entity's Board of Directors or advisory committees; Cellerant Therapeutics: Membership on an entity's Board of Directors or advisory committees. Shizuru:Jasper Therapeutics, Inc: Current equity holder in private company, Membership on an entity's Board of Directors or advisory committees.


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 3487-3487
Author(s):  
Paul Harker-Murray ◽  
Brent L. Wood ◽  
Meenakshi Devidas ◽  
Zhiguo Chen ◽  
Tal Schechter-Finkelstein ◽  
...  

Abstract Background: The prognosis for patients (pts) with relapsed T-ALL and T-LL is dismal. The primary goal of T-ALL/T-LL treatment is to prevent relapse. In the phase 3 Children's Oncology Group (COG) clinical trial AALL1231 (NCT02112916), children, adolescents and young adults (age 1-30 years) with T-ALL and T-LL were treated with a modified augmented BFM (aBFM) backbone that used dexamethasone as the only corticosteroid and included two (rather than one) doses of pegaspargase during induction and delayed intensification. Pts were stratified as standard (SR), intermediate (IR), or very high risk (VHR), primarily based on disease response: morphology, minimal residual disease (MRD) performed by multiparameter flow cytometry at a central reference laboratory) at end of induction and consolidation (T-ALL), and radiographic response for T-LL. Pts were randomized 1:1 to receive/not receive bortezomib during induction and delayed intensification (1.3mg/m 2 x 4 doses per block). VHR T-ALL pts were defined as having day 29 M3 marrow (&gt;25% blasts) or end of consolidation (EOC) MRD &gt;0.1%. 10-15% of T-ALL pts were predicted to be VHR based on COG AALL0434. Pts with induction failure (M3 marrow by morphology) or EOC MRD &gt;0.1% were expected to have 4-yr event-free survival (EFS) of ~66+/-16%. Following consolidation, VHR pts received 3 BFM-based intensification blocks in lieu of interim maintenance (IM). Detectable MRD following the intensification blocks was considered an event and these pts were removed from protocol therapy. VHR ALL pts who had undetectable MRD continued protocol therapy, received delayed intensification, an IM phase with Capizzi escalating methotrexate plus pegaspargase, and maintenance. A secondary aim of AALL1231 was to compare survival in VHR T-ALL pts with EOC MRD ≥ 0.1% but undetectable MRD after intensification of chemotherapy with those who continued to have detectable MRD and were eligible for other treatment strategies, including hematopoietic stem cell transplant (HSCT). This study also analyzed outcomes for pts with M3 marrow at the end of induction. Results: AALL1231 accrued 847 pts (824 eligible and evaluable) of 1400 anticipated from 2014 until early closure. The 3-year EFS for the bortezomib randomization for the SR and IR groups has been reported previously (Teachey, et. al ASH 2020). Because only 2 of 209 T-LL pts were VHR; this report focuses on the outcomes of the 5.2% (32/615) of T-ALL pts who were VHR. In total, 25 VHR T-ALL pts were EOC MRD &gt;0.1%, and 18 of these had MRD sent at the end of HR intensification. Of the 8 pts who became MRD undetectable and continued protocol therapy, only 2 survived (3-year overall survival [OS] 25+15.3%). In contrast, 10 pts who had detectable MRD were taken off protocol and underwent HSCT. Of these 10, only one relapsed (3-year OS 90+12.7%). The 3-year OS for the 10 pts who were M3 at Day 29 was 60.0±17.0%. As there were not enough pts to assess the impact of EOC MRD on pts who were M3 at Day 29, we assessed the impact of EOC MRD on outcomes in M2 (5-25% blasts at Day 29; n = 24) and M3 pts, which defines induction failure in other cooperative groups. M2+M3 T-ALL who were EOC MRD &lt;0.1% (n = 15) had 3-year OS of 86.7±10.0% vs 45.5±15.0% for those with EOC MRD &gt;0.1% (n = 12) pts. Conclusions: T-ALL pts treated on AALL1231 who are EOC MRD ≥0.1% with undetectable MRD after 3 BFM-based intensification blocks had a very poor outcome when treated with standard cytotoxic chemotherapy. In contrast, while patient numbers are small, those pts that remained MRD-positive after 3 intensification blocks and underwent HSCT had an excellent outcome. These data not only impact the recommended treatment for T-ALL pts who are induction and consolidation failures, but also support the importance of the graft-versus-leukemia (GVL) effect in refractory T-ALL. Disclosures Hayashi: Magenta Therapeutics: Consultancy. August: Jazz: Membership on an entity's Board of Directors or advisory committees. Hermiston: Sobi: Consultancy; Novartis: Consultancy. Bollard: Cabeletta Bio: Membership on an entity's Board of Directors or advisory committees; Catamaran Bio and Mana Therapeutics: Other: member and cofounder; SOBI: Other: DSMB. Loh: MediSix therapeutics: Membership on an entity's Board of Directors or advisory committees. Raetz: Pfizer: Research Funding; Celgene: Other: DSMB member. Teachey: BEAM Therapeutics: Consultancy, Research Funding; NeoImmune Tech: Research Funding; Sobi: Consultancy; Janssen: Consultancy.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 3764-3764
Author(s):  
Teresa Ezponda ◽  
Juan P. Romero ◽  
Marina Ainciburu ◽  
Ana Alfonso ◽  
Nerea Berastegui ◽  
...  

Myelodysplastic syndromes (MDS) are clonal hematopoietic stem cell (HSC) malignancies characterized by ineffective hematopoiesis. Genetic alterations do not fully explain the molecular pathogenesis of the disease, indicating that other types of lesions, such as transcriptional aberrations, may play a role in its development. Moreover, MDS prevalence is almost exclusive to older patients, suggesting that elderly-related alterations may predispose to the development of this clinical entity. Thus, study of the transcriptional lesions occurring in the aging-MDS axis could shed some light of the molecular bases of the disease. To characterize the transcriptional profile of HSCs in aging and MDS, we isolated CD34+, CD38-, CD90+, CD45RA- cells from 11 untreated MDS patients with unilineage and multilineage dysplasia (median of 75 y/o), as well as from 16 young and 8 elderly healthy donors (median of 21 and 70 y/o, respectively), and their expression profile was analyzed using MARS-seq. Unsupervised principal component analysis demonstrated that the three groups of HSCs clustered separately, indicating that different expression profiles characterize healthy young and elderly, and MDS-associated HSCs. To better understand the gene expression deregulation of HSCs, we analyzed the transcriptional dynamisms along the aging-MDS axis, detecting groups of genes following different patterns of expression. Some gene clusters showed exclusive alteration either in aging or in the progression from elderly HSCs to MDS-HSCs, other groups of genes presented a continuous alteration along the axis, and some displayed opposite regulation in aging and in the transition to MDS (Figure 1). Genes showing specific downregulation in aging were involved in DNA damage sensing and repair, and in cell cycle regulation, whereas genes overexpressed in this process were enriched in apoptosis regulators and in cancer-associated genes, including AML-related factors. These findings indicate that transcriptional changes in aging may predispose for MDS and AML, and potentially other malignancies. Interestingly, we detected a group of genes in which the age-mediated upregulation of gene expression was reversed to that of young HSCs in MDS, indicating a "rejuvenation" profile of malignant HSCs. These genes were involved in response to inflammation, to different types of stress conditions such as hypoxia or radiation, and to cytokines. Elderly HSCs may upregulate such genes in response to the known inflammatory microenvironment of elderly bone marrow. Intriguingly, the decrease in expression detected in MDS suggests that malignant HSCs lose the ability of reacting to such stimuli, possibly favoring their survival in a hostile microenvironment. Finally, the analyses performed allowed for the identification of genes showing MDS-specific deregulation. Genes specifically overexpressed in MDS compared to normal (both young and elderly) HSCs, we enriched in transcriptional and epigenetic regulators, and among them, we detected the presence of DDIT3/CHOP, a member of the CCAAT/enhancer-binding protein (C/EBP) family of transcription factors. To determine its potential effects on hematopoietic deregulation, DDIT3 was exogenously overexpressed in healthy HSCs. Notably, its upregulation produced an erythroid bias in an ex-vivo differentiation system, with an increase in the percentage of erythroblasts and a decrease in granulocytes and monocytes compared to HSCs transduced with the empty vector. Transcriptomic analysis of transduced HSCs not subjected to differentiation demonstrated how DDIT3 overexpression produced an erythroid-prone state of HSCs, suggesting it may act as a pioneer factor in MDS-HSCs. Furthermore, gene set enrichment analysis showed that DDIT3 overexpression produced an MDS-like transcriptional profile, suggesting this factor may be key in the acquisition of the disease. Altogether, our results demonstrate that HSCs undergo transcriptional changes in the aging-MDS axis that may alter their intrinsic functions as well as their response to the microenvironment, ultimately contributing to the acquisition of the disease. In particular, our data show that DDIT3 may be a potential driver of MDS transformation. Disclosures Paiva: Amgen, Bristol-Myers Squibb, Celgene, Janssen, Merck, Novartis, Roche, and Sanofi; unrestricted grants from Celgene, EngMab, Sanofi, and Takeda; and consultancy for Celgene, Janssen, and Sanofi: Consultancy, Honoraria, Research Funding, Speakers Bureau. Díez-Campelo:Celgene Corporation: Consultancy, Membership on an entity's Board of Directors or advisory committees, Research Funding; Novartis: Consultancy, Membership on an entity's Board of Directors or advisory committees, Research Funding.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 1962-1962
Author(s):  
Sandhya R. Panch ◽  
Brent R. Logan ◽  
Jennifer A. Sees ◽  
Bipin N. Savani ◽  
Nirali N. Shah ◽  
...  

Introduction: Approximately 7% of unrelated hematopoietic stem cell (HSC) donors are asked to donate a subsequent time to the same or different recipient. In a recent large CIBMTR study of second time donors, Stroncek et al. incidentally found that second peripheral blood stem cell (PBSC) collections had lower total CD34+ cells, CD34+ cells per liter of whole blood processed, and CD34+ cells per kg donor weight. Based on smaller studies, the time between the two independent PBSC donations (inter-donation interval) as well as donor sex, race and baseline lymphocyte counts appear to influence CD34+ cell yields at subsequent donations. Our objective was to retrospectively evaluate factors contributory to CD34+ cell yields at subsequent PBSC donation amongst NMDP donors. Methods. The study population consisted of filgrastim (G-CSF) mobilized PBSC donors through the NMDP/CIBMTR between 2006 and 2017, with a subsequent donation of the same product. evaluated the impact of inter-donation interval, donor demographics (age, BMI, race, sex, G-CSF dose, year of procedure, need for central line) and changes in complete blood counts (CBC), on the CD34+ cell yields/liter (x106/L) of blood processed at second donation and pre-apheresis (Day 5) peripheral blood CD34+ cell counts/liter (x106/L) at second donation. Linear regression was used to model log cell yields as a function of donor and collection related variables, time between donations, and changes in baseline values from first to second donation. Stepwise model building, along with interactions among significant variables were assessed. The Pearson chi-square test or the Kruskal-Wallis test compared discrete variables or continuous variables, respectively. For multivariate analysis, a significance level of 0.01 was used due to the large number of variables considered. Results: Among 513 PBSC donors who subsequently donated a second PBSC product, clinically relevant decreases in values at the second donation were observed in pre-apheresis CD34+ cells (73.9 vs. 68.6; p=0.03), CD34+cells/L blood processed (32.2 vs. 30.1; p=0.06), and total final CD34+ cell count (x106) (608 vs. 556; p=0.02). Median time interval between first and second PBSC donations was 11.7 months (range: 0.3-128.1). Using the median pre-apheresis peripheral blood CD34+ cell counts from donation 1 as the cut-off for high versus low mobilizers, we found that individuals who were likely to be high or low mobilizers at first donation were also likely to be high or low mobilizers at second donation, respectively (Table 1). This was independent of the inter-donation interval. In multivariate analyses, those with an inter-donation interval of >12 months, demonstrated higher CD34+cells/L blood processed compared to donors donating within a year (mean ratio 1.15, p<0.0001). Change in donor BMI was also a predictor for PBSC yields. If donor BMI decreased at second donation, so did the CD34+cells/L blood processed (0.74, p <0.0001). An average G-CSF dose above 960mcg was also associated with an increase in CD34+cells/L blood processed compared to donors who received less than 960mcg (1.04, p=0.005). (Table 2A). Pre-apheresis peripheral blood CD34+ cells on Day 5 of second donation were also affected by the inter-donation interval, with higher cell counts associated with a longer time interval (>12 months) between donations (1.23, p<0.0001). Further, independent of the inter-donation interval, GCSF doses greater than 960mcg per day associated with higher pre-apheresis CD34+ cells at second donation (1.26, p<0.0001); as was a higher baseline WBC count (>6.9) (1.3, p<0.0001) (Table 2B). Conclusions: In this large retrospective study of second time unrelated PBSC donors, a longer inter-donation interval was confirmed to be associated with better PBSC mobilization and collection. Given hematopoietic stem cell cycling times of 9-12 months in humans, where possible, repeat donors may be chosen based on these intervals to optimize PBSC yields. Changes in BMI are also to be considered while recruiting repeat donors. Some of these parameters may be improved marginally by increasing G-CSF dose within permissible limits. In most instances, however, sub-optimal mobilizers at first donation appear to donate suboptimal numbers of HSC at their subsequent donation. Disclosures Pulsipher: CSL Behring: Membership on an entity's Board of Directors or advisory committees; Miltenyi: Research Funding; Bellicum: Consultancy; Amgen: Other: Lecture; Jazz: Other: Education for employees; Adaptive: Membership on an entity's Board of Directors or advisory committees, Research Funding; Novartis: Consultancy, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Medac: Honoraria. Shaw:Therakos: Other: Speaker Engagement.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 3199-3199
Author(s):  
Ji Zha ◽  
Lori Kunselman ◽  
Hongbo Michael Xie ◽  
Brian Ennis ◽  
Jian-Meng Fan ◽  
...  

Hematopoietic stem cell (HSC) transplantation (HSCT) is required for curative therapy for patients with high-risk hematologic malignancies, and a number of non-malignant disorders including inherited bone marrow failure syndromes (iBMFS). Strategies to enhance bone marrow (BM) niche capacity to engraft donor HSC have the potential to improve HSCT outcome by decreasing graft failure rates and enabling reduction in conditioning intensity and regimen-associated complications. Several studies in animal models of iBMFS have demonstrated that BM niche dysfunction contributes to both the pathogenesis of iBMFS, as well as impaired graft function after HSCT. We hypothesize that such iBMFS mouse models are useful tools for discovering targetable niche elements critical for donor engraftment after HSCT. Here, we report the development of a novel mouse model of Shwachman-Diamond Syndrome (SDS) driven by conditional Sbds deletion, which demonstrates profound impairment of healthy donor hematopoietic engraftment after HSCT due to pathway-specific dysfunctional signaling within SBDS-deficient recipient niches. We first attempted to delete Sbds specifically in mature osteoblasts by crossing Sbdsfl/flmice with Col1a1Cre+mice. However, the Col1a1CreSbdsExc progenies are embryonic lethal at E12-E15 stage due to developmental musculoskeletal abnormalities. Alternatively, we generated an inducible SDS mouse model by crossing Sbdsfl/flmice with Mx1Cre+ mice, and inducing Sbds deletion in Mx1-inducible BM hematopoietic and osteolineage niche cells by polyinosinic-polycytidilic acid (pIpC) administration. Compared with Sbdsfl/flcontrols, Mx1CreSbdsExc mice develop significantly decreased platelet counts, an inverted peripheral blood myeloid/lymphoid cell ratio, and reduced long-term HSC within BM, consistent with stress hematopoiesis seen in BMF and myelodysplastic syndromes. To assess whether inducible SBDS deficiency impacts niche function to engraft donor HSC, we transplanted GFP+ wildtype donor BM into pIpC-treated Mx1CreSbdsExc mice and Sbdsfl/flcontrols after 1100 cGy of total body irradiation (TBI). Following transplantation, Mx1CreSbdsExc recipient mice exhibit significantly higher mortality than controls (Figure 1). The decreased survival was related to primary graft failure, as Mx1CreSbdsExc mice exhibit persistent BM aplasia after HSCT and decreased GFP+ reconstitution in competitive secondary transplantation assays. We next sought to identify the molecular and cellular defects within BM niche cells that contribute to the engraftment deficits in SBDS-deficient mice. We performed RNA-seq analysis on the BM stromal cells from irradiated Mx1CreSbdsExc mice versus controls, and the results revealed that SBDS deficiency in BM niche cells caused disrupted gene expression within osteoclast differentiation, FcγR-mediated phagocytosis, and VEGF signaling pathways. Multiplex ELISA assays showed that the BM niche of irradiated Mx1CreSbdsExc mice expresses lower levels of CXCL12, P-selectin and IGF-1, along with higher levels of G-CSF, CCL3, osteopontin and CCL9 than controls. Together, these results suggest that poor donor HSC engraftment in SBDS-deficient mice is likely caused by alterations in niche-mediated donor HSC homing/retention, bone metabolism, host monocyte survival, signaling within IGF-1 and VEGF pathways, and an increased inflammatory state within BM niches. Moreover, flow cytometry analysis showed that compared to controls, the BM niche of irradiated Mx1CreSbdsExc mice contained far fewer megakaryocytes, a hematopoietic cell component of BM niches that we previously demonstrated to be critical in promoting osteoblastic niche expansion and donor HSC engraftment. Taken together, our data demonstrated that SBDS deficiency in BM niches results in reduced capacity to engraft donor HSC. We have identified multiple molecular and cellular defects in the SBDS-deficient niche contributing to this phenotype. Such niche signaling pathway-specific deficits implicate these pathways as critical for donor engraftment during HSCT, and suggest their potential role as targets of therapeutic approaches to enhance donor engraftment and improve HSCT outcome in any condition for which HSCT is required for cure. Disclosures Olson: Merck: Membership on an entity's Board of Directors or advisory committees; Bluebird Bio: Membership on an entity's Board of Directors or advisory committees; Novartis: Membership on an entity's Board of Directors or advisory committees; Miltenyi: Honoraria.


Blood ◽  
2009 ◽  
Vol 114 (22) ◽  
pp. 3229-3229 ◽  
Author(s):  
Ivana N Micallef ◽  
Eric Jacobsen ◽  
Paul Shaughnessy ◽  
Sachin Marulkar ◽  
Purvi Mody ◽  
...  

Abstract Abstract 3229 Poster Board III-166 Introduction Low platelet count prior to mobilization is a significant predictive factor for mobilization failure in patients with non-Hodgkin's lymphoma (NHL) or Hodgkin's disease (HD) undergoing autologous hematopoietic stem cell (HSC) transplantation (auto-HSCT; Hosing C, et al, Am J Hematol. 2009). The purpose of this study is to assess the efficacy of HSC mobilization with plerixafor plus G-CSF in patients with concomitant thrombocytopenia undergoing auto-HSCT. Methods Patients who had failed successful HSC collection with any mobilization regimen were remobilized with plerixafor plus G-CSF as part of a compassionate use program (CUP). Mobilization failure was defined as the inability to collect 2 ×106 CD34+ cells/kg or inability to achieve a peripheral blood count of ≥10 CD34+ cells/μl without having undergone apheresis. As part of the CUP, G-CSF (10μg/kg) was administered subcutaneously (SC) every morning for 4 days. Plerixafor (0.24 mg/kg SC) was administered in the evening on Day 4, approximately 11 hours prior to the initiation of apheresis the following day. On Day 5, G-CSF was administered and apheresis was initiated. Plerixafor, G-CSF and apheresis were repeated daily until patients collected the minimum of 2 × 106 CD34+ cells/kg for auto-HSCT. Patients in the CUP with available data on pre-mobilization platelet counts were included in this analysis. While patients with a platelet count <85 × 109/L were excluded from the CUP, some patients received waivers and were included in this analysis. Efficacy of remobilization with plerixafor + G-CSF was evaluated in patients with platelet counts ≤ 100 × 109/L or ≤ 150 × 109/L. Results Of the 833 patients in the plerixafor CUP database, pre-mobilization platelet counts were available for 219 patients (NHL=115, MM=66, HD=20 and other=18.). Of these, 92 patients (NHL=49, MM=25, HD=8 and other=10) had pre-mobilization platelet counts ≤ 150 × 109/L; the median platelet count was 115 × 109/L (range, 50-150). The median age was 60 years (range 20-76) and 60.4% of the patients were male. Fifty-nine patients (64.1%) collected ≥2 × 109 CD34+ cells/kg and 13 patients (14.1%) achieved ≥5 × 106 CD34+ cells/kg. The median CD34+ cell yield was 2.56 × 106 CD34+ cells/kg. The proportion of patients proceeding to transplant was 68.5%. The median time to neutrophil and platelet engraftment was 12 days and 22 days, respectively. Similar results were obtained when efficacy of plerixafor + G-CSF was evaluated in 29 patients with platelet counts ≤ 100 × 109/L (NHL=12, MM=10, HD=3 and other=4). The median platelet count in these patients was 83 × 109/L (range, 50-100). The median age was 59 years (range 23-73) and 60.4% of the patients were male. The minimal and optimal cell dose was achieved in 19(65.5%) and 3(10.3%) patients, respectively. The median CD34+ cell yield was 2.92 × 106 CD34+ cells/kg. The proportion of patients proceeding to transplant was 62.1%. The median time to neutrophil and platelet engraftment was 12 days and 23 days, respectively. Conclusions For patients mobilized with G-CSF alone or chemotherapy ±G-CSF, a low platelet count prior to mobilization is a significant predictor of mobilization failure. These data demonstrate that in patients with thrombocytopenia who have failed prior mobilization attempts, remobilization with plerixafor plus G-CSF allows ∼65% of the patients to collect the minimal cell dose to proceed to transplantation. Thus, in patients predicted or proven to be poor mobilizers, addition of plerixafor may increase stem cell yields. Future studies should investigate the efficacy of plerixafor + G-CSF in front line mobilization in patients with low platelet counts prior to mobilization. Disclosures Micallef: Genzyme Corporation: Membership on an entity's Board of Directors or advisory committees, Research Funding. Jacobsen:Genzyme Corporation: Research Funding. Shaughnessy:Genzyme Corporation: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding. Marulkar:Genzyme Corporation: Employment, Equity Ownership. Mody:Genzyme Corporation: Employment, Equity Ownership. van Rhee:Genzyme Corporation: Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding.


Blood ◽  
2015 ◽  
Vol 126 (23) ◽  
pp. 5536-5536
Author(s):  
Yizel Elena Paz Nuñez ◽  
Beatriz Aguado Bueno ◽  
Isabel vicuña Andrés ◽  
Ángela Figuera Álvarez ◽  
Miriam González-Pardo ◽  
...  

Abstract Introduction The prognosis of patients with multiple myeloma (MM) has improved in the last years due to the important advances in the knowledge of the biology of the disease, the implementation of new drugs and the incorporation of autologous hematopoietic stem cell transplant (autoHSCT). The allogenic hematopoietic stem cell transplant (alloHSCT) continues to be controversial: it offers a curative potential but with the cost of high toxicity, limiting the procedure to those young patients with a high-risk disease. This procedure shall be performed in expert centers and, whenever possible, in the context of a clinical trial. In the following we describe the experience of our center with alloHSCT in advance multiple myeloma patients. Patients and methods A total of 18 patients were diagnosed with multiple myeloma received an alloHSCT during a 13 year period (1996-2013), with a median age of 46 ± 5.9 years. All of our patients received an allogenic HLA matched sibling donor with reduced-intensity conditioning. The majority of patients were transplanted because of advanced disease, relapse after an autologous transplant or as part of a sequential transplant in patient with a high risk disease. One patient received, in two occasions, an alloHSCT. Around 70% of patients had received more than 3 previous lines of treatment including, in nearly 95%, an autoHSCT. Patient's characteristics can be found on table 1, characteristics of the procedure can be found in table 2.Table 1.Patient«s CharacteristicsN (%)GenderMale Female10 (55,5%) 9 (44,4%)Secreted ProteinIgGκ IgG λ IgA κ BJ Plasmocitoma8 (44,4%) 4 (22,2%) 2 (11,1%) 3 (16,7%) 1 (5,6%)Debut DS stageII-A II-B III-A III-B Plasmocitoma5 (27,8%) 1 (5,6%) 8 (44,4%) 3 (16,7%) 1 (5,6%)Cytogentics at diagnosisMissing Unfavorable Favorable10 (55,5%) 6 (33,3%) 2 (11,1%)Previous lines of treatment²2 3-4 ³56 (33,3%) 10 (55,5%) 2 (11,1%)Previous autoHSCTYes No17 (94,5%) 1 (5,6%)Previous radiotherapyYes No8 (44,4%) 10 (55,6%)Disease status at transplantComplete remission Partial remission Relapse9 (50,0%) 3 (16,7%) 6 (33,3%)Table 2.Treatment characteristicsN (%)Conditioning regimenMyeloablative Reduced-intensity6 (33,3%) 12 (66.7%)Stem cell sourceBone marrow Peripheral blood4 (22.2%) 14 (77.8%)GVHD prophylaxisCsA+MTXCsA+CSCsA+MMF10 (55.6%) 3 (16.7%) 5 (27.8%)InfectionsYes No16 (88.9%) 2 (11.1%)MucositisYes No12 (66.7%) 6 (33.3%)Acute GVHDYes II-IV III-IV No4 (22.3%) 3 (16.7%) 1 (5.6%) 14 (77.8%)Chronic GVHDNo Limited Extensive8 (44.3%) 5 (27.8%) 5 (27.8%) Results: Transplant related mortality (TRM) before day 100th was one case due to a thromboembolic event. Global TRM was 16.6% (3 cases). The incidence of acute graft versus host disease (aGVHD) was 22%, controlled on most cases when corticosteroids were initiated. More than half of the patients developed chronic graft versus host disease (cGVHD), with an equal distribution on either presentation as limited or extensive. (Table 2) The total number of patients eligible for analysis was 17 (one patient was lost on follow-up). With a median follow up of 11 years, the overall survival (OS) was of 8.06 years [IC 95% 4,33-11,78] (figure 1.) and the estimated progression free survival (PFS) was of 25.83 months [IC 95% 8.87-42.79](figure 2). A total of 5 (29,4%) patients are still alive and 2 (11,7%) of them are in complete remission, of these 1 patient did not have a previous autoHSCT with a follow up of almost 15 years. Conclusions: Our results are similar to those reflected on the literature1-2. However we have to point out that our population is homogenous with advanced MM with more than 3 previous lines of treatment including in most cases auto-HSCT. In spite of this, morbility and mortality in our cohort was acceptable with the limitation of a high rate of cGVHD. There is a need of more studies including more patients to evaluate the role of alloHSCT in the era of new drugs for MM. References 1. Rosi-ol L et al. Allogeneic hematopoietic SCT in multiple myeloma: long-term results from a single institution. Bone Marrow Transplant. 2015. 2. Beaussant Y et al. Hematopoietic Stem Cell Transplantation in Multiple Myeloma: A Retrospective Study of the Société Française de Greffe de Moelle et de Thérapie Cellulaire (SFGM-TC). Biol Blood Marrow Transplant. 2015 Disclosures Alegre: Celgene Corporation: Membership on an entity's Board of Directors or advisory committees; Janssen: Membership on an entity's Board of Directors or advisory committees; Amgen: Membership on an entity's Board of Directors or advisory committees.


Sign in / Sign up

Export Citation Format

Share Document