A CD19/CD3 Bispecific Tandab, AFM11, Recruits T Cells To Potently and Safely Kill CD19+ Tumor Cells

Blood ◽  
2013 ◽  
Vol 122 (21) ◽  
pp. 4405-4405
Author(s):  
Eugene Zhukovsky ◽  
Uwe Reusch ◽  
Carmen Burkhardt ◽  
Stefan Knackmuss ◽  
Ivica Fucek ◽  
...  

Abstract To harness the potent tumor-killing capacity of T cells for the treatment of CD19+ malignancies, we developed a humanized bispecific tetravalent antibody, with two binding sites for CD3 and CD19, the CD19/CD3 RECRUIT-TandAb AFM11. CD19 is expressed from early B cell development through differentiation into plasma cells, and is an attractive alternative to CD20 as a target for the development of therapeutic antibodies to treat B cell malignancies such as Non Hodgkin Lymphoma. Since native antibodies cannot recruit T cells, we engineered a bispecific anti-CD19/anti-CD3 TandAb. The tumor-specific CD19 antigen module targets the TandAb to cancer cells, while simultaneously, the CD3 effector module recruits and activates T cells, leading to cancer cell lysis. The advantages of the TandAb technology, relative to other bi-functional fragment antibody scaffolds, include: improved pharmacokinetics (PK) enabling intravenous dosing, more drug-like properties, and avidity-enhanced efficacy for the targeting and killing of tumor cells. We evaluated in vitro efficacy and safety using CD19+ cell lines, and in vivo efficacy in a murine NOD/scid xenograft model reconstituted with human PBMC. Further, we used standard preclinical IND enabling assays to evaluate tissue cross reactivity, PK, and toxicological profile (local tolerance, hematocompatibility, effects on hematopoesis, etc). In vitro assays demonstrated the higher potency and efficacy of target cell lysis by AFM11 relative to a bispecific tandem scFv (that is currently in clinical evaluation). CD8+ T cells dominate early AFM11-mediated cytotoxicity (4 hrs) while after 24 hrs both CD4+ and CD8+ T cells equally contribute to tumor lysis with EC50 between 0.5 – 5 pM; cytotoxicity was independent of CD19 cell-surface density. AFM11 exhibited similar cytotoxicity over effector:target ratios ranging from 5:1 to 1:5, and facilitated serial T cell-killing of its targets. The advantage of AFM11 over the bispecific tandem scFv was most pronounced at lower effector:target ratios. AFM11 activated T cells only in the presence of CD19+ cells. In PBMC cultures, AFM11 induced CD69 and CD25 expression, T cell proliferation, and production of IFN-γ, TNF-α, IL-2, IL-6, and IL-10. Depletion of CD19+ cells from PBMC abrogated these effects, demonstrating that the T cell activation is strictly CD19+ target-dependent. Thus, AFM11 should not elicit the devastating cytokine release observed when full-length antibodies bind CD3. Up to one week co-incubation with AFM11 did not inhibit T cell cytotoxicity, suggesting that the TandAb does not induce anergy. In vivo, AFM11 induced dose-dependent growth inhibition of Raji tumors; a single 0.5 mg/kg dose exhibited efficacy similar to 5 daily injections. In the tissue cross reactivity study, only tissues containing CD19+ and CD3+ cells were stained by AFM11; all other tissues, including vital organs, displayed no cross reactivity. Similarly, no local intolerance was observed in rabbits, and no effect on myeloid and erythroid progenitors was observed in a colony-forming assay. Strong accumulation of 125I-labeled AFM11 was observed in the tumors of mice engrafted with CD19+ cancer cells, and no unspecific organ accumulation was observed. Finally, evaluated on the basis of Cmax and the area under the curve (AUC), AFM11 exhibited dose linearity (20 – 500 mg AFM11 dose range) upon single i.v. bolus administration in mice; half-life (T1/2) ranged from 18.4 to 22.9 hr. In summary, AFM11 is a highly efficacious novel drug candidate for the treatment of CD19+ malignancies with an advantageous safety profile and anticipated dosing regimen. Disclosures: Zhukovsky: Affimed Therapeutics AG: Employment, Equity Ownership. Reusch:Affimed Therapeutics AG: Employment. Burkhardt:Affimed Therapeutics AG: Employment. Knackmuss:Affimed Therapeutics AG: Employment. Fucek:Affimed Therapeutics AG: Employment. Eser:Affimed Therapeutics AG: Employment. McAleese:Affimed Therapeutics AG: Employment. Ellwanger:Affimed Therapeutics AG: Employment. Little:Affimed Therapeutics AG: Consultancy, Equity Ownership.

Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 4070-4070
Author(s):  
Harbani Malik ◽  
Ben Buelow ◽  
Udaya Rangaswamy ◽  
Aarti Balasubramani ◽  
Andrew Boudreau ◽  
...  

Introduction The restricted expression of CD19 in the B-cell lineage makes it an attractive target for the therapeutic treatment of B-cell malignancies. Many monoclonal antibodies and antibody drug conjugates targeting CD19 have been developed, including bispecific T-cell redirecting antibodies (T-BsAbs). In addition, anti-CD19 chimeric antigen receptor T-cells (CAR-T) have been approved to treat leukemia and lymphoma. However, despite the impressive depth of responses achieved by T-cell redirecting approaches such as T-BsAbs and CAR-T cells, toxicity from over-activation of T-cells remains a substantial limitation for this type of therapy, in particular neurotoxicity. In designing TNB-486, a novel CD19 x CD3 T-BsAb, we endeavored to retain activity against CD19-positive tumor cells while limiting the cytokine secretion thought to underlie toxicity from T-cell redirecting therapies. Utilizing TeneoSeek, a next generation sequencing (NGS)-based discovery pipeline that leverages in silico analysis of heavy chain only/fixed light chain antibody (HCA/Flic, respectively) sequences to enrich for antigen specific antibodies, we made a high affinity αCD19 HCA and a library of αCD3 Flic antibodies that showed a >2 log range of EC50s for T cell activation in vitro. Of note, the library contained a low-activating αCD3 that induced minimal cytokine secretion even at concentrations that mediated saturating T-cell dependent lysis of lymphoma cells (when paired with an αCD19 HCA). We characterized the relative efficacy and potential therapeutic window of this unique molecule, TNB-486, in vitro and in vivo and compared it to two strongly activating bispecific CD19 x CD3 antibodies similar to those currently available and in clinical development. Methods Affinity measurements of the αCD19 moiety were made via Biacore (protein) and flow cytometry (cell surface). Stability measurements were made by subjecting the molecule to thermal stress and the %aggregation was measured by Size Exclusion Chromatography. T-cell activation was measured via flow cytometry (CD69 and CD25 expression) and cytokine was measured by ELISA (IL-2, IL-6, IL-10, INF-ɣ, and TNFα) in vitro. Lysis of B-cell tumor cell lines (Raji, RI-1, and Nalm6) was measured via flow cytometry in vitro. In vivo, NOG mice were engrafted subcutaneously with NALM-6 or SUDHL-10 cells and intravenously with human peripheral blood mononuclear cells (huPBMC), and the mice treated with multiple doses of TNB-486 or negative or positive control antibody. Tumor burden was evaluated via caliper measurement. Pharmacodynamic/Pharmacokinetic (PK/PD) studies were performed in NOG mice. A pharmacokinetic (PK) study was performed in BALB/c mice, and a tolerability and PK study are ongoing in cynomolgus monkeys. Results TNB-486 bound to cell surface CD19 with single digit nanomolar affinity (~3nM). EC50s for cytotoxicity were in the single-digit nanomolar range for TNB-486, and sub-nanomolar for the strongly activating controls; TNB-486 maximum achievable lysis was identical to the positive controls. TNB-486 induced significantly less cytokine release for all cytokines tested compared to the positive controls even at doses saturating for tumor lysis. No off-target activation was observed in the absence of CD19 expressing target cells. In vivo, TNB-486 eradicated all CD19-positive tumors tested (NALM-6 and SUDHL10) at doses as little as 1µg administered every four days after tumors had reached ~200mm3. TNB-486 showed a PK profile consistent with other IgG molecules in mice (T1/2 ~6 days in mice). Conclusions TNB-486 induced comparable lysis of CD19-positive tumor cells as the strongly activating control bispecific antibodies while inducing significantly reduced cytokine secretion, even at doses saturating for tumor lysis in vitro. In vivo TNB-486 eradicated all tested CD19 positive tumor cell lines in established tumor models. No off-target binding was observed. In summary, TNB-486 shows promise as a lymphoma therapeutic differentiated from T-cell targeted therapies currently in the clinic and in clinical trials. Disclosures Malik: Teneobio, Inc.: Employment, Equity Ownership. Buelow:Teneobio, Inc.: Employment, Equity Ownership. Rangaswamy:Teneobio, Inc.: Employment, Equity Ownership. Balasubramani:Teneobio, Inc.: Employment, Equity Ownership. Boudreau:Teneobio, Inc.: Employment, Equity Ownership. Dang:Teneobio, Inc.: Employment, Equity Ownership. Davison:Teneobio, Inc.: Employment, Equity Ownership. Force Aldred:Teneobio, Inc.: Equity Ownership. Iyer:Teneobio, Inc.: Employment, Equity Ownership. Jorgensen:Teneobio, Inc.: Employment, Equity Ownership. Pham:Teneobio, Inc.: Employment, Equity Ownership. Prabhakar:Teneobio, Inc.: Employment, Equity Ownership. Schellenberger:Teneobio, Inc.: Employment, Equity Ownership. Ugamraj:Teneobio, Inc.: Employment, Equity Ownership. Trinklein:Teneobio, Inc.: Employment, Equity Ownership. Van Schooten:Teneobio, Inc.: Employment, Equity Ownership.


Blood ◽  
2020 ◽  
Vol 136 (Supplement 1) ◽  
pp. 34-35
Author(s):  
Moriah Rabin ◽  
Mengyan Li ◽  
Scott Garforth ◽  
Jacqueline Marino ◽  
Jian Hua Zheng ◽  
...  

Background: While chimeric antigen receptor T cells (CAR T-cells) induce dramatic remissions of refractory or recurrent B cell malignancies, the durability of these remissions is frequently limited by subsequent reduction in circulating CAR T-cells and/or by diminution of their effector function. We hypothesized that we could overcome this therapeutic limitation and increase the functional activity and longevity of CAR T-cells by selectively deriving them from virus-specific effector memory T cells. We have developed biologics we termed synTacs (artificial immunological synapse for T-cell activation), which selectively activate and expand antigen-specific CD8+ T cells in vitro and in vivo by recapitulating signals delivered at the immunological synapse. The synTacs consist of dimeric Fc domain scaffolds linking CD28- or 4-1BB-specific ligands to HLA-A2 MHC molecules covalently tethered to virus-derived peptides. Treatment of PBMCs from CMV-exposed donors with synTacs presenting a CMV-derived peptide (pp65-NLVPMVATV) induce vigorous and selective ex vivo and in vivo expansion of highly functional CMV-specific CD8+ T cells, with potent antiviral activity. We used these synTacs to selectively generate CAR T-cells from CMV-specific effector memory CD8+ T cells, which could be further expanded by restimulation with the CMV-specific synTacs. Methods: We treated PBMCs from CMV-exposed donors in media supplemented with either IL-2 or IL-7/12/15 with a synTac containing the CMV-derived pp65 peptide presented by HLA-A2 MHC molecules linked to ligands capable of stimulating CD28- or 4-1BB-dependent costimulatory pathways. PBMCs activated either with anti-CD3/CD28 or the CMV-specific synTacs were transduced with lentivirus expressing an anti-CD19 CAR and a GFP reporter gene. CMV-specific CD8+ T cells were quantified by tetramer staining and CAR T-cells were detected by GFP expression determined by flow cytometric analysis. The functional activity of the CD19 CAR T-cells was determined by a B cell-specific cytotoxic assay. Results: After 7 days, treatment of PBMCs with CMV-specific synTacs rapidly induced robust activation and >50-fold expansion of CMV-specific CD8+ T cells expressing effector memory markers. Treatment of the PBMCs with CMV-specific synTacs selectively activated CMV-specific T cells and enabled them to be specifically transduced with a CD19-specific CAR lentivirus and converted into CD19 CAR T-cells. These CMV-specific CD19 CAR T-cells displayed potent dose-responsive cytotoxic activity targeting purified primary B cells. Furthermore, these CMV-specific CD19 CAR T-cells could be selectively expanded by in vitro treatment with CMV-specific synTacs. Conclusions: SynTacs are versatile immunotherapeutics capable of selective in vitro and in vivo activation and expansion of virus-specific CD8+ T cells with potent antiviral cytotoxic activity. After selective lentiviral transduction and conversion into CD19 CAR T-cells, their co-expression of the CMV-specific T cell receptor enabled them to be potently stimulated and activated by in vitro treatment with CMV synTacs. The modular design of synTacs facilitates efficient coupling of other costimulatory ligands - such as OX40 or GITRL - or cytokines, such as IL-2, IL-7, or IL-15, to enable the selective in vivo delivery of defined costimulatory signals or cytokines to the CAR T-cells expressing CMV-specific TCR. This strategy has the potential to boost the in vivo activity of tumor-specific CAR T-cells after infusion and enable more durable and potent treatment of refractory/recurrent B cell malignancies. Disclosures Almo: Cue Biopharma: Current equity holder in publicly-traded company, Patents & Royalties: Patent number: 62/013,715, Research Funding. Goldstein:Cue Biopharma: Research Funding.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 1974-1974 ◽  
Author(s):  
Adam D. Cohen ◽  
J. Joseph Melenhorst ◽  
Alfred L. Garfall ◽  
Simon F Lacey ◽  
Megan Davis ◽  
...  

Abstract Background: Relapsed/refractory (rel/ref) MM is associated with progressive immune dysfunction, including reversal of CD4:CD8 T cell ratio and acquisition of terminally-differentiated T cell phenotypes. BCMA-directed CAR T cells have promising activity in MM, but the factors that predict for robust in vivo expansion and responses are not known. In a phase 1 study of CART-BCMA (autologous T cells expressing a human BCMA-specific CAR with CD3ζ/4-1BB signaling domains) in refractory MM patients (median 7 priors, 96% high-risk cytogenetics), we observed partial response (PR) or better in 12/25 (47%) (Cohen et al, ASH 2017, #505). Recently, we demonstrated in CLL pts receiving CD19-directed CAR T cells that certain T cell phenotypes prior to generation of the CAR T product were associated with improved in vivo expansion and clinical outcomes (Fraietta et al, Nat Med 2018). We thus sought to identify pre-treatment clinical or immunological features associated with CART-BCMA expansion and/or response. Methods: Three cohorts were enrolled: 1) 1-5 x 108 CART cells alone; 2) cyclophosphamide (Cy) 1.5 g/m2 + 1-5 x 107 CART cells; and 3) Cy 1.5 g/m2 + 1-5 x 108 CART cells. Phenotypic analysis of peripheral blood (PB) and bone marrow (BM) mononuclear cells, frozen leukapheresis aliquots, and phenotype and in vitro kinetics of CART-BCMA growth during manufacturing were performed by flow cytometry. CART-BCMA in vivo expansion was assessed by flow cytometry and qPCR. Responses were assessed by IMWG criteria. Results: Responses (≥PR) were seen in 4/9 pts (44%, 1 sCR, 2 VPGR, 1 PR) in cohort 1; 1/5 (20%, 1 PR) in cohort 2; and 7/11 (64%, 1 CR, 3 VGPR, 3 PR) in cohort 3. As of 7/9/18, 3/25 (12%) remain progression-free at 11, 14, and 32 months post-infusions. As previously described, responses were associated with both peak in vivo CART-BCMA expansion (p=0.002) as well as expansion over first month post-infusion (AUC-28, p=0.002). No baseline clinical or MM-related characteristic was significantly associated with expansion or response, including age, isotype, time from diagnosis, # prior therapies, being quad- or penta-refractory, presence of del 17p or TP53 mutation, serum hemoglobin, BM MM cell percentage, MM cell BCMA intensity, or soluble BCMA concentration. Treatment regimen given before leukapheresis or CART-BCMA infusions also had no predictive value. We did find, however, that higher CD4:CD8 T cell ratios within the leukapheresis product were associated with greater in vivo CART-BCMA expansion (Spearman's r=0.56, p=0.005) and clinical response (PR or better; p=0.014, Mann-Whitney). In addition, and similar to our CLL data, we found that a higher frequency of CD8 T cells within the leukapheresis product with an "early-memory" phenotype of CD45RO-CD27+ was also associated with improved expansion (Spearman's r=0.48, p=0.018) and response (p=0.047); Analysis of manufacturing data confirmed that higher CD4:CD8 ratio at culture start was associated with greater expansion (r=0.41, p=0.044) and, to a lesser degree, responses (p=0.074), whereas absolute T cell numbers or CD4:CD8 ratio in final CART-BCMA product was not (p=NS). In vitro expansion during manufacturing did associate with in vivo expansion (r=0.48, p=0.017), but was not directly predictive of response. At the time of CART-BCMA infusion, the frequency of total T cells, CD8+ T cells, NK cells, B cells, and CD3+CD56+ cells within the PB or BM was not associated with subsequent CART-BCMA expansion or clinical response; higher PB and BM CD4:CD8 ratio pre-infusion correlated with expansion (r=0.58, p=0.004 and r=0.64, p=0.003, respectively), but not with response. Conclusions: In this study, we found that CART-BCMA expansion and responses in heavily-pretreated MM patients were not associated with tumor burden or other clinical characteristics, but did correlate with certain immunological features prior to T cell collection and manufacturing, namely preservation of normal CD4:CD8 ratio and increased frequency of CD8 T cells with a CD45RO-CD27+ phenotype. This suggests that patients with less dysregulated immune systems may generate more effective CAR T cell products in MM, and has implications for optimizing patient selection, timing of T cell collection, and manufacturing techniques to try to overcome these limitations in MM patients. Disclosures Cohen: Celgene: Consultancy; Novartis: Research Funding; Oncopeptides: Consultancy; Janssen: Consultancy; Poseida Therapeutics, Inc.: Research Funding; Bristol Meyers Squibb: Consultancy, Research Funding; Kite Pharma: Consultancy; GlaxoSmithKline: Consultancy, Research Funding; Seattle Genetics: Consultancy. Melenhorst:Parker Institute for Cancer Immunotherapy: Research Funding; novartis: Patents & Royalties, Research Funding; Casi Pharmaceuticals: Consultancy; Incyte: Research Funding; Shanghai UNICAR Therapy, Inc: Consultancy. Garfall:Amgen: Research Funding; Kite Pharma: Consultancy; Bioinvent: Research Funding; Novartis: Research Funding. Lacey:Novartis Pharmaceuticals Corporation: Patents & Royalties; Parker Foundation: Research Funding; Tmunity: Research Funding; Novartis Pharmaceuticals Corporation: Research Funding. Davis:Novartis Institutes for Biomedical Research, Inc.: Patents & Royalties. Vogl:Karyopharm Therapeutics: Consultancy. Pruteanu:Novartis: Employment. Plesa:Novartis: Research Funding. Young:Novartis: Patents & Royalties, Research Funding. Levine:Novartis: Consultancy, Patents & Royalties, Research Funding; CRC Oncology: Consultancy; Incysus: Consultancy; Tmunity Therapeutics: Equity Ownership, Research Funding; Brammer Bio: Consultancy; Cure Genetics: Consultancy. June:Novartis Pharmaceutical Corporation: Patents & Royalties, Research Funding; Immune Design: Membership on an entity's Board of Directors or advisory committees; Tmunity Therapeutics: Equity Ownership, Membership on an entity's Board of Directors or advisory committees, Patents & Royalties, Research Funding; Novartis Pharmaceutical Corporation: Patents & Royalties, Research Funding; Immune Design: Membership on an entity's Board of Directors or advisory committees; Celldex: Consultancy, Membership on an entity's Board of Directors or advisory committees; Tmunity Therapeutics: Equity Ownership, Membership on an entity's Board of Directors or advisory committees, Patents & Royalties, Research Funding. Stadtmauer:Takeda: Consultancy; Celgene: Consultancy; Amgen: Consultancy; AbbVie, Inc: Research Funding; Janssen: Consultancy. Milone:Novartis: Patents & Royalties.


Blood ◽  
2010 ◽  
Vol 116 (21) ◽  
pp. 2330-2330
Author(s):  
Constantijn J.M. Halkes ◽  
Inge Jedema ◽  
Judith Olde Wolbers ◽  
Esther M van Egmond ◽  
Peter A. Von Dem Borne ◽  
...  

Abstract Abstract 2330 In vivo T cell depletion with anti-thymocyte globulin (ATG) or alemtuzumab (anti-CD52) before reduced intensity allogeneic stem cell transplantation (alloSCT) in combination with in vitro T cell depletion with alemtuzumab reduces the risk of GVHD. Detectable levels of circulating antibodies are present up to several months after the alloSCT, leading to a delayed immune reconstitution which is associated with an increased incidence of opportunistic infections and early relapses. Prior to 2007, combined in vitro (Alemtuzumab 20 mg added “to the bag”) and in vivo T cell depletion with horse-derived ATG (h-ATG) resulted in good engraftment without GVHD in the absence of GVHD prophylaxis after reduced intensity alloSCT using conditioning with fludarabine and busulphan. Due to the unavailability of h-ATG, rabbit-derived ATG (r-ATG) 10–14 mg/kg was introduced in the conditioning regimen in 2007. Strikingly, in this cohort of patients, early EBV reactivation and EBV-associated post-transplantation lymphoproliferative disease (PTLD) was observed in 10 out of 18 patients at a median time of 6 weeks after alloSCT (range 5 to 11 weeks) in the absence of GVHD or immunosuppressive treatment. Analysis of T and B cell recovery early after transplantation revealed preferential depletion of T cells as compared to B cells, thereby allowing unrestricted proliferation of EBV infected B cells. Due to this unacceptable high incidence of EBV-related complications, in the conditioning regimen r-ATG was replaced by low dose alemtuzumab (15 mg i.v. day -4 and -3) in 2008. In this cohort of 60 patients, only 2 patients experienced transient EBV reactivation during the first 3 months after alloSCT and one patient developed an EBV-associated lymphoma 4 weeks after alloSCT. To investigate the mechanisms underlying the low incidence of EBV reactivation using alemtuzumab for T cell depletion, we studied the in vivo and in vitro effects of alemtuzumab on different lymphocyte subsets. First, lineage-specific reconstitution was studied in 20 patients from the alemtuzumab cohort with known CD52 negative diseases (11 AML and 9 multiple myeloma) to exclude the confounding effect of antibody absorption by malignant cells. Whereas at 3 weeks after alloSCT detectable numbers of circulating NK cells and T cells were observed (medians 71 (range 6–378), and 12 (range 1–1164)E6/L, respectively), no circulating B cells could be detected (median 0, range 0–1 E6/L). At 6 weeks after alloSCT, NK and T cell numbers further increased (medians 212 (52-813), and 130 (range 25–1509)E6/L, respectively), whereas B cell numbers still remained low in the majority of patients (median 15, range 0–813E6/L). In all patients, T cells were detectable before the appearance of circulating B cells. Furthermore, the expression of CD52 and the sensitivity to alemtuzumab-mediated complement-dependent cell lysis (CDC) of B cells, T cells and NK cells was measured in vitro. The highest CD52 expression was observed on B cells (mean fluorescence intensity (MFI) 120), resulting in 95% lysis after incubation with 10ug/mL alemtuzumab and rabbit complement. NK cells showed a significantly lower CD52 expression (MFI 41), which was also reflected by a lower susceptibility to alemtuzumab-mediated CDC (62% lysis). Interestingly, differential expression of CD52 was observed on CD4 and CD8 T cells (MFI 120 and 101, respectively). Cytotoxicity analysis revealed relative protection of CD8 compared to CD4 T cells against alemtuzumab-mediated CDC, resulting in 52% and 90% lysis, respectively. Based on these results, we investigated in detail the presence and phenotype of the CD4 and CD8 subsets and EBV-specific CD8 T cells using tetramer staining at 6 weeks after alloSCT. In accordance with the in-vitro expression and susceptibility data, circulating CD52+ CD8 T cells including EBV-specific T cells were detectable. Interestingly, the majority of circulating CD4 T cells (64-93%, n=4) lacked CD52 expression, explaining their capacity to persist in the presence of alemtuzumab. We conclude that in vivo and in vitro T cell depletion with alemtuzumab is associated with a relatively low risk of EBV-associated PTLD because of efficient B cell depletion and persistent EBV immunity allowed by the relative insusceptibility for alemtuzumab of CD8 T cells and the development of CD52 negative escape variants of CD4 T cells. Disclosures: No relevant conflicts of interest to declare.


Blood ◽  
2012 ◽  
Vol 120 (21) ◽  
pp. 3722-3722
Author(s):  
Ruth A. Chenault ◽  
Rebecca Gottschalk ◽  
Gabriela Hernandez-Hoyos ◽  
Jennifer Wiens ◽  
Brian Gordon ◽  
...  

Abstract Abstract 3722 Background: Despite advances in treatments for B-cell leukemias and lymphomas, many patients ultimately relapse and succumb to disease following multiple courses of therapy. Bispecific antibody fragments that can simultaneously engage T cells and tumor cells have been shown, in the literature, to destroy tumor cells by effectively redirecting the cytotoxic function of T cells. T-cell engaging bispecific molecules linking anti-CD19 and anti-CD3 binding domains in the context of novel SCORPION™ (multi-specific protein therapeutic) proteins were evaluated both in vitro and in vivo for function and stability. Methods: Redirected T-cell cytotoxicity (RTCC) was measured by combining CD19 positive or negative cell lines with SCORPION proteins in the presence of human T cells. In a similar assay context, CFSE-labeled T cells were monitored for activation and proliferation. Functional RTCC assays were also used to analyze serum stability of SCORPION molecules in vitro and to complete an in vivo pharmacokinetic analysis. In vivo efficacy was assessed by monitoring the rate of tumor outgrowth of Ramos xenografts co-implanted with human peripheral blood mononuclear cells (PBMC) in NOD/SCID mice after treatment with SCORPION molecules. Results: SCORPION molecules potently mediate target-specific T-cell cytotoxicity toward tumor cell lines presenting cell surface CD19, with EC50 values for cytotoxicity at low pM concentrations. These molecules also demonstrate induction of T-cell activation and proliferation in the presence of target-bearing tumor cells but not in the absence of target expression. SCORPION molecules retain stable function following incubation at 37°C in mouse serum for up to a week in vitro, and pharmacokinetic analysis of SCORPION protein function in BALB/c mouse serum following intravenous administration resulted in half-life estimates of 69–84 hours. In efficacy studies conducted in NOD/SCID mice, SCORPION proteins significantly inhibited the outgrowth of Ramos tumor xenografts in the presence of human effector cells. Conclusion: SCORPION molecules targeting CD19 and CD3 effectively harness the cytotoxic activity of T cells to kill CD19 positive tumor cells both in vitro and in vivo and show potential for further investigation as possible therapeutic agents for B-cell malignancies. Disclosures: Chenault: Emergent BioSolutions: Employment. Gottschalk:Emergent BioSolutions: Employment. Hernandez-Hoyos:Emergent BioSolutions: Employment. Wiens:Emergent BioSolutions: Employment. Gordon:Emergent BioSolutions: Employment. Klee:Emergent BioSolutions: Employment, Equity Ownership. Bienvenue:Emergent BioSolutions: Employment. Dasovich:Emergent BioSolutions: Employment. Kumer:Emergent BioSolutions: Employment. Aguilar:Emergent BioSolutions: Employment. Bannink:Emergent BioSolutions: Employment, Equity Ownership. McMahan:Emergent BioSolutions: Employment, Equity Ownership. Natarajan:Emergent BioSolutions: Employment, Equity Ownership. Algate:Emergent BioSolutions: Employment, Equity Ownership. Blankenship:Emergent BioSolutions: Employment, Equity Ownership, Patents & Royalties.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 250-250
Author(s):  
Jan A Rath ◽  
Gagan Bajwa ◽  
Benoit Carreres ◽  
Isabelle Gruber ◽  
Elisabeth Hoyer ◽  
...  

Introduction:Transgenic co-expression of a major histocompatibility complex class I restricted tumor associated antigen specific TCR and CD8αβ (TCR8) has been previously proposed as a strategy to redirect CD4+ T cells to tumors. However, it is unknown whether forced TCR8 expression induces more fundamental transcriptional consequences in both CD4+ and CD8+ T cells, and whether T cell lineage origin affects this outcome. Here we deeply interrogate the effects of transgenic TCR and TCR8 in human CD4+ and CD8+ T cells upon leukemia challenge by single cell RNA sequencing (scRNAseq) and investigate T cell function in vitro and in vivo. We identify profound changes of gene expression that have significant functional consequences. Methods:A previously characterized HLA-A*02:01 restricted survivin-specific TCR was used (Arber et al, JCI, 2015 Jan;125(1):157-68) and a new polycistronic vector with this TCR and CD8αβ was generated. CD4+ and CD8+ T cells were isolated and scRNAseq (25'474 cells in total) was performed on (1) freshly isolated cells, (2) retrovirally transduced (TCR or TCR8) expanded cells, and (3) TCR+CD8+, TCR8+CD8+ and TCR8+CD4+ T cells co-cultured with BV173 leukemia cells (HLA-A*02:01+survivin+). scRNAseq results were cross-validated in independent experiments with FACS analysis of selected markers, in vitro stress-killing assays, analysis of cytokine production, and assessment of anti-tumor function in vivo in xenograft mice. Results:CD4+ T cells only killed BV173 leukemia cells when redirected with TCR8 but not with TCR alone (p=0.0004, n=7), while killing by TCR+CD8+ and TCR8+CD8+ T was comparable (p=NS). To explore some of the possible underlying mechanisms, we used dimensionality reduction and unsupervised clustering of the scRNASeq data and identified 19 distinct cell clusters. CD4+ and CD8+ lineage origin clearly separated the samples, but separation by transgene type only became apparent upon co-culture. Analyzing differentially expressed genes, we found that co-cultured samples contained clusters with high expression of cytotoxic markers but with significant differences between CD4+ and CD8+ lineages (e.g. transcription of GZMB in CD4+ T cells and GNLY, NKG7, GZMK in CD8+ T cells). Next, we analyzed which genes were upregulated from the expanded to co-cultured states. Co-cultured TCR8+CD4+ T cells had more upregulated genes with a broader diversity compared to TCR+CD8+ or TCR8+CD8+ T cells. Among these upregulated pathways were cytotoxicity, co-stimulation, oxidative phosphorylation, NFkB regulation, cell growth and transcription factors. TCR8+CD4+ T cells also retained a less differentiated phenotype (e.g. high IL7R, SELL, CCR7, CXCR4) with preservation of replicative potential. Furthermore, co-cultured TCR8+CD4+ T cells expressed more co-stimulatory and less activation/ exhaustion markers. In addition, co-cultured TCR8+CD4+ T cells heavily relied on oxidative phosphorylation and had higher mitochondrial activity compared to co-cultured TCR+ or TCR8+ CD8+ T cells. In stress co-cultures with multiple rounds of tumor challenge, TCR8+CD4+ T cells outperformed TCR+CD8+ T cells (number of killings TCR8+CD4+ vs TCR+CD8+: 3.3±0.5 vs 1.3±1.1, p=0.01, n=7), but were comparable to TCR8+CD8+ T cells (TCR8+CD4+ vs TCR8+CD8+: 3.3±0.5 vs 2±1.4,p=NS, n=7). TCR8+CD4+ T cells expanded significantly better than TCR+CD8+ T cells (p=0.002) and TCR8+CD8+ T cells (p=0.015) and produced TH1 type cytokines. In the xenograft mouse model, we observed significant BV173 leukemia control in mice treated with TCR+CD8+ T cells compared to controls (NT), and further enhancement in mice treated with TCR8+CD8+ T cells (NT vs TCR: p=0.0002, NT vs TCR8: p<0.0001, TCR vs TCR8: p=0.01, n=5). TCR8+CD4+ T cells also significantly delayed leukemia progression compared to TCR+CD4+ or NT T cells (p=0.001, n=5). Conclusion:Transgenic TCR8 expression has previously been proposed as a strategy to enhance TCR-pMHC recognition. Here we identify profound transcriptional changes involving multiple pathways that are important for sustained anti-tumor function upon adoptive T cell transfer in vivo, such as cytotoxicity, co-stimulation, cell cycle and metabolism. Our results point towards previously unrecognized mechanisms by which TCR8 transgenes mediate their beneficial effect in both CD4+ and CD8+ T cells. Disclosures Brenner: T Scan: Membership on an entity's Board of Directors or advisory committees; Marker Therapeutics: Equity Ownership; Allovir: Equity Ownership, Membership on an entity's Board of Directors or advisory committees; Tessa Therapeutics: Equity Ownership; Memgen: Membership on an entity's Board of Directors or advisory committees; Allogene: Membership on an entity's Board of Directors or advisory committees. Arber:Cell Medica: Patents & Royalties.


Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 2785-2785
Author(s):  
Li Guo ◽  
Rukhsana Aslam ◽  
Yajing Zhao ◽  
Edwin R. Speck ◽  
Heyu Ni ◽  
...  

Abstract Primary immune thrombocytopenia (ITP) is an autoimmune disease characterized by increased platelet destruction and/or impaired megakaryocyte production, mediated by autoreactive B cells and T cells. B cell depletion therapy by rituximab, a monoclonal human anti-CD20 antibody, has been shown effective in both anti-platelet antibody positive (B cell mediated) and negative (T cell mediated) ITP patients. Those patients responsive to rituximab therapy showed normalized CD4+ and CD8+ T cell responses (Stasi et al. Blood. 2007), however, the mechanism of T cell regulation by B cell depletion is not clear. One possibility is through normalization of CD4+ T helper cells or up-regulation of CD4+ regulatory T cells (Tregs) (Stasi et al. Blood. 2008). Another possibility is by suppression of activated conventional CD8+ T cells or the up-regulation of CD8+ Tregs. We examined the changes of both CD4+ and CD8+ T cells and Tregs (CD25highFoxp3+) after B cell depletion in vivo in our ITP mouse model. Briefly, BALB/c GPIIIa (CD61) KO mice were either given PBS (ND) or mouse monoclonal anti-CD20 antibody (B-dep, Biogen) at day -1 and day 13 (250ug/mouse, ip). Residual CD19+ B cells in peripheral blood were less than 0.1% within 24hours in the latter group. All mice were immunized by transfusions of wildtype (WT) platelets at day 0, 7, 14, and 21 (1×108/mouse, iv). At day 28, we examined the percentages of T cell subsets in the spleens of the immunized mice. B cell-depleted immune CD61 KO mice showed significantly higher percentages of both CD3+CD8+ T cells and CD8+CD25highFoxp3+ T cells (Table 1). There was no significant difference in the CD3+CD4+ and CD4+CD25highFoxp3+ T cell populations. Both ND and B-dep immune CD61 KO splenocytes showed increased cytotoxicity activity against CD61+ PU5-1.8 target cells in vitro compared with naïve CD61 KO splenocytes, indicating the activation of CD8+ T cells. To test their in vivo effect on ITP development, splenocytes were engrafted from immune mice into irradiated and AsialoGM-1 treated severe combined immunodeficient (SCID) mice at a dose of 2.5×104/mouse and the mice were monitored for weekly platelet counts. ND and in vitro B cell depleted splenocytes from immune KO mice induced persistent ITP during 3 weeks observation whereas splenocytes from B-dep immune mice did not. To further confirm the role of B cell depletion on CD8+ T cell responses, CD8+ T cells from either ND or B-dep immune CD61 KO splenocytes were purified and transferred into SCID mice at 3×104/mouse. CD4+ T cells from ND immune CD61 KO splenocytes were added at 3×104/mouse to all the SCID mice to support the CD8+ T cell survival in vivo. SCID mice received CD8+ T cells from B-dep group showed higher platelet count at Day 14. Overall, our results indicate a protective role of CD8+CD25highFoxp3+ T cells against the development of cell mediated ITP that is enhanced by B cell depleting therapy in vivo. Table 1. CD61 KO MouseSpleens CD3+CD8+(%) CD8+CD25highFoxp3+ (%) Naïve Control 9.12±0.37 0.12±0.08 Immune, ND 6.78±2.37 0.0925±0.03 Immune, B-dep 14.15±5.1 0.2367±0.11 P value (ND vs B-dep) 0.0007 0.0064 Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 3789-3789
Author(s):  
Kalpana Parvathaneni ◽  
Ai-Hong Zhang ◽  
David W. Scott

Abstract To modulate B-cell responsiveness to FVIII, we previously generated cytotoxic cells expressing FVIII C2 or A2 immunodominant domains as chimeric receptors. We termed these antigen-expressing engineered T cells, "BARs", for B-cell Antibody Receptor. These CD8 T cells directly interact and kill FVIII-specific B cells and anti-FVIII hybridomas in prophylactic experiments in vitro and in vivo. It was not known whether these BAR CD8s could function or would be blocked in the presence of circulating antibodies to the expressed BAR domains. To test this, we cultured FVIII C2 or A2 BAR CD8 T cell with a mixture of monoclonal antibodies specific for these domains (up to 10 BU), and then added them to spleen cells from FVIII-immunized mice. These spleen cells were then re-stimulated with FVIII and the antibody response was determined after 5 days. Our results showed that these BAR CD8 T cells were not blocked in their ability to suppress the antibody response to FVIII under these conditions. Coupled with the observation that BAR-T cells can be stimulated to proliferate by anti-FVIII monoclonals, these results suggest that BAR cytotoxic activity may still be effective in the presences of inhibitors. (Supported by NIH grant R01 HL126727) Disclosures No relevant conflicts of interest to declare.


2002 ◽  
Vol 197 (1) ◽  
pp. 19-26 ◽  
Author(s):  
Melanie S. Vacchio ◽  
Richard J. Hodes

Whereas ligation of CD28 is known to provide a critical costimulatory signal for activation of CD4 T cells, the requirement for CD28 as a costimulatory signal during activation of CD8 cells is less well defined. Even less is known about the involvement of CD28 signals during peripheral tolerance induction in CD8 T cells. In this study, comparison of T cell responses from CD28-deficient and CD28 wild-type H-Y–specific T cell receptor transgenic mice reveals that CD8 cells can proliferate, secrete cytokines, and generate cytotoxic T lymphocytes efficiently in the absence of CD28 costimulation in vitro. Surprisingly, using pregnancy as a model to study the H-Y–specific response of maternal T cells in the presence or absence of CD28 costimulation in vivo, it was found that peripheral tolerance does not occur in CD28KO pregnants in contrast to the partial clonal deletion and hyporesponsiveness of remaining T cells observed in CD28WT pregnants. These data demonstrate for the first time that CD28 is critical for tolerance induction of CD8 T cells, contrasting markedly with CD28 independence of in vitro activation, and suggest that the role of CD28/B7 interactions in peripheral tolerance of CD8 T cells may differ significantly from that of CD4 T cells.


2008 ◽  
Vol 205 (13) ◽  
pp. 2965-2973 ◽  
Author(s):  
Susan Gilfillan ◽  
Christopher J. Chan ◽  
Marina Cella ◽  
Nicole M. Haynes ◽  
Aaron S. Rapaport ◽  
...  

Natural killer (NK) cells and CD8 T cells require adhesion molecules for migration, activation, expansion, differentiation, and effector functions. DNAX accessory molecule 1 (DNAM-1), an adhesion molecule belonging to the immunoglobulin superfamily, promotes many of these functions in vitro. However, because NK cells and CD8 T cells express multiple adhesion molecules, it is unclear whether DNAM-1 has a unique function or is effectively redundant in vivo. To address this question, we generated mice lacking DNAM-1 and evaluated DNAM-1–deficient CD8 T cell and NK cell function in vitro and in vivo. Our results demonstrate that CD8 T cells require DNAM-1 for co-stimulation when recognizing antigen presented by nonprofessional antigen-presenting cells; in contrast, DNAM-1 is dispensable when dendritic cells present the antigen. Similarly, NK cells require DNAM-1 for the elimination of tumor cells that are comparatively resistant to NK cell–mediated cytotoxicity caused by the paucity of other NK cell–activating ligands. We conclude that DNAM-1 serves to extend the range of target cells that can activate CD8 T cell and NK cells and, hence, may be essential for immunosurveillance against tumors and/or viruses that evade recognition by other activating or accessory molecules.


Sign in / Sign up

Export Citation Format

Share Document