4-1BB costimulation ameliorates T cell exhaustion induced by tonic signaling of chimeric antigen receptors

Adrienne H Long, Waleed M Haso, Jack F Shern, Kelsey M Wanhainen, Meera Murgai, Maria Ingaramo, Jillian P Smith, Alec J Walker, M Eric Kohler, Vikas R Venkateshwara, Rosandra N Kaplan, George H Patterson, Terry J Fry, Rimas J Orentas, Crystal L Mackall, Adrienne H Long, Waleed M Haso, Jack F Shern, Kelsey M Wanhainen, Meera Murgai, Maria Ingaramo, Jillian P Smith, Alec J Walker, M Eric Kohler, Vikas R Venkateshwara, Rosandra N Kaplan, George H Patterson, Terry J Fry, Rimas J Orentas, Crystal L Mackall

Abstract

Chimeric antigen receptors (CARs) targeting CD19 have mediated dramatic antitumor responses in hematologic malignancies, but tumor regression has rarely occurred using CARs targeting other antigens. It remains unknown whether the impressive effects of CD19 CARs relate to greater susceptibility of hematologic malignancies to CAR therapies, or superior functionality of the CD19 CAR itself. We show that tonic CAR CD3-ζ phosphorylation, triggered by antigen-independent clustering of CAR single-chain variable fragments, can induce early exhaustion of CAR T cells that limits antitumor efficacy. Such activation is present to varying degrees in all CARs studied, except the highly effective CD19 CAR. We further determine that CD28 costimulation augments, whereas 4-1BB costimulation reduces, exhaustion induced by persistent CAR signaling. Our results provide biological explanations for the antitumor effects of CD19 CARs and for the observations that CD19 CAR T cells incorporating the 4-1BB costimulatory domain are more persistent than those incorporating CD28 in clinical trials.

Figures

Figure 1
Figure 1
GD2.28z CAR T cells have discrepant in vitro and in in vivo activity. (a) CD19 and GD2 antigen expression on the 143B-CD19 osteosarcoma line. Representative of n=5. (b) In vitro51chromium release assay evaluating cytolytic activity of mock-transduced, CD19.28z CAR, and GD2.28z CAR T cells against 143B-CD19. Assay performed 9 days after initial activation. n=3 replicates/point; representative of 4 donors. (c) Growth curves of 143B-CD19 tumors in NSG mice. Mice were inoculated with 106 143B-CD19 periosteally on day 0 followed by adoptive transfer of 107 CAR transduced or mock-transduced T cells on day 14. Mock n=5, CD19.28z CAR n=7, GD2.28z CAR n=7. (d) Left: composition of T cell product adoptively transferred into mice in (c). Middle and right: quantification of T cells within the spleen and tumor 14 days following adoptive T cell transfer into mice from (c). n=8 mice/group. * = p<0.05 by Student’s T-test. Bar graphs represent mean ± SEM. In vivo results are representative of four experiments.
Figure 2
Figure 2
GD2.28z CAR T cells become exhausted during ex vivo expansion. (a) Activation marker expression 4–7 days after initial activation. Representative of 3 donors. (b) Expansion during ex vivo culture. n=3 replicates/point; representative from 3 donors. (c) Quantification of apoptosis of CAR T cells generated from 4 unique donors evaluated 9–10 days following initial activation. (d) Representative exhaustion marker expression 9 days following initial activation. (e) Quantification of exhaustion marker expression pooled from 4 donors, 9–11 days following activation. (f) Exhaustion marker SPICE analysis from (e). (g) ΔΔCT q-RT-PCR expression levels of exhaustion transcription factors relative to mock, 9–11 days following initial activation. n=4 technical replicates; representative of 3 donors. (h) Cytokine production of flow-sorted single-transduced (CD19.28z CAR+ or GD2.28z CAR+) and co-transduced (Co-trans; CD19.28z CAR+ and GD2.28z CAR+) T cells, co-incubated with 143B-CD19 at 9 days after initial activation. n=3 replicates/group; representative of 3 donors. T cells with media: <5 pg/mL IL-2, TNF-α and IFN-γ. (i) Tumor growth curves of NSG mice inoculated with 106 143B-CD19 periosteally on day 0 followed by adoptive transfer of 107 transduced CAR T cells on day 14. n=5 mice/group. (j) Left: composition of T cell product adoptively transferred into mice in (i). Middle and right: quantification of T cells within the spleen and tumor 14 days following adoptive T cell transfer into mice from (i). n=6 mice/group. SPICE analysis: * = p<0.05 by Wilcoxon signed rank test. All other data: * = p<0.05 by Student’s T-test. Bar graphs represent mean ± SEM.
Figure 3
Figure 3
Tonic CAR signaling during ex vivo expansion leads to early exhaustion. (a) Western blot evaluating phosphorylation levels of CAR signaling domains versus total CAR signaling domains, using an anti-phospho-CD3ζ and anti-CD3ζ antibody, respectively. Evaluated day 5 after initial activation. “Basal” phosphorylation evaluated without further stimulation. “After crosslinking” evaluated following incubation with OKT3, anti-CD19 CAR idiotype, or anti-GD2 CAR idiotype antibodies. Representative of 3 donors. (b) Nine amino acid point mutations were introduced to eliminate signaling via the GD2.28z CAR (GD2.mut-28.mut-z CAR). (c) Activation and (d) exhaustion marker expression of GD2.mut-28.mut-z CAR T cells 10 days after initial activation. Representative of 3 donors. (e) Ex vivo expansion and (f-g) exhaustion marker expression of CD19.28z CAR T cells cultured ± anti-idiotype antibody (anti-Id; 0.25 ug/ml) and a crosslinking F(ab’)2 (2.5 ug/ml). n=4 replicates/group; representative of 3 donors. (h) Tumor growth curves of NSG mice inoculated with 106 143B-CD19 periosteally on day 0 followed by adoptive transfer of 107 transduced CAR T cells on day 14. Mice received mock-transduced or CD19.28z CAR cultured with or without anti-idiotype antibody. No anti-idiotype antibody was given to mice. n=8 mice/group. (i) CD19 expression on tumors 10 days following adoptive transfer into mice from (h). (j) Quantification of T cells within the blood 9 days following adoptive T cell transfer into mice from (h). n=3 mice/group. In vivo results are representative of two experiments. * = p<0.05 by Student’s T-test. Bar graphs represent mean ± SEM.
Figure 4
Figure 4
Tonic GD2.28z CAR signaling is antigen independent. (a) In vitro51Cr release assay against the 143B-CD19 osteosarcoma cell line. Both GD2.28z CAR and GD2.mutCDR.28z CAR had comparable transduction efficiencies (85 and 81%, respectively). Assay performed 9 days after initial T cell activation. n=3 replicates/point; representative of 3 donors. (b) Exhaustion marker expression on GD2.mutCDR.28z CAR vs. GD2.28z CAR T cells. Representative of 3 donors. (c) Fluorescence microscopy of T cells transduced with CAR-Cerulean fusion proteins (blue). (d) Quantification of the number of puncta per cell from (c). 30 cells per group; repeated for 3 donors. (e) Exhaustion marker expression levels of a version of the GD2.28z CAR designed to be structurally more like the CD19.28z CAR, based upon removal of the IgG1 hinge and substitution of the GD2 linker with the CD19 linker (GD2.sh.28z CAR). (f) Cytokine release levels upon co-incubation with 143B-CD19 on day 10 of ex vivo culture. # designates values >10,000 pg/mL. n=3 replicates/group; representative of 3 donors. (g) Structure and (h) exhaustion marker expression of a hybrid CAR combining the CDR regions of the CD19.28z CAR and the framework regions of the GD2.28z CAR (CD19-CDR.GD2-FW.28z CAR), 9 days post activation during ex vivo culture. Representative of 3 donors. For cell puncta analysis, * = p<0.0001 by Kolmogorov-Smirnov test. All other data, * = p<0.05 by Student’s T-test. Bar graphs represent mean ± SEM.
Figure 5
Figure 5
4-1BB endodomain ameliorates exhaustion in CAR T cells. (a) Activation marker expression 7 days following initial activation and (b) exhaustion marker expression 9 days after initial activation of GD2 CARs with the CD28 or 4-1BB co-stimulatory domains. (Representative of 4 donors). (c) Quantification of exhaustion marker expression pooled from 4 donors, 9–11 days following initial activation. (d) Exhaustion marker SPICE analysis of (c). (e) Cytokine release upon co-incubation with 143B-CD19 for 24 hours starting on day 10 following initial activation. n=3 replicates/group; representative of 3 donors. (f) Tumor growth curves of NSG mice inoculated with 5 × 105 143B-CD19 periosteally on day 0, then treated with 3 mg cyclophosphamide intraperitoneally on day 2 and with adoptive transfer of 107 transduced CAR T cells on day 4. n=10 mice/group. (g) Quantification of T cells within the blood 8 and 15 days following adoptive transfer into mice from (f). n=8 mice/group. (h) Exhaustion marker expression of CAR T cells from (f) on day 14 following adoptive transfer. (i) Quantification of T cells within the blood of NSG mice inoculated with 106 NALM6-GL followed by adoptive transfer of 5×106 CAR T cells three days later. T cells were quantified on day 8 following adoptive transfer. n=5 mice/group. (j) Exhaustion marker expression of CAR T cells from (i). In vivo results are representative of three experiments. SPICE analysis: * = p<0.05 by Wilcoxon signed rank test. All other data: * = p<0.05 by Student’s T-test. Bar graphs represent mean ± SEM.
Figure 6
Figure 6
Ameliorating effect of 4-1BB signaling is associated with a unique transcriptional profile. (a) Principal component analysis (PCA) of global transcriptional profiles of CAR T cells generated from 3 unique healthy donors, evaluated 9 days following initial activation. Transduction efficiency was >90% for all samples. (b) Heatmap of genes previously reported to impact T cells exhaustion. Exhaustion related transcription factors (TBX21, EOMES, PRDM1, IKZF2), inhibitory receptors (LAG3, HAVCR2, CTLA4, BTLA, CD244), and transcription factors reported to be preferentially expressed in memory vs. exhausted cells (KLF6, JUN, JUNB) were compared. (c) Representative GSEA results from running the unfiltered GD2.BBz CAR vs. GD2.28z CAR T cell rank list against the MSigDB C5 gene ontology sets. (d) Heatmap of genes uniquely dysregulated in exhausted GD2.28z CAR T cells. Genes are those up/downregulated > 2 fold in GD2.28z CAR vs. CD19.28z CAR T cells, up/downregulated > 2 fold in GD2.28z CAR vs. GD2.BBz CAR T cells, and < 2 fold difference in GD2.BBz CAR vs. CD19.28z CAR T cells. Genes associated with response to hypoxia (EGLN3, EGR, PTGIS, ID1), apoptosis (ID1), and metabolism (GLUL, ATP10D, SMPDL3A), or T cell suppressive pathways (CTLA4, CD38, LGMN, CLECL1, ENTPD1, KLRC1/2) were identified.

References

    1. Lee DW, et al. The Future Is Now: Chimeric Antigen Receptors as New Targeted Therapies for Childhood Cancer. Clinical Cancer Research. 2012;18(10):2780–2790.
    1. Sadelain M, Brentjens R, Rivière I. The Basic Principles of Chimeric Antigen Receptor Design. Cancer Discovery. 2013;3(4):388–398.
    1. Lee DW, et al. T cells expressing CD19 chimeric antigen receptors for acute lymphoblastic leukaemia in children and young adults: a phase 1 dose-escalation trial. The Lancet. 2014
    1. Maude SL, et al. Chimeric Antigen Receptor T Cells for Sustained Remissions in Leukemia. New England Journal of Medicine. 2014;371(16):1507–1517.
    1. Kochenderfer JN, et al. B-cell depletion and remissions of malignancy along with cytokine-associated toxicity in a clinical trial of anti-CD19 chimeric-antigen-receptor-transduced T cells. Blood. 2012;119(12):2709–2720.
    1. Porter DL, et al. Chimeric antigen receptor-modified T cells in chronic lymphoid leukemia. New England Journal of Medicine. 2011;365(8):725–733.
    1. Grupp SA, et al. Chimeric Antigen Receptor-Modified T Cells for Acute Lymphoid Leukemia. New England Journal of Medicine. 2013;368(16):1509–1518.
    1. Savoldo B, et al. CD28 costimulation improves expansion and persistence of chimeric antigen receptor-modified T cells in lymphoma patients. The Journal of Clinical Investigation. 2011;121(5):1822–1826.
    1. Brentjens RJ, et al. CD19-Targeted T Cells Rapidly Induce Molecular Remissions in Adults with Chemotherapy-Refractory Acute Lymphoblastic Leukemia. Science Translational Medicine. 2013;5(177):177ra138–177ra138.
    1. Davila ML, et al. Efficacy and Toxicity Management of 19-28z CAR T Cell Therapy in B Cell Acute Lymphoblastic Leukemia. Science Translational Medicine. 2014;6(224) 224ra225.
    1. Kershaw MH, et al. A phase I study on adoptive immunotherapy using gene-modified T cells for ovarian cancer. Clinical Cancer Research. 2006;12(20):6106–6115.
    1. Park JR, et al. Adoptive transfer of chimeric antigen receptor re-directed cytolytic T lymphocyte clones in patients with neuroblastoma. Molecular therapy. 2007;15(4):825–833.
    1. Pule MA, et al. Virus-specific T cells engineered to coexpress tumor-specific receptors: persistence and antitumor activity in individuals with neuroblastoma. Nature medicine. 2008;14(11):1264–1270.
    1. Louis CU, et al. Antitumor activity and long-term fate of chimeric antigen receptor-positive T cells in patients with neuroblastoma. Blood. 2011;118(23):6050–6056.
    1. Till BG, et al. CD20-specific adoptive immunotherapy for lymphoma using a chimeric antigen receptor with both CD28 and 4-1BB domains: pilot clinical trial results. Blood. 2012;119(17):3940–3950.
    1. Lamers CHJ, et al. Treatment of Metastatic Renal Cell Carcinoma With CAIX CAR-engineered T cells: Clinical Evaluation and Management of On-target Toxicity. Mol Ther. 2013;21(4):904–912.
    1. Robbins PF, et al. Cutting edge: persistence of transferred lymphocyte clonotypes correlates with cancer regression in patients receiving cell transfer therapy. The Journal of Immunology. 2004;173(12):7125–7130.
    1. Kowolik CM, et al. CD28 Costimulation Provided through a CD19-Specific Chimeric Antigen Receptor Enhances In vivo Persistence and Antitumor Efficacy of Adoptively Transferred T Cells. Cancer Research. 2006;66(22):10995–11004.
    1. Milone MC, et al. Chimeric receptors containing CD137 signal transduction domains mediate enhanced survival of T cells and increased antileukemic efficacy in vivo. Molecular Therapy. 2009;17(8):1453–1464.
    1. Haso W, et al. Anti-CD22-chimeric antigen receptors targeting B-cell precursor acute lymphoblastic leukemia. Blood. 2013;121(7):1165–1174.
    1. Ahmadzadeh M, et al. Tumor antigen-specific CD8 T cells infiltrating the tumor express high levels of PD-1 and are functionally impaired. Blood. 2009;114(8):1537–1544.
    1. Sakuishi K, et al. Targeting Tim-3 and PD-1 pathways to reverse T cell exhaustion and restore anti-tumor immunity. The Journal of experimental medicine. 2010;207(10):2187–2194.
    1. Baitsch L, et al. Exhaustion of tumor-specific CD8+ T cells in metastases from melanoma patients. The Journal of clinical investigation. 2011;121(6):2350.
    1. Zhou Q, et al. Coexpression of Tim-3 and PD-1 identifies a CD8+ T-cell exhaustion phenotype in mice with disseminated acute myelogenous leukemia. Blood. 2011;117(17):4501–4510.
    1. Woo S-R, et al. Immune inhibitory molecules LAG-3 and PD-1 synergistically regulate T-cell function to promote tumoral immune escape. Cancer research. 2012;72(4):917–927.
    1. Topalian SL, et al. Safety, activity, and immune correlates of anti-PD-1 antibody in cancer. New England Journal of Medicine. 2012;366(26):2443–2454.
    1. Virgin HW, Wherry EJ, Ahmed R. Redefining chronic viral infection. Cell. 2009;138(1):30–50.
    1. Wherry EJ. T cell exhaustion. Nat Immunol. 2011;12(6):492–499.
    1. Rossig C, et al. Targeting of GD2-positive tumor cells by human T lymphocytes engineered to express chimeric T-cell receptor genes. International journal of cancer. 2001;94(2):228–236.
    1. Pulè MA, et al. A chimeric T cell antigen receptor that augments cytokine release and supports clonal expansion of primary human T cells. Molecular Therapy. 2005;12(5):933–941.
    1. Hudecek M, et al. The non-signaling extracellular spacer domain of chimeric antigen receptors is decisive for in vivo antitumor activity. Cancer immunology research. 2014 canimm. 0127.2014.
    1. Kochenderfer JN, et al. Construction and Pre-clinical Evaluation of an Anti-CD19 Chimeric Antigen Receptor. Journal of immunotherapy (Hagerstown, Md: 1997) 2009;32(7):689.
    1. Shin H, et al. A Role for the Transcriptional Repressor Blimp-1 in CD8+ T Cell Exhaustion during Chronic Viral Infection. Immunity. 2009;31(2):309–320.
    1. Doering TA, et al. Network Analysis Reveals Centrally Connected Genes and Pathways Involved in CD8< sup>+</sup> T Cell Exhaustion versus Memory. Immunity. 2012;37(6):1130–1144.
    1. Paley MA, et al. Progenitor and terminal subsets of CD8+ T cells cooperate to contain chronic viral infection. Science. 2012;338(6111):1220–1225.
    1. Friend LD, et al. A dose-dependent requirement for the proline motif of CD28 in cellular and humoral immunity revealed by a targeted knockin mutant. The Journal of experimental medicine. 2006;203(9):2121–2133.
    1. Dodson LF, et al. Targeted knock-in mice expressing mutations of CD28 reveal an essential pathway for costimulation. Molecular and cellular biology. 2009;29(13):3710–3721.
    1. Zhao Y, et al. A herceptin-based chimeric antigen receptor with modified signaling domains leads to enhanced survival of transduced T lymphocytes and antitumor activity. The Journal of Immunology. 2009;183(9):5563–5574.
    1. Nieba L, Honegger A, Krebber C, Plückthun A. Disrupting the hydrophobic patches at the antibody variable/constant domain interface: improved in vivo folding and physical characterization of an engineered scFv fragment. Protein Engineering. 1997;10(4):435–444.
    1. Dolezal O, et al. Single-chain Fv multimers of the anti-neuraminidase antibody NC10: the residue at position 15 in the VL domain of the scFv-0 (VL− VH) molecule is primarily responsible for formation of a tetramer-trimer equilibrium. Protein engineering. 2003;16(1):47–56.
    1. Whitlow M, et al. Multivalent Fvs: characterization of single-chain Fv oligomers and preparation of a bispecific Fv. Protein Engineering. 1994;7(8):1017–1026.
    1. Wherry EJ, et al. Molecular Signature of CD8+ T Cell Exhaustion during Chronic Viral Infection. Immunity. 2007;27(4):670–684.
    1. Crawford A, et al. Molecular and Transcriptional Basis of CD4+ T Cell Dysfunction during Chronic Infection. Immunity. 40(2):289–302.
    1. Guest RD, et al. The role of extracellular spacer regions in the optimal design of chimeric immune receptors: evaluation of four different scFvs and antigens. Journal of Immunotherapy. 2005;28(3):203.
    1. James SE, et al. Antigen sensitivity of CD22-specific chimeric TCR is modulated by target epitope distance from the cell membrane. The Journal of Immunology. 2008;180(10):7028–7038.
    1. Frigault MJ, et al. Identification of chimeric antigen receptors that mediate constitutive or inducible proliferation of T cells. Cancer immunology research. (in the press)
    1. Agnellini P, et al. Impaired NFAT nuclear translocation results in split exhaustion of virus-specific CD8+ T cell functions during chronic viral infection. Proceedings of the National Academy of Sciences. 2007;104(11):4565–4570.
    1. Vezys V, et al. 4-1BB signaling synergizes with programmed death ligand 1 blockade to augment CD8 T cell responses during chronic viral infection. The Journal of Immunology. 2011;187(4):1634–1642.
    1. Wang C, et al. 4-1BBL induces TNF receptor-associated factor 1-dependent Bim modulation in human T cells and is a critical component in the costimulation-dependent rescue of functionally impaired HIV-specific CD8 T cells. The Journal of Immunology. 2007;179(12):8252–8263.
    1. Porter D, et al. Chimeric Antigen Receptor Modified T Cells Directed Against CD19 (CTL019 cells) Have Long-Term Persistence and Induce Durable Responses In Relapsed, Refractory CLL. Blood. 2013;122(21):4162–4162.
    1. Doedens AL, et al. Hypoxia-inducible factors enhance the effector responses of CD8+ T cells to persistent antigen. Nat Immunol. 2013;14(11):1173–1182.
    1. Pearce EL, et al. Enhancing CD8 T-cell memory by modulating fatty acid metabolism. Nature. 2009;460(7251):103–107.
    1. van der Windt Gerritje JW, et al. Mitochondrial Respiratory Capacity Is a Critical Regulator of CD8+ T Cell Memory Development. Immunity. 2012;36(1):68–78.
    1. Araki K, et al. mTOR regulates memory CD8 T-cell differentiation. Nature. 2009;460(7251):108–112.
    1. Roederer M, Nozzi JL, Nason MC. SPICE: Exploration and analysis of post-cytometric complex multivariate datasets. Cytometry Part A. 2011;79(2):167–174.
Methods-only references
    1. Morgan RA, et al. Recognition of glioma stem cells by genetically modified T cells targeting EGFRvIII and development of adoptive cell therapy for glioma. Human gene therapy. 2012;23(10):1043–1053.
    1. Brochet X, Lefranc M-P, Giudicelli V. IMGT/V-QUEST: the highly customized and integrated system for IG and TR standardized V-J and V-D-J sequence analysis. Nucleic Acids Research. 2008;36(suppl 2):W503–W508.
    1. Sen G, et al. Induction of IgG antibodies by an anti-idiotype antibody mimicking disialoganglioside GD2. Journal of Immunotherapy. 1998;21(1):75–83.
    1. Jena B, et al. Chimeric Antigen Receptor (CAR)-Specific Monoclonal Antibody to Detect CD19-Specific T Cells in Clinical Trials. PLoS ONE. 2013;8(3):e57838.
    1. Edelstein A, et al. Computer control of microscopes using µManager. Current protocols in molecular biology. 2010 14.20. 11-14.20. 17.
    1. Elangovan M, et al. Characterization of one-and two-photon excitation fluorescence resonance energy transfer microscopy. Methods. 2003;29(1):58–73.
    1. Schindelin J, et al. Fiji: an open-source platform for biological-image analysis. Nature methods. 2012;9(7):676–682.
    1. Olenych SG, Claxton NS, Ottenberg GK, Davidson MW. The fluorescent protein color palette. Current Protocols in Cell Biology. 2007
    1. Martin CE, et al. IL-7/anti-IL-7 mAb complexes augment cytokine potency in mice through association with IgG-Fc and by competition with IL-7R. Blood. 2013;121(22):4484–4492.

Source: PubMed

3
Prenumerera