BOXR1030, an anti-GPC3 CAR with exogenous GOT2 expression, shows enhanced T cell metabolism and improved anti-cell line derived tumor xenograft activity

Taylor L Hickman, Eugene Choi, Kathleen R Whiteman, Sujatha Muralidharan, Tapasya Pai, Tyler Johnson, Avani Parikh, Taylor Friedman, Madaline Gilbert, Binzhang Shen, Luke Barron, Kathleen E McGinness, Seth A Ettenberg, Greg T Motz, Glen J Weiss, Amy Jensen-Smith, Taylor L Hickman, Eugene Choi, Kathleen R Whiteman, Sujatha Muralidharan, Tapasya Pai, Tyler Johnson, Avani Parikh, Taylor Friedman, Madaline Gilbert, Binzhang Shen, Luke Barron, Kathleen E McGinness, Seth A Ettenberg, Greg T Motz, Glen J Weiss, Amy Jensen-Smith

Abstract

Purpose: The solid tumor microenvironment (TME) drives T cell dysfunction and inhibits the effectiveness of immunotherapies such as chimeric antigen receptor-based T cell (CAR T) cells. Early data has shown that modulation of T cell metabolism can improve intratumoral T cell function in preclinical models.

Experimental design: We evaluated GPC3 expression in human normal and tumor tissue specimens. We developed and evaluated BOXR1030, a novel CAR T therapeutic co-expressing glypican-3 (GPC3)-targeted CAR and exogenous glutamic-oxaloacetic transaminase 2 (GOT2) in terms of CAR T cell function both in vitro and in vivo.

Results: Cell surface expression of tumor antigen GPC3 was observed by immunohistochemical staining in tumor biopsies from hepatocellular carcinoma, liposarcoma, squamous lung cancer, and Merkel cell carcinoma patients. Compared to control GPC3 CAR alone, BOXR1030 (GPC3-targeted CAR T cell that co-expressed GOT2) demonstrated superior in vivo efficacy in aggressive solid tumor xenograft models, and showed favorable attributes in vitro including an enhanced cytokine production profile, a less-differentiated T cell phenotype with lower expression of stress and exhaustion markers, an enhanced metabolic profile and increased proliferation in TME-like conditions.

Conclusions: Together, these results demonstrated that co-expression of GOT2 can substantially improve the overall antitumor activity of CAR T cells by inducing broad changes in cellular function and phenotype. These data show that BOXR1030 is an attractive approach to targeting select solid tumors. To this end, BOXR1030 will be explored in the clinic to assess safety, dose-finding, and preliminary efficacy (NCT05120271).

Conflict of interest statement

I have read the journal’s policy and the authors of this manuscript have the following competing interests: TLH is an employee of Takeda Pharmaceuticals, a former employee of Unum Therapeutics, has ownership interest in Unum Therapeutics (now Cogent Biosciences), and has issued patents: PCT/US18/00028 and PCT/US2018/015999-all outside the submitted work. EC is an employee of Catamaran Bio, Inc, a former employee of Unum Therapeutics; has ownership interest in Unum Therapeutics (now Cogent Biosciences) and Catamaran Bio, Inc; and has issued patents: PCT/US2010/0041032A1, PCT/US2019/0153061A1, PCT/US2019/0105348A1, and PCT/US2019/0284298A1-all outside the submitted work. KRW is an employee of SOTIO Biotech Inc., a former employee of Unum Therapeutics; has ownership interest in Unum Therapeutics (now Cogent Biosciences) and has issued patents: PCT/US2012/0282175A1 and PCT/US2010/0028346A1 all outside the submitted work. SM is an employee of SOTIO Biotech Inc., a former employee of Kiniksa Pharmaceuticals. TP is an employee of Novartis and a former employee of Unum therapeutics. TJ is an employee of Novartis, a former employee of Unum Therapeutics, and has ownership interest in Unum Therapeutics (now Cogent Biosciences). TF is an employee of Biogen and a former employee of Unum therapeutics. MG is an employee of Arrakis Therapeutics and a former employee of Unum therapeutics. BS is an employee of Tango Therapeutics, a former employee of Unum Therapeutics, and has issued patents: US9464333, US9388422, US9303250, WO2014055778A2, WO2016069774A1, and WO2017049094A1-all outside the submitted work. LB is an employee of Catamaran Bio, Inc. and a former employee of Unum therapeutics. KEM is an employee of Arrakis Therapeutics, a former employee of Unum Therapeutics, and a consultant for SOTIO Biotech Inc.; has ownership interest in Unum Therapeutics (now Cogent Biosciences), and ownership interests in Baxter International and Takeda Pharmaceutical Co. outside the submitted work; is an inventor multiple patents PCT/US2019/046550, PCT/US2019/050013, PCT/US2019/040346, and PCT/US2019/060287 related to the submitted work; and US 8,252,913, US 8,461,318, US 8,598,327, US 10,144,770, PCT/US2019/044512, PCT/US2018/015999, PCT/US2017/023064, PCT/US2010/023599-outside of the submitted work. SAE is an employee of BlueRock Therapeutics, a former employee of Unum Therapeutics, has ownership interest in Unum Therapeutics (now Cogent Biosciences), and is a co-inventor on patents filed related to this work. GTM is an employee of BlueRock Therapeutics, a former employee of Unum Therapeutics; has ownership interest in Unum Therapeutics (now Cogent Biosciences), and is a named inventor on many CAR-T patents. GJW is an employee of SOTIO Biotech Inc., a former employee of Unum Therapeutics; reports personal fees from Spring Bank Pharmaceuticals, Imaging Endpoints II, MiRanostics Consulting, Gossamer Bio, Paradigm, International Genomics Consortium, Angiex, IBEX Medical Analytics, GLG Council, Guidepoint Global, Genomic Health, Rafael Pharmaceuticals, SPARC-all outside this submitted work; has ownership interest in Unum Therapeutics (now Cogent Biosciences), and ownership interests in MiRanostics Consulting, Exact Sciences, Moderna, Agenus, Aurinia Pharmaceuticals, and Circulogene-outside the submitted work; and has issued patents: PCT/US2008/072787, PCT/US2010/043777, PCT/US2011/020612, and PCT/US2011/037616-all outside the submitted work and is an inventor on a patent filed related to this work. AJS is an employee of SOTIO Biotech Inc.; a former employee of bluebird bio; has ownership interest in bluebird bio, Pfizer, Regeneron, Exelixis, Bristol Myers Squibb, Celgene, and Alkermes-all outside the submitted work. All other authors have no other competing interests related to this work. This does not alter our adherence to PLOS ONE policies on sharing data and materials.

Figures

Fig 1. GPC3 target expression in human…
Fig 1. GPC3 target expression in human normal and tumor tissues by IHC.
(A) Immunohistochemical staining of normal healthy TMA for GPC3 expression. Representative images (20x magnification) of different organs with different levels of GPC3 are shown (B) IHC staining of tissues from hepatocellular carcinoma (HCC), squamous cell lung cancer (SCC), Merkel cell carcinoma (MCC) and liposarcoma patients for GPC3 expression. Representative images (20x magnification) show samples with different levels of GPC3 (and H-scores).
Fig 2. Co-expression of GPC3-targeted CAR with…
Fig 2. Co-expression of GPC3-targeted CAR with exogenous GOT2 in BOXR1030 T cells.
(A) Depiction of expression construct and domains for the GPC3 CAR and BOXR1030 used in studies. (B) Representative surface expression of the anti-GPC3 scFv following transduction, measured on day 9 by flow cytometry from a single donor. (C) Summarized CAR expression by flow cytometry for n = 5 healthy donors. (D) mRNA copy number of endogenous and exogenous (codon optimized) GOT2 measured by qPCR (n = 3 donors). (E) The frequency of CD45RA+ CD27+, CD45RA+ CD27-, CD45RA- CD27+, and CD45RA- CD27- T cells in the indicated subsets at baseline (n = 11 healthy donors; no stimulation). P-values were determined by paired t-test. Data represented as mean+/- standard deviation (SD).
Fig 3. AST enzymatic activity in BOXR1030…
Fig 3. AST enzymatic activity in BOXR1030 CAR T cells.
(A) Schematic describing the role of GOT2 in mitochondrial conversion of glutamate to aKG and OAA to aspartate. Image created with biorender.com. (B) Intracellular aspartate levels were measured using a plate-based colorimetric assay comparing BOXR1030 and control CAR T cells under non-stimulated conditions. (C) BOXR1030 and control CAR T cells were activated for 8 hours with GPC3+ Hep3B target cells and evaluated for AST activity. Data represented as mean +/- SD of two technical replicates.
Fig 4. In vitro BOXR1030 activity.
Fig 4. In vitro BOXR1030 activity.
(A) Expression of GPC3 target antigen on target cell lines were measured by flow cytometry and represented as median fluorescence intensity (MFI). (B) Histogram plots of GPC3 MFI for selected cell lines. (C-F) T cell cytokine release of BOXR1030 and control CAR T cells was measured by ELISA following co-culture with target cell lines for 24 hours. Results for IFN-g (C) and IL-2 (D) are the averages across 3 donors, and error bars indicate SD. Results for TNF-a (E) and IL-17A (F) are for a single donor, and error bars indicate SD between technical replicates. (G) Target cell cytotoxicity was measured with a luciferase-based assay, and percent cytotoxicity was normalized to samples treated with untransduced T cells following 24 hours of co-culture (n = 3 donors, error bars indicate SD). (H) T cell proliferation was evaluated after a 7 day incubation with target cell lines. CAR+ T cell counts were measured by flow cytometry. (n = 3 donors, error bars indicate SD). (I) Intracellular cytokine levels were measured by flow cytometry in CD4+ and CD8+ T cells (n = 10, error bars indicate SEM).
Fig 5. In vitro activity of BOXR1030…
Fig 5. In vitro activity of BOXR1030 under conditions simulating the solid tumor microenvironment.
BOXR1030 or control CAR T cells were repeat stimulated with GPC3+ target cell lines on day 0 and on day 3 in hypoxic conditions (A) or low glucose conditions (B). (A and B) The gMFI of CellTrace Violet (CTV) was measured for BOXR1030 and control CAR stimulated with target cells on day 7 in the indicated culture conditions and proliferation was plotted as 1/gMFI. (n = 5 for JHH7 stimulated conditions and n = 11 for Hep3B stimulated conditions, statistical analysis was performed using a 2-tailed paired t-test, and p-values

Fig 6. In vivo anti-tumor activity of…

Fig 6. In vivo anti-tumor activity of BOXR1030 T cells.

(A) Plot on the right…

Fig 6. In vivo anti-tumor activity of BOXR1030 T cells.
(A) Plot on the right shows Hep3B tumor-bearing model (mean tumor volume 108.7±34.1mm3) NSG mice were treated with two weekly doses of 1 x 106 CAR+ control or BOXR1030 T cells each (total dose of 2 x 106 CAR+ cells) (dosing days indicated by arrows) and tumor volumes were measured over the course of 110 days. Plot on the left shows data for individual mice and plot on the right shows mean data. Mean tumor volume plots are discontinued when less than 50% of group is remaining. (B) Plot on the right shows JHH7 tumor-bearing mice (mean tumor volume 49.8±7.2mm3) were treated with two weekly doses of 5 x 106 CAR+ control or BOXR1030 T cells each (total dose of 10 x 106 CAR+ cells) (dosing days indicated by arrows) and tumor volumes were measured out to 50 days. Plot on the left shows data for individual mice and plot on the right shows mean data. Mean tumor volume plots are discontinued when less than 50% of group is remaining. (C) CAR+ T cells were measured in peripheral blood of Hep3G tumor bearing mice on days 15, 25, 40 and 60 post T cell treatment and data are reported as counts per ul of blood. (D) Percent PD1+TIM3+CD4+ and PD1+TIM3+CD8+ tumor infiltrating T cells were measured by FACS on days 7 and 14 following T cell administration. (E) Biodistribution of BOXR1030 T cells was measured in mouse tissues by qPCR at days 15 and 45 post treatment. Data are reported as BOXR1030 copies / 100ng DNA.
Fig 6. In vivo anti-tumor activity of…
Fig 6. In vivo anti-tumor activity of BOXR1030 T cells.
(A) Plot on the right shows Hep3B tumor-bearing model (mean tumor volume 108.7±34.1mm3) NSG mice were treated with two weekly doses of 1 x 106 CAR+ control or BOXR1030 T cells each (total dose of 2 x 106 CAR+ cells) (dosing days indicated by arrows) and tumor volumes were measured over the course of 110 days. Plot on the left shows data for individual mice and plot on the right shows mean data. Mean tumor volume plots are discontinued when less than 50% of group is remaining. (B) Plot on the right shows JHH7 tumor-bearing mice (mean tumor volume 49.8±7.2mm3) were treated with two weekly doses of 5 x 106 CAR+ control or BOXR1030 T cells each (total dose of 10 x 106 CAR+ cells) (dosing days indicated by arrows) and tumor volumes were measured out to 50 days. Plot on the left shows data for individual mice and plot on the right shows mean data. Mean tumor volume plots are discontinued when less than 50% of group is remaining. (C) CAR+ T cells were measured in peripheral blood of Hep3G tumor bearing mice on days 15, 25, 40 and 60 post T cell treatment and data are reported as counts per ul of blood. (D) Percent PD1+TIM3+CD4+ and PD1+TIM3+CD8+ tumor infiltrating T cells were measured by FACS on days 7 and 14 following T cell administration. (E) Biodistribution of BOXR1030 T cells was measured in mouse tissues by qPCR at days 15 and 45 post treatment. Data are reported as BOXR1030 copies / 100ng DNA.

References

    1. June CH, Sadelain M. Chimeric Antigen Receptor Therapy. N Engl J Med. N Engl J Med; 2018;379: 64–73. doi: 10.1056/NEJMra1706169
    1. Kershaw MH, Westwood JA, Parker LL, Wang G, Eshhar Z, Mavroukakis SA, et al.. A phase I study on adoptive immunotherapy using gene-modified T cells for ovarian cancer. Clin Cancer Res. Clin Cancer Res; 2006;12: 6106–15. doi: 10.1158/1078-0432.CCR-06-1183
    1. Morgan RA, Yang JC, Kitano M, Dudley ME, Laurencot CM, Rosenberg SA. Case report of a serious adverse event following the administration of T cells transduced with a chimeric antigen receptor recognizing ERBB2. Mol Ther. Elsevier; 2010;18: 843–51. doi: 10.1038/mt.2010.24
    1. O’Rourke DM, Nasrallah MP, Desai A, Melenhorst JJ, Mansfield K, Morrissette JJD, et al.. A single dose of peripherally infused EGFRvIII-directed CAR T cells mediates antigen loss and induces adaptive resistance in patients with recurrent glioblastoma. Sci Transl Med. Sci Transl Med; 2017;9. doi: 10.1126/scitranslmed.aaa0984
    1. D’Aloia MM, Zizzari IG, Sacchetti B, Pierelli L, Alimandi M. CAR-T cells: the long and winding road to solid tumors. Cell Death Dis. Cell Death Dis; 2018;9: 282. doi: 10.1038/s41419-018-0278-6
    1. Le Bourgeois T, Strauss L, Aksoylar H-I, Daneshmandi S, Seth P, Patsoukis N, et al.. Targeting T Cell Metabolism for Improvement of Cancer Immunotherapy. Front Oncol. Front Oncol; 2018;8: 237. doi: 10.3389/fonc.2018.00237
    1. Knochelmann HM, Smith AS, Dwyer CJ, Wyatt MM, Mehrotra S, Paulos CM. CAR T Cells in Solid Tumors: Blueprints for Building Effective Therapies. Front Immunol. Front Immunol; 2018;9: 1740. doi: 10.3389/fimmu.2018.01740
    1. Martinez M, Moon EK. CAR T Cells for Solid Tumors: New Strategies for Finding, Infiltrating, and Surviving in the Tumor Microenvironment. Front Immunol. Front Immunol; 2019;10: 128. doi: 10.3389/fimmu.2019.00128
    1. Henze J, Tacke F, Hardt O, Alves F, Al Rawashdeh W. Enhancing the Efficacy of CAR T Cells in the Tumor Microenvironment of Pancreatic Cancer. Cancers (Basel). Multidisciplinary Digital Publishing Institute (MDPI); 2020;12. doi: 10.3390/cancers12061389
    1. Motz GT, Coukos G. Deciphering and reversing tumor immune suppression. Immunity. Immunity; 2013;39: 61–73. doi: 10.1016/j.immuni.2013.07.005
    1. Thommen DS, Schumacher TN. T Cell Dysfunction in Cancer. Cancer Cell. Cancer Cell; 2018;33: 547–562. doi: 10.1016/j.ccell.2018.03.012
    1. Jenkins Y, Zabkiewicz J, Ottmann O, Jones N. Tinkering under the Hood: Metabolic Optimisation of CAR-T Cell Therapy. Antibodies (Basel, Switzerland). Antibodies (Basel); 2021;10. doi: 10.3390/antib10020017
    1. Rostamian H, Fallah-Mehrjardi K, Khakpoor-Koosheh M, Pawelek JM, Hadjati J, Brown CE, et al.. A metabolic switch to memory CAR T cells: Implications for cancer treatment. Cancer Lett. Cancer Lett; 2021;500: 107–118. doi: 10.1016/j.canlet.2020.12.004
    1. Pellegrino M, Del Bufalo F, De Angelis B, Quintarelli C, Caruana I, de Billy E. Manipulating the Metabolism to Improve the Efficacy of CAR T-Cell Immunotherapy. Cells. Cells; 2020;10. doi: 10.3390/cells10010014
    1. Bengsch B, Johnson AL, Kurachi M, Odorizzi PM, Pauken KE, Attanasio J, et al.. Bioenergetic Insufficiencies Due to Metabolic Alterations Regulated by the Inhibitory Receptor PD-1 Are an Early Driver of CD8(+) T Cell Exhaustion. Immunity. Immunity; 2016;45: 358–73. doi: 10.1016/j.immuni.2016.07.008
    1. Fultang L, Booth S, Yogev O, Martins da Costa B, Tubb V, Panetti S, et al.. Metabolic engineering against the arginine microenvironment enhances CAR-T cell proliferation and therapeutic activity. Blood. Content Repository Only!; 2020;136: 1155–1160. doi: 10.1182/blood.2019004500
    1. Rostamian H, Khakpoor-Koosheh M, Fallah-Mehrjardi K, Mirzaei HR, Brown CE. Mitochondria as Playmakers of CAR T-cell Fate and Longevity. Cancer Immunol Res. Cancer Immunol Res; 2021;9: 856–861. doi: 10.1158/2326-6066.CIR-21-0110
    1. Chang C-H, Qiu J, O’Sullivan D, Buck MD, Noguchi T, Curtis JD, et al.. Metabolic Competition in the Tumor Microenvironment Is a Driver of Cancer Progression. Cell. Cell; 2015;162: 1229–41. doi: 10.1016/j.cell.2015.08.016
    1. Ma EH, Bantug G, Griss T, Condotta S, Johnson RM, Samborska B, et al.. Serine Is an Essential Metabolite for Effector T Cell Expansion. Cell Metab. Cell Metab; 2017;25: 345–357. doi: 10.1016/j.cmet.2016.12.011
    1. Johnson MO, Wolf MM, Madden MZ, Andrejeva G, Sugiura A, Contreras DC, et al.. Distinct Regulation of Th17 and Th1 Cell Differentiation by Glutaminase-Dependent Metabolism. Cell. Cell; 2018;175: 1780–1795.e19. doi: 10.1016/j.cell.2018.10.001
    1. Klysz D, Tai X, Robert PA, Craveiro M, Cretenet G, Oburoglu L, et al.. Glutamine-dependent α-ketoglutarate production regulates the balance between T helper 1 cell and regulatory T cell generation. Sci Signal. Sci Signal; 2015;8: ra97. doi: 10.1126/scisignal.aab2610
    1. Geiger R, Rieckmann JC, Wolf T, Basso C, Feng Y, Fuhrer T, et al.. L-Arginine Modulates T Cell Metabolism and Enhances Survival and Anti-tumor Activity. Cell. Cell; 2016;167: 829–842.e13. doi: 10.1016/j.cell.2016.09.031
    1. Son J, Lyssiotis CA, Ying H, Wang X, Hua S, Ligorio M, et al.. Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature. Nature; 2013;496: 101–5. doi: 10.1038/nature12040
    1. van Karnebeek CDM, Ramos RJ, Wen X-Y, Tarailo-Graovac M, Gleeson JG, Skrypnyk C, et al.. Bi-allelic GOT2 Mutations Cause a Treatable Malate-Aspartate Shuttle-Related Encephalopathy. Am J Hum Genet. Am J Hum Genet; 2019;105: 534–548. doi: 10.1016/j.ajhg.2019.07.015
    1. Bailis W, Shyer JA, Zhao J, Canaveras JCG, Al Khazal FJ, Qu R, et al.. Distinct modes of mitochondrial metabolism uncouple T cell differentiation and function. Nature. Nature; 2019;571: 403–407. doi: 10.1038/s41586-019-1311-3
    1. Chisolm DA, Savic D, Moore AJ, Ballesteros-Tato A, León B, Crossman DK, et al.. CCCTC-Binding Factor Translates Interleukin 2- and α-Ketoglutarate-Sensitive Metabolic Changes in T Cells into Context-Dependent Gene Programs. Immunity. NIH Public Access; 2017;47: 251–267.e7. doi: 10.1016/j.immuni.2017.07.015
    1. Hiemer S, Jatav S, Jussif J, Alley J, Lathwal S, Piotrowski M, et al.. Integrated Metabolomic and Transcriptomic Profiling Reveals Novel Activation-Induced Metabolic Networks in Human T Cells. SSRN Electron J. 2019; doi: 10.2139/ssrn.3473264
    1. Bettonville M, d’Aria S, Weatherly K, Porporato PE, Zhang J, Bousbata S, et al.. Long-term antigen exposure irreversibly modifies metabolic requirements for T cell function. Elife. 2018;7. doi: 10.7554/eLife.30938
    1. Ho M, Kim H. Glypican-3: a new target for cancer immunotherapy. Eur J Cancer. Eur J Cancer; 2011;47: 333–8. doi: 10.1016/j.ejca.2010.10.024
    1. Baumhoer D, Tornillo L, Stadlmann S, Roncalli M, Diamantis EK, Terracciano LM. Glypican 3 expression in human nonneoplastic, preneoplastic, and neoplastic tissues: a tissue microarray analysis of 4,387 tissue samples. Am J Clin Pathol. Am J Clin Pathol; 2008;129: 899–906. doi: 10.1309/HCQWPWD50XHD2DW6
    1. Ishiguro T, Sano Y, Komatsu SI, Kamata-Sakurai M, Kaneko A, Kinoshita Y, et al.. An anti-glypican 3/CD3 bispecific T cell-redirecting antibody for treatment of solid tumors. Sci Transl Med. Sci Transl Med; 2017;9. doi: 10.1126/scitranslmed.aal4291
    1. Moek KL, Fehrmann RSN, van der Vegt B, de Vries EGE, de Groot DJA. Glypican 3 Overexpression across a Broad Spectrum of Tumor Types Discovered with Functional Genomic mRNA Profiling of a Large Cancer Database. Am J Pathol. Am J Pathol; 2018;188: 1973–1981. doi: 10.1016/j.ajpath.2018.05.014
    1. Lu C, Rong D, Zhang B, Zheng W, Wang X, Chen Z, et al.. Current perspectives on the immunosuppressive tumor microenvironment in hepatocellular carcinoma: challenges and opportunities. Mol Cancer. BioMed Central; 2019;18: 130. doi: 10.1186/s12943-019-1047-6
    1. Wang X, He Q, Shen H, Xia A, Tian W, Yu W, et al.. TOX promotes the exhaustion of antitumor CD8+ T cells by preventing PD1 degradation in hepatocellular carcinoma. J Hepatol. J Hepatol; 2019;71: 731–741. doi: 10.1016/j.jhep.2019.05.015
    1. Engels B, Cam H, Schüler T, Indraccolo S, Gladow M, Baum C, et al.. Retroviral vectors for high-level transgene expression in T lymphocytes. Hum Gene Ther. Hum Gene Ther; 2003;14: 1155–68. doi: 10.1089/104303403322167993
    1. Cerami E, Gao J, Dogrusoz U, Gross BE, Sumer SO, Aksoy BA, et al.. The cBio cancer genomics portal: an open platform for exploring multidimensional cancer genomics data. Cancer Discov. Cancer Discov; 2012;2: 401–4. doi: 10.1158/-12-0095
    1. Gao J, Aksoy BA, Dogrusoz U, Dresdner G, Gross B, Sumer SO, et al.. Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci Signal. Sci Signal; 2013;6: pl1. doi: 10.1126/scisignal.2004088
    1. Al-saraireh Y, Alrawashdeh F, Al-shuneigat J, Alsbou M, Alnawaiseh N, Al-shagahin H, et al.. Screening of Glypican- 3 Expression in Human Normal versus Benign and Malignant Tissues: A Comparative Study Glypican- 3 expression in cancers. Biosci Biotechnol Res Asia. 2016;13: 687–692. doi: 10.13005/bbra/2085
    1. Seo H, Chen J, González-Avalos E, Samaniego-Castruita D, Das A, Wang YH, et al.. TOX and TOX2 transcription factors cooperate with NR4A transcription factors to impose CD8+ T cell exhaustion. Proc Natl Acad Sci U S A. National Academy of Sciences; 2019;116: 12410–12415. doi: 10.1073/pnas.1905675116
    1. Khan O, Giles JR, McDonald S, Manne S, Ngiow SF, Patel KP, et al.. TOX transcriptionally and epigenetically programs CD8+ T cell exhaustion. Nature. Nature Publishing Group; 2019;571: 211–218. doi: 10.1038/s41586-019-1325-x
    1. Gerlach C, Moseman EA, Loughhead SM, Garg R, De La Torre JC, Von UH, et al.. The Chemokine Receptor CX3CR1 Defines Three Antigen-Experienced CD8 T Cell Subsets with Distinct Roles in Immune Surveillance and Homeostasis Article The Chemokine Receptor CX3CR1 Defines Three Antigen-Experienced CD8 T Cell Subsets with Distinct Roles in Immune Surveillance and Homeostasis. Immunity. 2016;45: 1270–1284. doi: 10.1016/j.immuni.2016.10.018
    1. Majchrzak K, Nelson MH, Bailey SR, Bowers JS, Yu X-Z, Rubinstein MP, et al.. Exploiting IL-17-producing CD4+ and CD8+ T cells to improve cancer immunotherapy in the clinic. Cancer Immunol Immunother. NIH Public Access; 2016;65: 247. doi: 10.1007/s00262-016-1797-6
    1. Araki K, Turner AP, Shaffer VO, Gangappa S, Keller SA, Bachmann MF, et al.. mTOR regulates memory CD8 T-cell differentiation. Nature. Nature; 2009;460: 108–12. doi: 10.1038/nature08155
    1. Alizadeh D, Wong RA, Yang X, Wang D, Pecoraro JR, Kuo C-F, et al.. IL15 Enhances CAR-T Cell Antitumor Activity by Reducing mTORC1 Activity and Preserving Their Stem Cell Memory Phenotype. Cancer Immunol Res. American Association for Cancer Research; 2019;7: 759–772. doi: 10.1158/2326-6066.CIR-18-0466
    1. Zhao Z, Condomines M, van der Stegen SJC, Perna F, Kloss CC, Gunset G, et al.. Structural Design of Engineered Costimulation Determines Tumor Rejection Kinetics and Persistence of CAR T Cells. Cancer Cell. Cancer Cell; 2015;28: 415–428. doi: 10.1016/j.ccell.2015.09.004
    1. Scharping NE, Menk AV, Moreci RS, Whetstone RD, Dadey RE, Watkins SC, et al.. The Tumor Microenvironment Represses T Cell Mitochondrial Biogenesis to Drive Intratumoral T Cell Metabolic Insufficiency and Dysfunction. Immunity. Immunity; 2016;45: 374–88. doi: 10.1016/j.immuni.2016.07.009
    1. Patsoukis N, Bardhan K, Chatterjee P, Sari D, Liu B, Bell LN, et al.. PD-1 alters T-cell metabolic reprogramming by inhibiting glycolysis and promoting lipolysis and fatty acid oxidation. Nat Commun. Nat Commun; 2015;6: 6692. doi: 10.1038/ncomms7692
    1. Macintyre AN, Gerriets VA, Nichols AG, Michalek RD, Rudolph MC, Deoliveira D, et al.. The glucose transporter Glut1 is selectively essential for CD4 T cell activation and effector function. Cell Metab. NIH Public Access; 2014;20: 61–72. doi: 10.1016/j.cmet.2014.05.004
    1. Ron-Harel N, Ghergurovich JM, Notarangelo G, LaFleur MW, Tsubosaka Y, Sharpe AH, et al.. T Cell Activation Depends on Extracellular Alanine. Cell Rep. Cell Rep; 2019;28: 3011–3021.e4. doi: 10.1016/j.celrep.2019.08.034
    1. Xu T, Stewart KM, Wang X, Liu K, Xie M, Ryu JK, et al.. Metabolic control of TH17 and induced Treg cell balance by an epigenetic mechanism. Nature. Nature; 2017;548: 228–233. doi: 10.1038/nature23475

Source: PubMed

3
Předplatit