Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Metabolic conditioning of CD8+ effector T cells for adoptive cell therapy

Abstract

CD8+ effector T (TE) cell proliferation and cytokine production depends on enhanced glucose metabolism. However, circulating T cells continuously adapt to glucose fluctuations caused by diet and inter-organ metabolite exchange. Here we show that transient glucose restriction (TGR) in activated CD8+ TE cells metabolically primes effector functions and enhances tumour clearance in mice. Tumour-specific TGR CD8+ TE cells co-cultured with tumour spheroids in replete conditions display enhanced effector molecule expression, and adoptive transfer of these cells in a murine lymphoma model leads to greater numbers of immunologically functional circulating donor cells and complete tumour clearance. Mechanistically, TE cells treated with TGR undergo metabolic remodelling that, after glucose re-exposure, supports enhanced glucose uptake, increased carbon allocation to the pentose phosphate pathway (PPP) and a cellular redox shift towards a more reduced state—all indicators of a more anabolic programme to support their enhanced functionality. Thus, metabolic conditioning could be used to promote efficiency of T-cell products for adoptive cellular therapy.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: TGR enhances CD8+ TE-cell molecule expression after glucose re-exposure.
Fig. 2: TGR TE cells have depleted sugar metabolism but sustain TCA cycle metabolites.
Fig. 3: TGR CD8+ TE cells are metabolically active and energetically balanced.
Fig. 4: TGR TE cells have reversible redox balance without accumulating cellular ROS.
Fig. 5: TGR CD8+ TE cells have enhanced anabolic metabolism after glucose re-exposure.
Fig. 6: TGR enhances tumour-specific CD8+ TE-cell function in vivo.

Similar content being viewed by others

Data availability

The data that support the findings of this study, as well as further information and requests for resources and reagents, will be made available on reasonable request by the corresponding author. scRNA-seq data have been deposited in the Gene Ontology Omnibus (accession no. GSE152018). Source data are provided with this paper.

References

  1. Michalek, R. D. et al. Cutting edge: distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T-cell subsets. J. Immunol. 186, 3299–3303 (2011).

    Article  CAS  PubMed  Google Scholar 

  2. Maciver, N. J. et al. Glucose metabolism in lymphocytes is a regulated process with significant effects on immune cell function and survival. J. Leukoc. Biol. 84, 949–957 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Chang, C.-H. et al. Metabolic competition in the tumor microenvironment is a driver of cancer progression. Cell 162, 1229–1241 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Chang, C.-H. et al. Post-transcriptional control of T-cell effector function by aerobic glycolysis. Cell 153, 1239–1251 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Blagih, J. et al. The energy sensor AMPK regulates T-cell metabolic adaptation and effector responses in vivo. Immunity 42, 41–54 (2015).

    Article  CAS  PubMed  Google Scholar 

  6. Rolf, J. et al. AMPKalpha1: a glucose sensor that controls CD8 T-cell memory. Eur. J. Immunol. 43, 889–896 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Klein Geltink, R. I. et al. Mitochondrial Priming by CD28. Cell 171, 385–397 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Buck, M. D. et al. Mitochondrial dynamics controls T-cell fate through metabolic programming. Cell 166, 63–76 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Sukumar, M. et al. Inhibiting glycolytic metabolism enhances CD8+ T-cell memory and anti-tumour function. J. Clin. Invest. 123, 4479–4488 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Sukumar, M. et al. Mitochondrial membrane potential identifies cells with enhanced stemness for cellular therapy. Cell Metab. 23, 63–76 (2016).

    Article  CAS  PubMed  Google Scholar 

  11. Scharping, N. E. et al. The tumor microenvironment represses T cell mitochondrial biogenesis to drive intratumoral T cell metabolic insufficiency and dysfunction. Immunity 45, 374–388 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Chowdhury, P. S., Chamoto, K., Kumar, A. & Honjo, T. PPAR-induced fatty acid oxidation in T cells increases the number of tumour-reactive CD8. Cancer Immunol. Res. 6, 1375–1387 (2018).

    Article  CAS  PubMed  Google Scholar 

  13. Buck, M. D., Sowell, R. T., Kaech, S. M. & Pearce, E. L. Metabolic instruction of immunity. Cell 169, 570–586 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Qiu, J. et al. Acetate promotes T-cell effector function during glucose restriction. Cell Rep. 27, 2063–2074 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Crompton, J. G., Sukumar, M. & Restifo, N. P. Targeting Akt in cell-transfer immunotherapy for cancer. Oncoimmunology 5, e1014776 (2016).

    Article  PubMed  CAS  Google Scholar 

  16. Nabe, S. et al. Reinforce the anti-tumour activity of CD8+ T cells via glutamine restriction. Cancer Sci. 109, 3737–3750 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Hosios, A. M. et al. Amino acids rather than glucose account for the majority of cell mass in proliferating mammalian cells. Dev. Cell 36, 540–549 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Garcia, D. & Shaw, R. J. AMPK: mechanisms of cellular energy sensing and restoration of metabolic balance. Mol. Cell 66, 789–800 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Saxton, R. A. & Sabatini, D. M. mTOR signaling in growth, metabolism, and disease. Cell 168, 960–976 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Jacobs, S. R. et al. Glucose uptake is limiting in T-cell activation and requires CD28-mediated Akt-dependent and independent pathways. J. Immunol. 180, 4476–4486 (2008).

    Article  CAS  PubMed  Google Scholar 

  21. Macintyre, A. N. et al. The glucose transporter Glut1 is selectively essential for CD4 T cell activation and effector function. Cell Metab. 20, 61–72 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Siska, P. J. et al. Suppression of Glut1 and glucose metabolism by decreased Akt/mTORC1 signaling drives T-cell impairment in B cell leukemia. J. Immunol. 197, 2532–2540 (2016).

    Article  CAS  PubMed  Google Scholar 

  23. Waypa, G. B. et al. Hypoxia triggers subcellular compartmental redox signaling in vascular smooth muscle cells. Circ. Res. 106, 526–535 (2010).

    Article  CAS  PubMed  Google Scholar 

  24. Schieber, M. & Chandel, N. S. ROS function in redox signaling and oxidative stress. Curr. Biol. 24, R453–R462 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Hosios, A. M. & Vander Heiden, M. G. The redox requirements of proliferating mammalian cells. J. Biol. Chem. 293, 7490–7498 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Kong, H. & Chandel, N. S. Regulation of redox balance in cancer and T cells. J. Biol. Chem. 293, 7499–7507 (2018).

    Article  CAS  PubMed  Google Scholar 

  27. Mak, T. W. et al. Glutathione primes T-cell metabolism for inflammation. Immunity 46, 675–689 (2017).

    Article  CAS  PubMed  Google Scholar 

  28. Smith, C. A., Want, E. J., O’Maille, G., Abagyan, R. & Siuzdak, G. XCMS: processing mass spectrometry data for metabolite profiling using nonlinear peak alignment, matching and identification. Anal. Chem. 78, 779–787 (2006).

    Article  CAS  PubMed  Google Scholar 

  29. Huang, X. et al. X13CMS: global tracking of isotopic labels in untargeted metabolomics. Anal. Chem. 86, 1632–1639 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Ma, E. H. et al. Serine is an essential metabolite for effector T-cell expansion. Cell Metab. 25, 345–357 (2017).

    Article  CAS  PubMed  Google Scholar 

  31. Crompton, J. G. et al. Akt inhibition enhances expansion of potent tumour-specific lymphocytes with memory-cell characteristics. Cancer Res. 75, 296–305 (2015).

    Article  CAS  PubMed  Google Scholar 

  32. Vander Heiden, M. G., Cantley, L. C. & Thompson, C. B. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324, 1029–1033 (2009).

    Article  CAS  Google Scholar 

  33. Warburg, O. The metabolism of carcinoma cells. J. Cancer Res. 9, 148–163 (1925).

    Article  CAS  Google Scholar 

  34. Liberti, M. V. & Locasale, J. W. The Warburg effect: how does it benefit cancer cells?. Trends Biochem. Sci. 41, 211–218 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Zhao, J., Song, Y. & Liu, D. Clinical trials of dual-target CAR T cells, donor-derived CAR T cells, and universal CAR T cells for acute lymphoid leukemia. J. Hematol. Oncol. 12, 17 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Sena, L. A. et al. Mitochondria are required for antigen-specific T-cell activation through reactive oxygen species signaling. Immunity 38, 225–236 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Nathan, C. & Cunningham-Bussel, A. Beyond oxidative stress: an immunologist’s guide to reactive oxygen species. Nat. Rev. Immunol. 13, 349–361 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Chatterjee, S. et al. CD38NAD+ axis regulates immunotherapeutic anti-tumour T-cell response. Cell Metab. 27, 85–100 (2018).

    Article  CAS  PubMed  Google Scholar 

  39. Johnson, M. O. et al. Distinct regulation of Th17 and Th1 cell differentiation by glutaminase-dependent metabolism. Cell 175, 1780–1795 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Ghergurovich, J. M. et al. A small-molecule G6PD inhibitor reveals immune dependence on pentose phosphate pathway. Nat. Chem. Biol. 16, 731–739 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ma, E. H. et al. Metabolic profiling using stable isotope tracing reveals distinct patterns of glucose utilization by physiologically activated CD8. Immunity 51, 856–870 (2019).

    Article  CAS  PubMed  Google Scholar 

  42. Yang, J. C. & Rosenberg, S. A. Adoptive T-cell therapy for cancer. Adv. Immunol. 130, 279–294 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Chamoto, K. et al. Mitochondrial activation chemicals synergize with surface receptor PD-1 blockade for T-cell-dependent anti-tumour activity. Proc. Natl Acad. Sci. USA 114, E761–E770 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Kumar, A., Chamoto, K., Chowdhury, P. S. & Honjo, T. Tumours attenuating the mitochondrial activity in T cells escape from PD-1 blockade therapy. eLife 9, e52330 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Wills, J., Edwards-Hicks, J. & Finch, A. J. AssayR: a simple mass spectrometry software tool for targeted metabolic and stable isotope tracer analyses. Anal. Chem. 89, 9616–9619 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. van der Windt, G. J. W., Chang, C. H. & Pearce, E. L. Measuring bioenergetics in T cells using a Seahorse extracellular flux analyser. Curr. Protoc. Immunol. 113, 3.16B.1–3.16B.14 (2016).

    Google Scholar 

  47. Pearce, E. L. et al. Enhancing CD8 T-cell memory by modulating fatty acid metabolism. Nature 460, 103–107 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Butler, A., Hoffman, P., Smibert, P., Papalexi, E. & Satija, R. Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nat. Biotechnol. 36, 411–420 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Diaz-Papkovich, A., Anderson-Trocmé, L., Ben-Eghan, C. & Gravel, S. UMAP reveals cryptic population structure and phenotype heterogeneity in large genomic cohorts. PLoS Genet. 15, e1008432 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Aoki-Kinoshita, K. F. & Kanehisa, M. Gene annotation and pathway mapping in KEGG. Methods Mol. Biol. 396, 71–91 (2007).

    Article  CAS  PubMed  Google Scholar 

  51. Szklarczyk, D. et al. STRING v11: protein–protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 47, D607–D613 (2019).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank members of the Pearce laboratories for support and helpful discussions and A. Quintana and J. Sutherland mouse colony management. This work was funded by the National Institutes of Health (NIH; CA181125 to E.L.P. and AI110481 to E.J.P.) and the Max Planck Society. R.Z. was supported by the Deutsche Forschungsgemeinschaft (DFG; SFB1160, B09; TRR167, B06) and the European Research Council (ERC; GvHDCure no. 681012).

Author information

Authors and Affiliations

Authors

Contributions

R.I.K.G., J.E.-H., P.A., D.O., D.E.S., D.J.P., J.M.B., K.M.G., A.M.K., M.S., F.M.U., M.F., R.Z., E.J.P. and E.L.P. designed the research, analysed data and provided conceptual input. A.E.P., N.A.M.L., J.D.C. and F.H. analysed data and performed the experiments. R.I.K.G., J.E.H. and E.L.P. wrote the manuscript.

Corresponding author

Correspondence to Erika L. Pearce.

Ethics declarations

Competing interests

E.L.P. is a SAB member of ImmunoMet Therapeutics, and E.L.P. and E.J.P. are founders of Rheos Medicines. The other authors declare no competing interests.

Additional information

Peer review information Primary Handling Editor: Christoph Schmitt.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 TGR CD8+ TE do not sustain mitochondrial metabolism by incorporating more glutamine-derived carbons into the TCA cycle.

WT CD8+ T cells isolated from spleens of C7Bl/6 mice were activated with anti-CD3 (5 μg/mL), anti-CD28 (0.5 μg/mL), IL-2 (100 U/mL), expanded for a total of 72 h, and exposed to 10 mM, 3 mM, or 1 mM glucose as indicated in cultures set to 1 million per ml. a, Polar metabolites were extracted from 500 µl media supernatant from 20 h cultures of 10 mM and 1 mM cells. Bar graphs represent glutamine concentration in the indicated conditions compared to complete culture media for n = 3 biological replicates. Significance was calculated using 2-tailed Student’s t tests, no significant changes (ns) were observed. b, TE were generated as above, but during the final 20 h all groups were cultured in 4 mM heavy labelled (U-13C) glutamine (100% U-13C glutamine). Polar metabolites were extracted and isotopologue distribution assessed by targeted mass spectrometry. TCA intermediates are plotted as percent label from newly metabolized U-13C (open bars) or remaining U-12C (black bars) glutamine carbons. Data are from n = 3 biological replicates, representative of 2 independent experiments. Significance was calculated using 2-tailed Student’s t tests, no significant changes were observed. c, TE were generated as above, but during the final 6 h 2 mM U-13C-lactate was added to the culture. Polar metabolites were extracted and isotopologue distribution assessed by targeted mass spectrometry. Significance was calculated using 2-tailed Student’s t tests comparing the predominant m + 2 isotopologue group in both conditions. Data shown for n = 3 biological replicates. Significance was calculated using 2-tailed Student’s t tests, No significant changes were observed. All error bars show SEM.

Source data

Extended Data Fig. 2 Limited transcriptional reprogramming during TGR.

WT CD8+ T cells isolated from spleens of C57BL/6 mice were activated with anti-CD3 (5 μg/mL), anti-CD28 (0.5 μg/mL), IL-2 (100 U/mL), expanded for a total of 72 h, and exposed to 10 mM (black) or 1 mM (orange) glucose in cultures set to 1 million per ml. RNA isolated from single cells was sequenced using the 10X genomics platform and analysed to explore transcriptional changes in each population of cells. a, A Uniform Manifold Approximation and Projection (UMAP) of the overlapped cells from each treatment (control, black; TGR, orange), clustered on the basis of transcriptional similarity is shown highlighting clusters and treatment distribution. Bar graphs depicting cluster distribution for each condition are also shown. b, Clusters 0 and 1, where least overlap between treatments was observed, were contrasted to look for differentially expressed genes (> 1.2-fold change and < 0.1 adjusted p-value); a heatmap of top up and down regulated genes is shown, along with UMAPs of example genes. c, The top 5 most enriched pathways in differentially regulated genes between cluster 0 and 1. d, Module scores based on average expression levels of gene programs were calculated for 18 differentially expressed OXPHOS genes (Atp5g1, Atp5g3, Cox5a, Cox6a1, Cyc1, Ndufa11, Ndufa12, Ndufa4, Ndufa8, Ndufab1, Ndufb7, Ndufb8, Ndufc1, Ndufc2, Ndufs6, Sdhb, Uqcr10, Uqcr11). Violin plots depict the global OXPHOS module score per condition (left) or per cluster and condition (right). e, Module scores based on average expression levels of gene programs were also calculated for enzyme-coding genes involved in glycolysis (mmu00010) or the pentose phosphate pathway (PPP, mmu00030) based on Kegg annotation. UMAP and Violin plots illustrate the OXPHOS, glycolysis and PPP modules, with the violin plots divided per cluster.

Extended Data Fig. 3 Signalling timecourse during TGR.

Protein isolates were prepared as described in (Fig. 3f), taking samples every 2 h over the 20 h exposure to limiting glucose concentration. Immunoblot analysis of protein extracts from equal cell numbers probed for phosphorylated acetyl-coA carboxylase at Ser79 (p-ACC1Thr79), total ACC1, phosphorylated AMPK at Thr172 (p-AMPkThr172), total AMPK, phosphorylated ribosomal protein S6 at Ser235/236 (p-S6Ser235/236), total S6, phosphorylated 4E-binding protein 1 at Thr37/46 (p-4E-BP1Thr37/46), total 4E-BP1, and Glut1. Tubulin was used as a loading control. Representative of 3 biological independent samples. Biological replicate data is shown in Fig. 3f.

Source data

Extended Data Fig. 4 TGR TE have altered glucose reallocation upon re-exposure to drive anabolic metabolism.

WT CD8+ T cells isolated from spleens of C57BL/6 mice were activated with anti-CD3 (5 μg/mL), anti-CD28 (0.5 μg/mL), IL-2 (100 U/mL), expanded for a total of 72 h, and exposed to 10 mM (control) or 1 mM (TGR) glucose during the final 5 min cells were re-exposed to normal (U-12C) glucose or heavy labelled (U-13C) glucose. 20 h cultures were started at 1 × 106 (control) and 1.5 × 106 (TGR) cells per/ml to generate similar end concentrations for glucose pulse experiments. a, Table showing the results of a Kegg-pathway analysis of the significantly increased and decreased metabolite pools in TGR TE after normal (U-12C) glucose re-feeding when compared to re-fed control TE. b, Table shows the results of a Kegg pathway analysis of the metabolites that exhibited increased glucose-derived U-13C assimilation (as determined by X13 CMS software) after re-feeding TGR TE with heavy labelled (U-13C) glucose compared to re-fed control TE. Data are from 3 biological replicates, representative of 3 independent experiments. Kegg Pathways were ranked according to the number of metabolite hits in that pathway (from highest to lowest). Metabolites in the highest ranked pathways were further validated by targeted analysis. c, Polar metabolites were extracted and untargeted metabolomic analysis was performed using XCMS on cells re-exposed to normal (U-12C) glucose. Volcano plot depicts relative metabolite pools compared between control and TGR cells. Orange circle represents a pentose phosphate pathway metabolite Pentose-phosphate, significantly increased in glucose re-fed TGR TE. Statistical significance was calculated using a Welch’s t-test, comparing relative intensities of each isotopologue in labelled samples of control TE versus those of TGR TE.

Source data

Extended Data Fig. 5 6h U-13C-glucose re-exposure following TGR.

WT CD8+ T cells isolated from spleens of C57BL/6 mice were activated with anti-CD3 (5 μg/mL), anti-CD28 (0.5 μg/mL), IL-2 (100 U/mL), expanded for a total of 72 h, and exposed to 10 mM (black) or 1 mM (orange) glucose. Cells were re-exposed to 10 mM U-13C-glucose for 6 h prior to polar metabolite extraction and analysis by LC-MS. 20 h cultures were started at 1 × 106 (control) and 1.5 × 106 (TGR) cells per/ml to generate similar end concentrations for glucose pulse experiments. a, Percent 13C label incorporation into nucleotide monophosphates was analysed. Data are from n = 3 biological replicates. Significance was calculated using 2-tailed Student’s t tests. * p < 0.05. b, Relative 13C label incorporation into serine was analysed. Data are from n = 3 biological replicates. Significance was calculated using 2way ANOVA. Significance is indicated for the 13C portion (open bars). *** p < 0.001. The 12C portion (filled bars) was also significant. ** p < 0.01. c, The GSH/GSSG ratio was calculated from summed isotopologues for each metabolite. Data are from n = 3 biological replicates. Significance was calculated using 2-tailed Student’s t tests, no significant difference was found. d, Relative 13C label incorporation into lactate was analysed. Data are from n = 3 biological replicates. Significance was calculated using 2way ANOVA. Significance is indicated for the 13C portion (open bars), which was not significant. The 12C portion (filled bars) was significant. *** p < 0.001. All error bars show SEM.

Source data

Extended Data Fig. 6 TGR enhances CD8+ tumour-specific antitumour function in vivo.

As depicted in Fig. 6f, congenically marked (CD45.1) C57BL/6 female recipient mice were injected subcutaneously with 1×106 B16-OVA melanoma cells and tumours established for 5 d. 5 ×106 control (black) or TGR (orange) OT-I+ TE were injected intravenously per tumour bearing mouse. Mice receiving no TE were used as controls. 3, 6 and 9 d after tumour inoculation, mice received 200 µg anti-PD-1 antibody or IgG2a isotype control in 100 µl PBS intraperitoneally. a, 21 d after tumour inoculation mice were humanely euthanized, blood was collected, red cells were lysed, and the white blood cell fraction stained for congenic markers. Data are presented as % of donor-derived (CD45.2+) CD8+ T cells as a fraction to total circulating CD8+ T cells and each dot represents an individual mouse (n = 5). Statistical significance was calculated by 2-tailed Student’s t test. * p < 0.05; ** p < 0.01. b, 21 d after tumour inoculation mice were humanely euthanized and the tumour mass was excised. Single cell suspensions were analysed by flow cytometry. The number of host (CD45.1) CD8+ T cells per mg tissue was quantified and each dot represents an individual mouse (n = 5). Statistical significance was calculated by 2-tailed Student’s t test. ns not significant. All error bars show SEM.

Source data

Supplementary information

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data and western blot optical density values.

Source Data Fig. 3

Uncropped western blots.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 3

Western blot optical density values.

Source Data Extended Data Fig. 3

Uncropped western blots.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Klein Geltink, R.I., Edwards-Hicks, J., Apostolova, P. et al. Metabolic conditioning of CD8+ effector T cells for adoptive cell therapy. Nat Metab 2, 703–716 (2020). https://doi.org/10.1038/s42255-020-0256-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-020-0256-z

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer