Cancer-cell stiffening via cholesterol depletion enhances adoptive T-cell immunotherapy

Kewen Lei, Armand Kurum, Murat Kaynak, Lucia Bonati, Yulong Han, Veronika Cencen, Min Gao, Yu-Qing Xie, Yugang Guo, Mélanie T M Hannebelle, Yangping Wu, Guanyu Zhou, Ming Guo, Georg E Fantner, Mahmut Selman Sakar, Li Tang, Kewen Lei, Armand Kurum, Murat Kaynak, Lucia Bonati, Yulong Han, Veronika Cencen, Min Gao, Yu-Qing Xie, Yugang Guo, Mélanie T M Hannebelle, Yangping Wu, Guanyu Zhou, Ming Guo, Georg E Fantner, Mahmut Selman Sakar, Li Tang

Abstract

Malignant transformation and tumour progression are associated with cancer-cell softening. Yet how the biomechanics of cancer cells affects T-cell-mediated cytotoxicity and thus the outcomes of adoptive T-cell immunotherapies is unknown. Here we show that T-cell-mediated cancer-cell killing is hampered for cortically soft cancer cells, which have plasma membranes enriched in cholesterol, and that cancer-cell stiffening via cholesterol depletion augments T-cell cytotoxicity and enhances the efficacy of adoptive T-cell therapy against solid tumours in mice. We also show that the enhanced cytotoxicity against stiffened cancer cells is mediated by augmented T-cell forces arising from an increased accumulation of filamentous actin at the immunological synapse, and that cancer-cell stiffening has negligible influence on: T-cell-receptor signalling, production of cytolytic proteins such as granzyme B, secretion of interferon gamma and tumour necrosis factor alpha, and Fas-receptor-Fas-ligand interactions. Our findings reveal a mechanical immune checkpoint that could be targeted therapeutically to improve the effectiveness of cancer immunotherapies.

Conflict of interest statement

Competing interests

The authors declare no competing interests.

© 2021. The Author(s), under exclusive licence to Springer Nature Limited.

Figures

Fig. 1. Cholesterol is enriched in the…
Fig. 1. Cholesterol is enriched in the plasma membrane of cancer cells.
a, B16F10 tumour tissues (indicated with dash lines) and the adjacent normal tissues were stained with hematoxylin and eosin (H&E) and Filipin III (shown in blue colour). Scale bar, 500 μm. b, Cholesterol levels in B16F10 tumour tissues and the adjacent skin and muscle tissues (n = 3). c, d, Membrane cholesterol levels of tumour-infiltrating leukocytes (CD45.2+) and cancer cells (tdTomato+) in 4T1-Fluc-tdTomato tumour (c), and murine T lymphoma cells (EG7-OVA) and normal murine T-cells (d) determined by Filipin III staining (n = 3). The displayed MFI values were normalized by forward scatter area (FSC-A) of corresponding samples. e, f, Relative intracellular and plasma membrane levels of cholesterol (normalized to that of native cells) in murine B16F10 and human Me275 cancer cells treated with water-soluble cholesterol/methyl-β-cyclodextrin complex (Chol) or methyl-β-cyclodextrin (MeβCD) in vitro (e, n = 3) and in vivo (f, B16F10, n = 5). Data are one representative of at least two independent experiments with biological replicates. P values were determined by unpaired Student’s t test. Error bars represent standard error of the mean (SEM). MFI, mean fluorescence intensity; n.s., not significant.
Fig. 2. Cancer-cell stiffness can be manipulated…
Fig. 2. Cancer-cell stiffness can be manipulated via the supplementation or depletion of cholesterol in the cell membrane.
a, Schematic illustration of the correlation between cellular stiffness and membrane cholesterol level. b, Relative cortical stiffness determined by nanoindentation measurements using atomic force microscopy (AFM) for native, Chol- or MeβCD-treated B16F10 cancer cells (n = 9 ~ 10 individual cells). Each data point is the average of at least twenty force curve measurements of a single cancer cell. Native B16F10 cancer cells serve as a standard (100%). Error bars represent SEM. c, Schematic illustration of the optical tweezer setting for cell cortical stiffness measurement. d, Cortical stiffness of native, Chol- or MeβCD-treated murine B16F10 and human Me275 cancer cells measured by the optical tweezer (n = 14 ~ 17 individual cells). e-g, Cellular deformation was measured using deformability cytometry to compare cellular stiffness in a high throughput manner. Shown are representative scatter plots (e; indicated are sample size, outliers not shown) and quantitative deformation of native, Chol- or MeβCD-treated murine B16F10 (f) and human Me275 (g) cancer cells. In all the violin plots (d, f, g), the middle solid line shows median, and lower and upper dash lines show 25th and 75th percentiles, respectively. P values were determined by unpaired Student’s t test. a.u., arbitrary unit; n.s., not significant.
Fig. 3. Cancer-cell softness impairs T-cell mediated…
Fig. 3. Cancer-cell softness impairs T-cell mediated cytotoxicity in vitro and in vivo.
a, Lysis percentage of B16F10 cancer cells pre-treated with Chol (softened) or PBS (native) and co-cultured with activated Pmel CD8+ T-cells at an effector:target (E:T) ratio of 10:1 for 5 h (n = 3). b, Relative membrane cholesterol levels of B16F10 cancer cells with ACAT1 knock-down (ACAT1 KD) and ACAT1 overexpression (ACAT1 OE) (n = 3). Native B16F10 cancer cells serve as a standard (100%). c, Cortical stiffness of native, ACAT1 KD, and ACAT1 OE B16F10 cells measured by the optical tweezer (n = 19 ~ 21 individual cells). In the violin plots, the middle solid line shows median, and lower and upper dash lines show 25th and 75th percentiles, respectively. d, Representative scatter plots for native, ACAT1 KD, and ACAT1 OE B16F10 cells by deformability cytometry analysis, and their 50%-density contour plots (the inner contours correspond to 95% event density) with iso-elasticity lines dividing the diagrams into areas of different stiffness. e, Lysis percentage of ACAT1 KD and ACAT1 OE B16F10 cancer cells after 5-h co-culture with Pmel CD8+ T-cells at an E:T ratio of 10:1 (n = 4). Data are one representative of at least two independent experiments with biological replicates (a, b, e). f-h, Mice bearing native, ACAT1 KD, or ACAT1 OE B16F10 tumours were treated with adoptive transfer of Pmel CD8+ T-cells (5 × 106 per mouse), as outlined in the experimental scheme (f) (n = 5 and 10 animals for PBS- and ACT-treated groups, respectively). Shown are tumour growth curves (g) and survival curves (h) of pooled data from two independent experiments with biological replicates. P values were determined by unpaired Student’s t test in (a-c, e), two-way ANOVA in (g), or log-rank test in (h). Error bars represent SEM. ACAT1, acyl-CoA:cholesterol acyltransferase 1; PBS, phosphate-buffered saline; ACT, adoptive cell transfer; s.c., subcutaneous; i.v., intravenous; n.s., not significant.
Fig. 4. Cancer-cell stiffening by MeβCD enhances…
Fig. 4. Cancer-cell stiffening by MeβCD enhances the efficacy of ACT immunotherapy.
a, Lysis percentage of B16F10 cancer cells pre-treated with MeβCD (stiffened) or PBS (native) and co-cultured with activated Pmel CD8+ T-cells at an E:T ratio of 10:1 for 5 h (n = 5). b, Lysis percentage of EG7-OVA cancer cells pre-treated with MeβCD (stiffened) or PBS (native) and co-cultured with activated OT-I CD8+ T-cells at indicated E:T ratios for 5 h (n = 3). Data in (a, b) are one representative of at least three independent experiments with biological replicates. c-e, B16F10 tumour-bearing mice were treated with adoptive transfer of Pmel CD8+ T-cells (5 × 106 per mouse) adjuvanted by interleukin-15 super-agonist (IL-15SA, 10 μg per injection) with or without daily MeβCD administration (1 mg per injection) as outlined in the experimental scheme (c). Mice receiving injections of PBS or MeβCD only serve as controls (n = 12 animals per group). Shown are survival curves (d), and individual tumour growth curves (e, indicated are the number of mice with durable responses out of all treated mice) of pooled data of two independent experiments with biological replicates. f-j, Tumour-infiltrating Pmel CD8+ T-cells were analysed by flow cytometry on day 14 (experimental scheme is shown in Supplementary Fig. 12b). Shown are counts (f), frequencies of granzyme B (GrzmB)+ (g), polyfunctional (h), and PD-1+ (j), and Ki67 expression level (i) of tumour-infiltrating Pmel CD8+ T-cells (n = 6 animals per group). Data are one representative of two independent experiments with biological replicates. P values were determined by unpaired Student’s t test in (a, b, f-j) or log-rank test in (d). Error bars represent SEM. PBS, phosphate-buffered saline; ACT, adoptive cell transfer; MFI, mean fluorescence intensity; s.c., subcutaneous; i.v., intravenous; i.t., intratumoural; n.s., not significant.
Fig. 5. Cancer-cell stiffening has negligible influence…
Fig. 5. Cancer-cell stiffening has negligible influence on biochemical cancer-cell killing pathways mediated by T-cells.
a, Schematic illustration of T-cell mediated killing pathways. b, c, Fold change of MFI of phosphorylated ZAP70 (pZAP70, b) and Erk1/2 (pErk1/2, c) in activated Pmel CD8+ T-cells stimulated by native or MeβCD-treated (stiffened) B16F10 cancer cells at 37 °C for 5 min (n = 6). d, Expression levels of Fas of native and stiffened B16F10 cancer cells (n = 6). e-i, Activated Pmel CD8+ T-cells were co-cultured with native or stiffened B16F10 cancer cells (E:T ratio = 10:1) at 37 °C for 5 h. Shown are expression levels of Fas ligand (FasL) (e) and granzyme B (GrzmB) (i), and frequencies of IFN-γ+ (f), TNF-α+ (g), and CD107a+ (h) of Pmel CD8+ T-cells (n = 6). j, k, Fold change of frequencies of apoptotic native and stiffened B16F10 cancer cells after incubation with FasL (j) or TNF-α (k) at indicated concentrations at 37 °C for 5 h (n = 5). l, Viability of native and stiffened B16F10 cancer cells after incubation with perforin of indicated concentrations at 37 °C for 20 min (n = 3). P values were determined by unpaired Student’s t test. Error bars represent SEM. MFI, mean fluorescence intensity; n.s., not significant. All data are one representative of at least three independent experiments with biological replicates.
Fig. 6. Enhanced cytotoxicity against stiffened cancer…
Fig. 6. Enhanced cytotoxicity against stiffened cancer cells is mediated by T-cell forces.
a-c, Forces exerted by activated Pmel CD8+ T-cells on polyacrylamide (PA) hydrogel substrates of indicated stiffness coated with anti-CD3 and anti-CD28 antibodies were measured using traction force microscopy. Shown are representative bright field images (a) and the corresponding traction stress maps (b), and average total force per cell (c) (n = 29 individual cells). The colour bar indicates the magnitude of stress. Scale bar, 5 μm. d, Representative deconvoluted confocal fluorescence images of F-actin of activated Pmel CD8+ T-cells on PA hydrogel substrates of indicated stiffness coated with anti-CD3 and anti-CD28 antibodies. The upper row shows the side view (XZ plane); the lower row shows the top view (XY plane) of F-actin at the T-cell immunological synapse (IS, defined as the structure between the surface of hydrogel and a height of 2 μm above the surface of the hydrogel). The colour bar indicates the intensity of the F-actin fluorescence signal. Scale bar, 2 μm. e, Relative total fluorescence intensity (normalized by the mean value at 260 Pa) of F-actin at the IS in the images from (d) (n = 66, 119, and 179 individual cells for 260, 510 and 890 Pa, respectively). f, MFI of phosphorylated Pyk2 (pPyk2) in activated Pmel CD8+ T-cells co-cultured with native, Chol-treated (softened) or MeβCD-treated (stiffened) B16F10 cancer cells (n = 5). g, h, Lysis percentage of native and stiffened B16F10 cancer cells co-cultured with activated Pmel CD8+ T-cells (E:T ratio = 10:1), which were pre-treated with latrunculin A (LatA, g) or blebbistatin (Bleb, h) (n = 5). P values were determined by Kruskal-Wallis test in (c, e) or unpaired Student’s t test in (f-h). Error bars represent SEM. In the violin plots (c, e), the middle solid line shows median, and lower and upper dash lines show 25th and 75th percentiles, respectively. MFI, mean fluorescence intensity; n.s., not significant. All data are one representative of at least two independent experiments with biological replicates.
Fig. 7. Illustration of mechanical immuno-suppression induced…
Fig. 7. Illustration of mechanical immuno-suppression induced by the softness of cancer cells. Stiffening the cancer cells enhances cancer-cell killing by T-cells.

References

    1. Ribas A, Wolchok JD. Cancer immunotherapy using checkpoint blockade. Science. 2018;359:1350–1355.
    1. Wang H, Mooney DJ. Biomaterial-assisted targeted modulation of immune cells in cancer treatment. Nat Mater. 2018;17:761–772.
    1. Eil R, et al. Ionic immune suppression within the tumour microenvironment limits T cell effector function. Nature. 2016;537:539–543.
    1. André P, et al. Anti-NKG2A mAb is a checkpoint inhibitor that promotes anti-tumor immunity by unleashing both T and NK cells. Cell. 2018;175:1731–1743.
    1. Levental KR, et al. Matrix crosslinking forces tumor progression by enhancing integrin signaling. Cell. 2009;139:891–906.
    1. Cross SE, Jin YS, Rao J, Gimzewski JK. Nanomechanical analysis of cells from cancer patients. Nat Nanotechnol. 2007;2:780–783.
    1. Fritsch A, et al. Are biomechanical changes necessary for tumour progression? Nat Phys. 2010;6:730–732.
    1. Swaminathan V, et al. Mechanical Stiffness grades metastatic potential in patient tumor cells and in cancer cell lines. Cancer Res. 2011;71:5075–5080.
    1. Alibert C, Goud B, Manneville JB. Are cancer cells really softer than normal cells? Biol Cell. 2017;109:167–189.
    1. Händel C, et al. Cell membrane softening in human breast and cervical cancer cells. New J Phys. 2015;17:083008
    1. Köster DV, Mayor S. Cortical actin and the plasma membrane: Inextricably intertwined. Curr Opin Cell Biol. 2016;38:81–89.
    1. Mossman KD, Campi G, Groves JT, Dustin ML. Altered TCR signaling from geometrically repatterned immunological synapses. Science. 2005;310:1191–1193.
    1. Judokusumo E, Tabdanov E, Kumari S, Dustin ML, Kam LC. Mechanosensing in T lymphocyte activation. Biophys J. 2012;102:L5–L7.
    1. Liu B, Chen W, Evavold BD, Zhu C. Accumulation of dynamic catch bonds between TCR and agonist peptide-MHC triggers T cell signaling. Cell. 2014;157:357–368.
    1. Pan Y, et al. Mechanogenetics for the remote and noninvasive control of cancer immunotherapy. Proc Natl Acad Sci U S A. 2018;115:992–997.
    1. Hickey JW, et al. Engineering an artificial T-cell stimulating matrix for immunotherapy. Adv Mater. 2019;31:1807359
    1. Meng KP, Majedi FS, Thauland TJ, Butte MJ. Mechanosensing through YAP controls T cell activation and metabolism. J Exp Med. 2020;217:e20200053.
    1. Basu R, et al. Cytotoxic T cells use mechanical force to potentiate target cell killing. Cell. 2016;165:100–110.
    1. Hui KL, Balagopalan L, Samelson LE, Upadhyaya A. Cytoskeletal forces during signaling activation in Jurkat T-cells. Mol Biol Cell. 2014;26:685–695.
    1. Saitakis M, et al. Different TCR-induced T lymphocyte responses are potentiated by stiffness with variable sensitivity. Elife. 2017;6:e23190.
    1. Byfield FJ, Aranda-Espinoza H, Romanenko VG, Rothblat GH, Levitan I. Cholesterol depletion increases membrane stiffness of aortic endothelial cells. Biophys J. 2004;87:3336–3343.
    1. Khatibzadeh N, Gupta S, Farrell B, Brownell WE, Anvari B. Effects of cholesterol on nanomechanical properties of the living cell plasma membrane. Soft Matter. 2012;8:8350–8360.
    1. Behnke O, Tranum-Jensen J, Van Deurs B. Filipin as a cholesterol probe. II. Filipin-cholesterol interaction in red blood cell membranes. Eur J Cell Biol. 1984;35:200–215.
    1. Jambhekar SS, Breen P. Cyclodextrins in pharmaceutical formulations I: Structure and physicochemical properties, formation of complexes, and types of complex. Drug Discov Today. 2016;21:356–362.
    1. Zidovetzki R, Levitan I. Use of cyclodextrins to manipulate plasma membrane cholesterol content: Evidence, misconceptions and control strategies. Biochim Biophys Acta - Biomembr. 2007;1768:1311–1324.
    1. Vahabikashi A, et al. Probe sensitivity to cortical versus intracellular cytoskeletal network stiffness. Biophys J. 2019;116:518–529.
    1. Guo M, et al. Cell volume change through water efflux impacts cell stiffness and stem cell fate. Proc Natl Acad Sci U S A. 2017;114:E8618–E8627.
    1. Otto O, et al. Real-time deformability cytometry: On-the-fly cell mechanical phenotyping. Nat Methods. 2015;12:199–202.
    1. Azadi S, Tafazzoli-Shadpour M, Soleimani M, Warkiani ME. Modulating cancer cell mechanics and actin cytoskeleton structure by chemical and mechanical stimulations. J Biomed Mater Res Part A. 2019;107A:1569–1581.
    1. Lekka M. Discrimination between normal and cancerous cells using AFM. Bionanoscience. 2016;6:65–80.
    1. Butcher DT, Alliston T, Weaver VM. A tense situation: Forcing tumour progression. Nat Rev Cancer. 2009;9:108–122.
    1. Zhu M, et al. ACAT1 regulates the dynamics of free cholesterols in plasma membrane which leads to the APP-α-processing alteration. Acta Biochim Biophys Sin. 2015;47:951–959.
    1. Yang W, et al. Potentiating the antitumour response of CD8+ T cells by modulating cholesterol metabolism. Nature. 2016;531:651–655.
    1. Barry M, Bleackley RC. Cytotoxic T lymphocytes: All roads lead to death. Nat Rev Immunol. 2002;2:401–409.
    1. Golstein P, Griffiths GM. An early history of T cell-mediated cytotoxicity. Nat Rev Immunol. 2018;18:527–535.
    1. Chow CW, Rincón M, Davis RJ. Requirement for transcription factor NFAT in interleukin-2 expression. Mol Cell Biol. 1999;19:2300–2307.
    1. Oeckinghaus A, Ghosh S. The NF-κB family of transcription factors and its regulation. Cold Spring Harb Perspect Biol. 2009;1:a000034.
    1. Style RW, et al. Traction force microscopy in physics and biology. Soft Matter. 2014;10:4047–4055.
    1. Weder G, et al. Increased plasticity of the stiffness of melanoma cells correlates with their acquisition of metastatic properties. Nanomed-Nanotechnol Biol Med. 2014;10:141–148.
    1. Luo Q, Kuang D, Zhang B, Song G. Cell stiffness determined by atomic force microscopy and its correlation with cell motility. Biochim Biophys Acta - Gen Subj. 2016;1860:1953–1960.
    1. Liu Y, et al. Cell softness prevents cytolytic T-cell killing of tumor-repopulating cells. Cancer Res. 2021;81:476–488.
    1. Blanchoin L, Boujemaa-Paterski R, Sykes C, Plastino J. Actin dynamics, architecture, and mechanics in cell motility. Physiol Rev. 2014;94:235–263.
    1. Bashour KT, et al. CD28 and CD3 have complementary roles in T-cell traction forces. Proc Natl Acad Sci U S A. 2014;111:2241–2246.
    1. Vaahtomeri K, et al. Locally triggered release of the chemokine CCL21 promotes dendritic cell transmigration across lymphatic endothelia. Cell Rep. 2017;19:902–909.
    1. Leithner A, et al. Dendritic cell actin dynamics control contact duration and priming efficiency at the immunological synapse. J Cell Biol. 2021;220:e202006081.
    1. Stack T, et al. Targeted delivery of cell softening micelles to Schlemm’s canal endothelial cells for treatment of glaucoma. Small. 2020;16:2004205
    1. Vorselen D, et al. Microparticle traction force microscopy reveals subcellular force exertion patterns in immune cell–target interactions. Nat Commun. 2020;11:20.
    1. Blumenthal D, Chandra V, Avery L, Burkhardt JK. Mouse T cell priming is enhanced by maturation-dependent stiffening of the dendritic cell cortex. Elife. 2020;9:e55995.
    1. Tello-Lafoz M, et al. Cytotoxic lymphocytes target characteristic biophysical vulnerabilities in cancer. Immunity. 2021;54:1037–1054.:e7.
    1. Janmey PA, McCulloch CA. Cell mechanics: Integrating cell responses to mechanical stimuli. Annu Rev Biomed Eng. 2007;9:1–34.
    1. Hanahan D, Weinberg RA. Hallmarks of cancer: The next generation. Cell. 2011;144:646–674.
    1. Chen DS, Mellman I. Oncology meets immunology: The cancer-immunity cycle. Immunity. 2013;39:110.
    1. Huse M. Mechanical forces in the immune system. Nat Rev Immunol. 2017;17:679–690.
    1. Zhu C, Chen W, Lou J, Rittase W, Li K. Mechanosensing through immunoreceptors. Nat Immunol. 2019;20:1269–1278.
    1. Jain N, Moeller J, Vogel V. Mechanobiology of macrophages: How physical factors coregulate macrophage plasticity and phagocytosis. Annu Rev Biomed Eng. 2019;21:267–297.
    1. Lei K, Kurum A, Tang L. Mechanical immunoengineering of T cells for therapeutic applications. Acc Chem Res. 2020;53:2777–2790.
    1. Moffat J, et al. A lentiviral RNAi library for human and mouse genes applied to an arrayed viral high-content screen. Cell. 2006;124:1283–1298.
    1. Rosen K, Shi W, Calabretta B, Filmus J. Cell detachment triggers p38 mitogen-activated protein kinase-dependent overexpression of Fas ligand: A novel mechanism of anoikis of intestinal epithelial cells. J Biol Chem. 2002;277:46123–46130.
    1. Hawley TS, Hawley RG. Flow Cytometry Protocols. Humana Press; New York: 2018.
    1. Tse JR, Engler AJ. Preparation of hydrogel substrates with tunable mechanical properties. Curr Protoc Cell Biol. 2010;47:1–16.
    1. Tseng Q, et al. Spatial organization of the extracellular matrix regulates cell-cell junction positioning. Proc Natl Acad Sci U S A. 2012;109:1506–1511.

Source: PubMed

3
Abonnere