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:

Cooperative phagocytosis of solid tumours by macrophages triggers durable anti-tumour responses

Abstract

In solid tumours, the abundance of macrophages is typically associated with a poor prognosis. However, macrophage clusters in tumour-cell nests have been associated with survival in some tumour types. Here, by using tumour organoids comprising macrophages and cancer cells opsonized via a monoclonal antibody, we show that highly ordered clusters of macrophages cooperatively phagocytose cancer cells to suppress tumour growth. In mice with poorly immunogenic tumours, the systemic delivery of macrophages with signal-regulatory protein alpha (SIRPα) genetically knocked out or else with blockade of the CD47–SIRPα macrophage checkpoint was combined with the monoclonal antibody and subsequently triggered the production of endogenous tumour-opsonizing immunoglobulin G, substantially increased the survival of the animals and helped confer durable protection from tumour re-challenge and metastasis. Maximizing phagocytic potency by increasing macrophage numbers, by tumour-cell opsonization and by disrupting the phagocytic checkpoint CD47–SIRPα may lead to durable anti-tumour responses in solid cancers.

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: Macrophages cooperatively phagocytose IgG-opsonized cancer cells in engineered tumouroids.
Fig. 2: Macrophage clusters in tumouroids are consistent with cooperative phagocytosis.
Fig. 3: Macrophages infiltrate, cluster and repress CD47-depleted syngeneic tumours in vivo only in combination with therapeutic IgG opsonization.
Fig. 4: Engineered macrophages phagocytose cancer cells efficiently and clear WT tumours in a dose-dependent fashion.
Fig. 5: Acquired anti-tumour immunity is required for long-term survival and protects against tumour re-challenge, experimental metastases and target antigen loss.
Fig. 6: Convalescent serum IgG drives B16-specific phagocytosis and clustering in tumouroids, and suppresses tumour initiation in vivo.
Fig. 7: Cooperative macrophages eliminate tumour cells to initiate an acquired immune response with phagocytic feedback.

Similar content being viewed by others

Data availability

The main data supporting the results in the study are available within the paper and its Supplementary Information. Source data are provided with this paper. Source data for the tumour-growth curves in Figs. 3f and 6c and in Extended Data Figs. 6a and 9a,b are provided with this paper. No large sequencing datasets were generated as part of this study, yet publicly available data from The Cancer Genome Atlas, The Human Protein Atlas and the National Center for Biotechnology Information GEO repository were accessed via the identifiers provided in Methods and in the relevant figure captions.

Code availability

Python code to generate random and structured simulated images is provided in Supplementary Information.

References

  1. Chen, J. et al. SLAMF7 is critical for phagocytosis of haematopoietic tumour cells via Mac-1 integrin. Nature 544, 493 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Morioka, S. et al. Efferocytosis induces a novel SLC program to promote glucose uptake and lactate release. Nature 563, 714–718 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Vorselen, D. et al. Microparticle traction force microscopy reveals subcellular force exertion patterns in immune cell–target interactions. Nat. Commun. 11, 20 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Nia, H. T., Munn, L. L. & Jain, R. K. Physical traits of cancer. Science 370, eaaz0868 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Manning, M. L., Foty, R. A., Steinberg, M. S. & Schoetz, E. M. Coaction of intercellular adhesion and cortical tension specifies tissue surface tension. Proc. Natl Acad. Sci. USA 107, 12517–12522 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Ohno, S. et al. The degree of macrophage infiltration into the cancer cell nest is a significant predictor of survival in gastric cancer patients. Anticancer Res. 23, 5015–5022 (2003).

    PubMed  Google Scholar 

  7. Ohno, S. et al. Correlation of histological localization of tumor-associated macrophages with clinicopathological features in endometrial cancer. Anticancer Res. 24, 3335–3342 (2004).

    PubMed  Google Scholar 

  8. Zindel, J. et al. Primordial GATA6 macrophages function as extravascular platelets in sterile injury. Science 371, 0–13 (2021).

    Article  CAS  Google Scholar 

  9. Flores-Toro, J. A. et al. CCR2 inhibition reduces tumor myeloid cells and unmasks a checkpoint inhibitor effect to slow progression of resistant murine gliomas. Proc. Natl Acad. Sci. USA 117, 1129–1138 (2020).

    Article  CAS  PubMed  Google Scholar 

  10. Georgouli, M. et al. Regional activation of myosin II in cancer cells drives tumor progression via a secretory cross-talk with the immune microenvironment. Cell 176, 757–774.e23 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Alvey, C. M. et al. SIRPA-inhibited, marrow-derived macrophages engorge, accumulate, and differentiate in antibody-targeted regression of solid tumors. Curr. Biol. 27, 2065–2077.e6 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Willingham, S. B. et al. The CD47-signal regulatory protein alpha (SIRPa) interaction is a therapeutic target for human solid tumors. Proc. Natl Acad. Sci. USA 109, 6662–6667 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Morrissey, M. A., Kern, N. & Vale, R. D. CD47 ligation repositions the inhibitory receptor SIRPA to suppress integrin activation and phagocytosis. Immunity 53, 290–302.e6 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Bakalar, M. H. et al. Size-dependent segregation controls macrophage phagocytosis of antibody-opsonized targets. Cell 174, 131–142.e13 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Advani, R. et al. CD47 blockade by Hu5F9-G4 and rituximab in non-Hodgkin’s lymphoma. N. Engl. J. Med. 379, 1711–1721 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Jalil, A. R., Andrechak, J. C. & Discher, D. E. Macrophage checkpoint blockade: results from initial clinical trials, binding analyses, and CD47-SIRPα structure–function. Antib. Ther. 3, 80–94 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Lakhani, N. J. et al. Evorpacept alone and in combination with pembrolizumab or trastuzumab in patients with advanced solid tumours (ASPEN-01): a first-in-human, open-label, multicentre, phase 1 dose-escalation and dose-expansion study. Lancet Oncol. 22, 1740–1751 (2021).

    Article  CAS  PubMed  Google Scholar 

  18. Kamber, R. A. et al. Inter-cellular CRISPR screens reveal regulators of cancer cell phagocytosis. Nature 597, 549–554 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Chao, M. P. et al. Calreticulin is the dominant pro-phagocytic signal on multiple human cancers and is counterbalanced by CD47. Sci. Transl. Med. 2, 63ra94 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Ingram, J. R. et al. Localized CD47 blockade enhances immunotherapy for murine melanoma. Proc. Natl Acad. Sci. USA 114, 10184–10189 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Tsai, R. K., Rodriguez, P. L. & Discher, D. E. Self inhibition of phagocytosis: the affinity of ‘marker of self’ CD47 for SIRPα dictates potency of inhibition but only at low expression levels. Blood Cells. Mol. Dis. 45, 67–74 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Suter, E. C. et al. Antibody:CD47 ratio regulates macrophage phagocytosis through competitive receptor phosphorylation. Cell Rep. 36, 109587 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Moynihan, K. D. et al. Eradication of large established tumors in mice by combination immunotherapy that engages innate and adaptive immune responses. Nat. Med. 22, 1402–1410 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Zhu, E. F. et al. Synergistic innate and adaptive immune response to combination immunotherapy with anti-tumor antigen antibodies and extended serum half-life IL-2. Cancer Cell 27, 489–501 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Sockolosky, J. T. et al. Durable antitumor responses to CD47 blockade require adaptive immune stimulation. Proc. Natl Acad. Sci. USA 113, E2646–E2654 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Wang, S., He, Z., Wang, X., Li, H. & Liu, X. S. Antigen presentation and tumor immunogenicity in cancer immunotherapy response prediction. eLife 8, e49020 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Liu, X. et al. CD47 blockade triggers T cell–mediated destruction of immunogenic tumors. Nat. Med. 21, 1209–1215 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Xu, M. M. et al. Dendritic cells but not macrophages sense tumor mitochondrial DNA for cross-priming through signal regulatory protein α signaling. Immunity 47, 363–373.e5 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Higgins, H. W., Lee, K. C., Galan, A. & Leffell, D. J. Melanoma in situ: part II. Histopathology, treatment, and clinical management. J. Am. Acad. Dermatol. 73, 193–203 (2015).

    Article  PubMed  Google Scholar 

  30. Majkut, S. et al. Heart-specific stiffening in early embryos parallels matrix and myosin expression to optimize beating. Curr. Biol. 23, 2434–2439 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Foty, R. A. & Steinberg, M. S. The differential adhesion hypothesis: a direct evaluation. Dev. Biol. 278, 255–263 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Guevorkian, K., Colbert, M. J., Durth, M., Dufour, S. & Brochard-Wyart, F. Aspiration of biological viscoelastic drops. Phys. Rev. Lett. 104, 218101 (2010).

    Article  PubMed  Google Scholar 

  33. Khalil, D. N. et al. An open-label, dose-escalation phase I study of anti-TYRP1 monoclonal antibody IMC-20D7S for patients with relapsed or refractory melanoma. Clin. Cancer Res. 22, 5204–5211 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Matlung, H. L. et al. Neutrophils kill antibody-opsonized cancer cells by trogoptosis. Cell Rep. 23, 3946–3959.e6 (2018).

    Article  CAS  PubMed  Google Scholar 

  35. Velmurugan, R., Challa, D. K., Ram, S., Ober, R. J. & Ward, E. S. Macrophage-mediated trogocytosis leads to death of antibody-opsonized tumor cells. Mol. Cancer Ther. 15, 1879–1889 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Martiniani, S., Chaikin, P. M. & Levine, D. Quantifying hidden order out of equilibrium. Phys. Rev. X 9, 11031 (2019).

    CAS  Google Scholar 

  37. Kakuta, S., Tagawa, Y. I., Shibata, S., Nanno, M. & Iwakura, Y. Inhibition of B16 melanoma experimental metastasis by interferon-γ through direct inhibition of cell proliferation and activation of antitumour host mechanisms. Immunology 105, 92–100 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Ammon, C. et al. Comparative analysis of integrin expression on monocyte-derived macrophages and monocyte-derived dendritic cells. Immunology 100, 364–369 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. McNally, A. K. & Anderson, J. M. β1 and β2 integrins mediate ddhesion during macrophage fusion and multinucleated foreign body giant cell formation. Am. J. Pathol. 160, 621–630 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Lehmann, B. et al. Tumor location determines tissue-specific recruitment of tumor-associated macrophages and antibody-dependent immunotherapy response. Sci. Immunol. 2, eaah6413 (2017).

    Article  PubMed  Google Scholar 

  41. Okazawa, H. et al. Negative regulation of phagocytosis in macrophages by the CD47-SHPS-1 system. J. Immunol. 174, 2004–2011 (2005).

    Article  CAS  PubMed  Google Scholar 

  42. Yanagita, T. et al. Anti-SIRPα antibodies as a potential new tool for cancer immunotherapy. JCI Insight 2, e89140 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Motegi, S.-I. et al. Essential roles of SHPS-1 in induction of contact hypersensitivity of skin. Immunol. Lett. 121, 52–60 (2008).

    Article  CAS  PubMed  Google Scholar 

  44. Tsai, R. K. & Discher, D. E. Inhibition of “self” engulfment through deactivation of myosin-II at the phagocytic synapse between human cells. J. Cell Biol. 180, 989–1003 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Wang, G. G. et al. Quantitative production of macrophages or neutrophils ex vivo using conditional Hoxb8. Nat. Methods 3, 287–293 (2006).

    Article  CAS  PubMed  Google Scholar 

  46. Roberts, A. W. et al. Cas9+ conditionally-immortalized macrophages as a tool for bacterial pathogenesis and beyond. eLife 8, e45957 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Trcka, J. et al. Redundant and alternative roles for activating Fc receptors and complement in an antibody-dependent model of autoimmune vitiligo. Immunity 16, 861–868 (2002).

    Article  CAS  PubMed  Google Scholar 

  48. Meeth, K., Wang, J. X., Micevic, G., Damsky, W. & Bosenberg, M. W. The YUMM lines: a series of congenic mouse melanoma cell lines with defined genetic alterations. Pigment Cell Melanoma Res. 29, 590–597 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Bruhns, P. Properties of mouse and human IgG receptors and their contribution to disease models. Blood 119, 5640–5649 (2012).

    Article  CAS  PubMed  Google Scholar 

  50. Carmi, Y. et al. Allogeneic IgG combined with dendritic cell stimuli induce antitumour T-cell immunity. Nature 521, 99–104 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Theruvath, J. et al. Anti-GD2 synergizes with CD47 blockade to mediate tumor eradication. Nat. Med 28, 333–344 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Lin, K.-Y. et al. Treatment of established tumors with a novel vaccine that enhances Major Histocompatibility Class II presentation of tumor antigen. Cancer Res. 56, 21–26 (1996).

    CAS  PubMed  Google Scholar 

  53. Oldenborg, P.-A., Gresham, H. D. & Lindberg, F. P. Cd47-signal regulatory protein α (Sirpα) regulates Fcγ and complement receptor–mediated phagocytosis. J. Exp. Med. 193, 855–862 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Ma, L. et al. Enhanced CAR–T cell activity against solid tumors by vaccine boosting through the chimeric receptor. Science 365, 162–168 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Oldenborg, P.-A., Gresham, H. D., Chen, Y., Izui, S. & Lindberg, F. P. Lethal autoimmune hemolytic anemia in CD47-deficient nonobese diabetic (NOD) mice. Blood 99, 3500–3504 (2002).

    Article  CAS  PubMed  Google Scholar 

  56. Mattes, M. J., Thomson, T. M., Old, L. J. & Lloyd, K. O. A pigmentation‐associated, differentiation antigen of human melanoma defined by a precipitating antibody in human serum. Int. J. Cancer 32, 717–721 (1983).

    Article  CAS  PubMed  Google Scholar 

  57. Fässler, M. et al. Antibodies as biomarker candidates for response and survival to checkpoint inhibitors in melanoma patients. J. Immunother. Cancer 7, 50 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  58. van Vloten, J. P. et al. Quantifying antigen-specific T cell responses when using antigen-agnostic immunotherapies. Mol. Ther. Methods Clin. Dev. 13, 154–166 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Rodríguez, D. et al. Tumor microenvironment profoundly modifies functional status of macrophages: peritoneal and tumor-associated macrophages are two very different subpopulations. Cell. Immunol. 283, 51–60 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Salmi, S. et al. The number and localization of CD68+ and CD163+ macrophages in different stages of cutaneous melanoma. Melanoma Res. 29, 237–247 (2019).

    Article  CAS  PubMed  Google Scholar 

  61. Fiumara, A. et al. In situ evidence of neoplastic cell phagocytosis by macrophages in papillary thyroid cancer. J. Clin. Endocrinol. Metab. 82, 1615–1620 (1997).

    CAS  PubMed  Google Scholar 

  62. Taskinen, M., Karjalainen-Lindsberg, M. L. & Leppä, S. Prognostic influence of tumor-infiltrating mast cells in patients with follicular lymphoma treated with rituximab and CHOP. Blood 111, 4664–4667 (2008).

    Article  CAS  PubMed  Google Scholar 

  63. Canioni, D. et al. High numbers of tumor-associated macrophages have an adverse prognostic value that can be circumvented by rituximab in patients with follicular lymphoma enrolled onto the GELA-GOELAMS FL-2000 trial. J. Clin. Oncol. 26, 440–446 (2008).

    Article  CAS  PubMed  Google Scholar 

  64. Morrissey, M. A. et al. Chimeric antigen receptors that trigger phagocytosis. eLife 7, e36688 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Sun, J. et al. Engineered proteins with sensing and activating modules for automated reprogramming of cellular functions. Nat. Commun. 8, 477 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Burkhardt, A. & Gebbers, J.O. Giant cell stromal reaction in squamous cell carcinomata. Virchows Arch. A Pathol. Anat. Histopathol 375, 263–280 (1977).

    Article  CAS  Google Scholar 

  67. Hayes, B. H. et al. Macrophages show higher levels of engulfment after disruption of cis interactions between CD47 and the checkpoint receptor SIRPα. J. Cell Sci. 133, jcs237800 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Edelstein, A., Amodaj, N., Hoover, K., Vale, R. & Stuurman, N. Computer control of microscopes using µManager. Mol. Biol. 92, 14.20.1–14.20.17 (2010).

    Google Scholar 

  69. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  PubMed  Google Scholar 

  70. Hoeksema, M. A. et al. IFN-γ priming of macrophages represses a part of the inflammatory program and attenuates neutrophil recruitment. J. Immunol. 194, 3909–3916 (2015).

    Article  CAS  PubMed  Google Scholar 

  71. Jablonski, K. A. et al. Novel markers to delineate murine M1 and M2 macrophages. PLoS ONE 10, 5–11 (2015).

    Article  Google Scholar 

  72. Dubrot, J. et al. In vivo screens using a selective CRISPR antigen removal lentiviral vector system reveal immune dependencies in renal cell carcinoma. Immunity 54, 571–585.e6 (2021).

    Article  CAS  PubMed  Google Scholar 

  73. Sato, M., Theret, D. P., Wheeler, L. T., Ohshima, N. & Nerem, R. M. Application of the micropipette technique to the measurement of cultured porcine aortic endothelial cell viscoelastic properties. J. Biomech. Eng. 112, 263–268 (1990).

    Article  CAS  PubMed  Google Scholar 

  74. Watanabe, T. et al. Higher cell stiffness indicating lower metastatic potential in B16 melanoma cell variants and in (−)-epigallocatechin gallate-treated cells. J. Cancer Res. Clin. Oncol. 138, 859–866 (2012).

    Article  CAS  PubMed  Google Scholar 

  75. Liu, J. et al. Soft fibrin gels promote selection and growth of tumorigenic cells. Nat. Mater. 11, 734–741 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Uhlén, M. et al. Tissue-based map of the human proteome. Science 347, 1260419 (2015).

    Article  PubMed  Google Scholar 

  77. Yesmin, F. et al. Ganglioside GD2 enhances the malignant phenotypes of melanoma cells by cooperating with integrins. Int. J. Mol. Sci. 23, 423 (2022).

    Article  CAS  Google Scholar 

  78. Mosely, S. I. S. et al. Rational selection of syngeneic preclinical tumor models for immunotherapeutic drug discovery. Cancer Immunol. Res. 5, 29–41 (2017).

    Article  CAS  PubMed  Google Scholar 

  79. Lechner, M. G. et al. Immunogenicity of murine solid tumor models as a defining feature of in vivo behavior and response to immunotherapy. J. Immunother. 36, 477–489 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by funding from the National Institutes of Health (NIH) and the National Science Foundation (NSF). D.E.D. discloses support for the research described in this study from NIH (U54 CA193417, U01 CA254886 and R01 HL124106), NSF (Materials Research Science and Engineering Center DMR-1720530 and DMR-1420530 and Grant Agreements CMMI 1548571 and 154857) and the Pennsylvania Department of Health (HRFF 4100083101). L.J.D. discloses support for the research described in this study from NIH (F32 CA228285). J.C.A and B.H.H. disclose support from the NSF (DGE-1845298). The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH or NSF. We thank C. Van Dang of the Wistar Institute for providing YUMM2.1 cells, S. Singhal for providing TC-1 cells and I. Brodsky of Penn Vet for providing editable CIM lines. We additionally thank J. Lee for aiding in development of the gene editing pipeline and S. Shin of Penn Vet for support on CIM progenitors. We acknowledge the following University of Pennsylvania core facilities: Cell Center, Stem Cell & Xenograft Core, Molecular Pathology & Imaging Core (supported by the Center for Molecular Studies in Digestive and Liver Diseases NIH/NIDDK Grant P30 DK050306), the Penn Vet Comparative Pathology Core, the Cell & Developmental Biology Microscopy Core, the Penn Cytomics & Cell Sorting Shared Research Laboratory (supported by the Abramson Cancer Center NIH/NCI Grant P30 CA016520) and the Penn Genomic Analysis Core.

Author information

Authors and Affiliations

Authors

Contributions

J.C.A., L.J.D. and D.E.D. designed the research. J.C.A., L.J.D., B.H.H., S.K., W.Z., R.P. and J.I. performed experiments. J.C.A., L.J.D., B.H.H., S.K., W.Z., R.P., M.V. and J.I. analysed data. J.C.A., L.J.D., B.H.H., J.I. and C.M.A. developed methodology. L.M. provided materials. D.E.D. supervised the work. J.C.A., L.J.D. and D.E.D. wrote the paper.

Corresponding author

Correspondence to Dennis E. Discher.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Biomedical Engineering thanks George Georgiou, In-San Kim and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data

Extended Data Fig. 1 Characterization of B16 tumouroids and subcutaneous tumours.

a, Brightfield and confocal microscopy of B16 tumouroids. Left: Brightfield image of a B16 tumouroid three days after 1000 cells were seeded atop an agarose gel. Scale bar: 0.5 mm. Right: Confocal max intensity projections (xy, xz planes) of GFP (green) and Hoechst 33342 (blue) in a B16 tumouroid fixed ~24 h after 100 cells were seeded in a non-adhesive well. Scale bar: 100 µm. b-d, Tumouroid cohesion depends on Ca2+ and the actin cytoskeleton. Transcriptomic analysis (b-i) of cadherin gene expression in cultured B16 cells (GSE162105) (ref. 72) and immunoblotting (b-ii) of B16 membrane, cytoplasmic, and nuclear protein fractions for N-cadherin, lamin-A/C, and tubulin. The uncropped blots are shown in Supplementary Fig. 5. Tumouroid (GFP+) area (c) following addition of 2 mM EDTA or EGTA (n = 16 tumouroids for PBS control or EGTA and n = 14 tumouroids for EDTA). In the box plots, the centreline is the median, the box is the interquartile range (IQR) from the 25th to 75th percentiles, and the whiskers extend over the entire range of the data. Statistical significance was assessed by one-way ANOVA at t = 1 h or 3 h, and adjusted p values were determined by Dunnett’s multiple comparisons test between PBS and EDTA or EGTA (p < 0.001 for all comparisons). Representative images depict GFP fluorescence in tumouroids at t = 3 h after EDTA (top) and PBS (bottom) treatment. Scale bar: 200 µm. Tumouroid area (d) following addition of 1 µM Latrunculin A (LatA) or DMSO vehicle (n = 16 tumouroids per condition, box plot as described in c). Statistical significance was assessed by Welch’s t-test (unpaired, two-tailed) between DMSO control and LatA treatment at t = 1 h or 3 h (p < 0.001 for both comparisons). Representative images depict GFP fluorescence in tumouroids at t = 3 h after treatment with LatA (top) and DMSO (bottom). Scale bar: 200 µm. e, Photographs of subcutaneous (s.c.) B16 tumours in mice dissected on day 4, 8, or 12 after tumour cell injection. Elongation of early-stage tumours was often observed in the direction in which the needle was inserted during s.c. injection (left). Tumour area calculated in ImageJ is reported beneath each photograph. Scale bars: 10 mm (left) and 5 mm (right). f, Trichrome-stained B16 tumour section. Insets depict the cellular (maroon cytoplasm and black nuclei) tumour interior (region 1) and fibrous ECM (blue) and large, round fat droplets (white) at the tumour periphery (region 2). Scale bar: 1 mm. Inset scale bar: 50 µm. g,h, Ex vivo pipette aspiration rheology of B16 tumours. Brightfield images (g-i) depict the extension (L) of tumour into the pipette after 90 s of aspiration at the indicated ΔP for an interior tumour region that exhibited large L and required low ΔP (left) and for a peripheral tumour region (fibrotic capsule and adjacent subcutaneous tissue) that exhibited small L and required higher ΔP for measurable deformation (right). Scale bar: 50 µm. Tissue strain L/RP, where RP is the radius of the pipette, is plotted (g-ii) for the constant pressure creep phase (0–90 s) and after the pressure was released back to ΔP = 0 for a representative interior tumour region. The black line depicts non-linear regression of L(t)/RP to the standard linear solid model, which includes the elastic modulus E as a fitted parameter. Elastic moduli (h) fit from the plateau in the strain vs. time plot for interior and peripheral regions of tumours excised days 4, 8, and 12 after engraftment. In the box plot, the centreline depicts the median, the box depicts the IQR, and the whiskers depict the 5th and 95th percentiles. The number of measurements n is combined from two or three tumours per time point. The solid lines are fits of simple exponential growth for the periphery (E ~ 0.8et/10) and exponential decay for the interior (E ~ 0.5e-t/10 + 0.3), where non-linear regression of the data was constrained so rates and y-intercepts would be shared for the periphery and interior. The shaded box denotes the range of E reported for B16 cells74,75 that is also typical for B16 tumouroids.

Source data

Extended Data Fig. 2 Analysis of CD47 and Tyrp1 expression and engulfment of B16 tumouroids and cell suspensions.

a, Flow cytometry analysis of CD47 and Tyrp1 expression on B16 cell lines. Representative histograms (a-i) show binding of 10 μg/mL anti-CD47 (left) and 20 μg/mL anti-Tyrp1 (right) to CD47 KO (top) and WT (bottom) B16 cell lines. Binding was detected with secondary antibodies conjugated with Alexa Fluor (AF) 647. The populations shown in the histograms were gated on live, single cells using forward/side scatter and DAPI staining. Median fluorescence intensity (MFI) values (a-ii) of CD47 KO and WT B16 incubated with varying concentrations of anti-CD47 or anti-Tyrp1 followed by secondary antibody (n = 3 tests per antibody concentration, mean ± s.d.). The solid line is a fit of MFI to a hyperbolic model of the form y = A*x/(K + x). The shaded regions correspond to blocking and opsonizing antibody concentrations that are required for tumouroid elimination in Fig. 1g. b, Tyrp1 protein expression on B16 melanoma cells. Immunofluorescence image (b-i) of a B16 cell surface-stained (fixed, unpermeabilized) with anti-Tyrp1 and anti-mouse IgG AF647 (green) and counterstained with Hoechst 33342 (blue). Brightfield image (b-ii) of a B16 cell. Inset: Pigmented melanosomes. Scale bars: 25 µm (left) and 10 µm (right). c, Protein expression data from The Human Protein Atlas76 for CD47 (https://www.proteinatlas.org/ENSG00000196776-CD47/tissue) and Tyrp1 (https://www.proteinatlas.org/ENSG00000107165-TYRP1/tissue) across different normal tissues. d, Conventional 2D phagocytosis (d-i) of CD47 KO and WT B16 from single-cell suspensions by primary mouse bone marrow-derived macrophages (BMDMs). Representative fluorescence image (d-ii) of phagocytosis of opsonized CD47 KO (green) by BMDMs (magenta). Yellow arrows denote phagocytic events. Scale bar: 25 µm. The phagocytic index (d-iii) is calculated as the percentage of macrophages engulfing a target cell multiplied by the number of engulfed cells per macrophage (mean ± s.e.m., n = 3 wells per condition). Statistical significance was assessed by two-way ANOVA and Tukey’s multiple comparisons test. e, Putative tumouroid trogocytic engulfment events consisting of GFP signal without co-localizing DNA signal (yellow arrows) and putative necrotic cell death observed as large, swollen B16 cells in tumouroids (orange arrow). Scale bars: 10 µm (left) and 100 µm (right). f, Co-localization of phagocytic macrophages in tumouroids with melanin pigments, which are visualized as a dark shadow in the GFP channel where the signal intensity is below background as illustrated in the colour heat map. Scale bar: 100 µm.

Extended Data Fig. 3 Cooperative tumouroid phagocytosis and non-cooperative phagocytosis in the conventional 2D assay.

a, CD47 KO and WT tumouroid growth curves with or without anti-Tyrp1 added on day 1 in the absence of macrophages (mean ± s.e.m., n = 6 tumouroids per condition). b, CD47 KO or WT tumouroid growth curves and effective growth rates at varying macrophage:B16 ratios without anti-Tyrp1 (mean ± s.e.m., n = 8 tumouroids in a representative experiment). c-e, Growth curves for WT tumouroids treated with 133 nM anti-CD47 and/or 133 nM anti-Tyrp1 at varying macrophage:WT ratios (c), 3:1 macrophage:WT tumouroids with 133 nM anti-Tyrp1 and varying concentrations of anti-CD47 (d), and 3:1 macrophage:CD47 KO tumouroids with varying concentrations of anti-Tyrp1 (e) (mean ± s.e.m. with exact sample numbers given in the plot legend). The tumouroids in panels (c-e) correspond to those used to determine keff as a function of macrophage:WT ratio, anti-CD47 concentration, or anti-Tyrp1 concentration plotted in Fig. 1f, g. f, Max intensity projection of confocal images of a WT tumouroid one day after addition of macrophages, anti-Tyrp1, and anti-CD47. Scale bar: 100 µm. The expansion depicts a single z slice with a cluster of macrophages that have engulfed B16 (white arrows). Scale bar: 25 µm. g, Phagocytic index for conventional 2D phagocytosis (g-i) with adherent macrophages and single-cell suspensions of CD47 KO or WT B16 opsonized with anti-Tyrp1 and treated with either anti-CD47 or rat IgG2a isotype control antibody (mean ± s.e.m., n = 3 wells per condition). Statistical significance was assessed by two-way ANOVA and Tukey’s multiple comparisons test. The phagocytic index (g-ii) calculated for each field of view (FOV) across all four conditions (2 cell lines x 2 blocking conditions) is plotted against the number of macrophages per FOV (Pearson r = −0.32, p = 0.01, two-tailed). h, Conventional 2D phagocytosis assay with varying macrophage surface densities and a fixed density of CD47 KO B16 opsonized with anti-Tyrp1. Representative FOVs with low and high macrophage density are shown (h-i). Scale bar: 50 µm. The phagocytic index (h-ii) calculated for each FOV is plotted against the number of macrophages per FOV as in g-ii (Pearson r = −0.36, p = 0.02, two-tailed). Alternatively, the data are fit with a segmental linear regression model (h-iii) relating the number of engulfed B16 to the number of macrophages per FOV (see Supplementary Fig. 2f for details). The dashed line approximates the expected number of B16 per FOV based on the density of the cell suspension, which is the theoretical maximum number of B16 that can be engulfed even if macrophage density increases.

Extended Data Fig. 4 Informational entropy calculation of macrophage order based on image compression.

a, Simulated 1024 × 1024 pixel (px) images containing 10,000 dark px on a white background that are arranged in random squares varying in size from 10,000 1 × 1 px squares to one 100 × 100 px square. The images were compressed in the lossless portable network graphics (PNG) file format. The PNG file size (in bytes) below each image is the average of 50 randomly generated images with the same specifications. b, To compare images with different numbers of dark pixels (N), a normalization procedure involved subtracting the PNG file size of an equal dimension white image and dividing this value by the PNG file size of an equal dimension random image with N dark pixels (minus a white image). c,d, Macrophages disperse spontaneously following phagocytosis of tumouroids. Representative images (c) of a 3:1 macrophage:CD47 KO tumouroid with anti-Tyrp1 on days 1–4. Scale bar: 100 µm. Macrophage entropy (d) during and after elimination of CD47 KO cells from tumouroids. The centreline depicts the median, the box depicts the IQR, and the whiskers depict the range of the data (n = 18 tumouroids). Statistical significance was assessed by one-way repeated measures ANOVA followed by Tukey’s multiple comparisons test.

Extended Data Fig. 5 Effects of cytokine priming or myosin-II inhibition on tumouroid growth suppression and macrophage clustering.

a, Tumouroid growth curves (a-i) following addition of macrophages primed for 48 h with 20 ng ml−1 IFNγ or 20 ng ml−1 IL-4 (mean ± s.e.m, n = 8 tumouroids). Statistical significance was assessed by one-way ANOVA and Dunnett’s multiple comparisons test between the mean area of tumouroids with unprimed macrophages and tumouroids with cytokine-primed macrophages at each time point. Growth curves for CD47 KO tumouroids (a-ii) in the presence of 20 µM p-amino-blebbistatin (NH2-blebb) to inhibit myosin-II or DMSO vehicle control (mean ± s.e.m., n = 4 tumouroids). b, Transcriptomic microarray analyses of BMDMs treated with IFNγ (GSE60290) (ref. 70) or IL-4 (GSE69607) (ref. 71) versus untreated BMDMs. The p values for differentially expressed genes were adjusted for multiple comparisons by the method of Benjamini and Hochberg to control the false discovery rate. The heat map depicts log2(fold change) for selected genes. Proteins encoded by differentially expressed genes are depicted in the schematic detailing a putative cytoskeleton and nuclear mechanosensing pathway and membrane receptors involved in phagocytosis and macrophage adhesion. c-e, Measurements of protein expression of selected differentially expressed genes in BMDMs treated with IFNγ or IL-4 for 48 h. Flow cytometry histograms (c) for surface staining of MHCII and CD206 in IFNγ- or IL-4-treated or untreated BMDMs. Similar staining was performed with antibodies against FcR’s, CD47, and SIRPα. The ratio of lamin-A to lamin-B and the projected nuclear area were quantified (d) by immunofluorescence microscopy (individual values for n > 140 cells per condition across 8 fields of view with a line denoting the median). Statistical significance was assessed by one-way ANOVA followed by Tukey’s multiple comparisons test. Scatterplot (e) of the fold change (cytokine-primed vs. untreated) of indicated proteins from flow cytometry (panel c) (mean ± s.d., replicates were BMDMs from n = 3 mice for MHCII staining and 4 mice for all other proteins) and lamin-A immunofluorescence staining (panel d) versus the fold change in gene expression from microarrays (panel b). f,g, Representative fluorescence images and entropy analysis for cytokine-primed (f) and blebbistatin-treated (g) macrophage monocultures on non-adhesive surfaces (mean ± s.e.m, n = 8 wells per condition). Cytokines and drugs were added after 6 h and images were acquired 20 h later. Scale bars: 100 µm. Statistical significance in panels f,g was assessed by one-way ANOVA followed by Tukey’s multiple comparisons test.

Extended Data Fig. 6 Isotype control, vehicle injections, and FcR-primed, SIRPα-blocked marrow cell injections; Biophysical crosslinking mechanism for SIRPα blockade by the anti-mouse SIRPα mAb clone P84.

a, Growth curves of tumours in mice injected with 2 × 105 CD47 KO cells and treated i.v. on days 4, 5, 7, 9, 11, 13, and 15 with 250 µg mouse IgG2a isotype control, with PBS vehicle, or left untreated (mean ± s.e.m., n = 6 mice per group). b, WT tumouroid growth curves and effective growth rates at varying macrophage:B16 ratios with anti-Tyrp1 and/or anti-SIRPα (mean ± s.e.m., n = 6–8 tumouroids in a representative experiment with exact n for anti-SIRPα + anti-Tyrp1 tumouroids reported in the plot legend). Statistical significance was assessed at each macrophage:B16 ratio by one-way ANOVA and Tukey’s multiple comparisons test. The p values reported in the plot correspond to differences in mean area between tumouroids treated with anti-SIRPα + anti-Tyrp1 and with anti-Tyrp1 only. c, Survival curves of mice bearing WT B16 tumours (parental line) treated with 2 × 107 A’PB (n = 22 mice), A’ (FcR-primed only, no SIRPα block; n = 14), or B (no FcR priming, only SIRPα block; n = 5) marrow cells i.v. on day 4 after tumour engraftment or left untreated (n = 21 mice). A’PB and A’ conditions included additional 250 µg anti-Tyrp1 i.v. on days 5, 7, 9, 11, 13, and 15. Statistical significance was determined by the log-rank (Mantel-Cox) test. d-g, Anti-mouse SIRPα antibody clone P84 binds to mouse macrophages and immobilizes SIRPα. Representative fluorescence images (d) and quantification (e) of fluorescence recovery after photobleaching (FRAP) for SIRPα-GFP ± anti-mSIRPα clone P84. Bleached regions (yellow arrow) are shown prior to bleaching and at early and late time points of recovery (n = 3). Scale bar: 10 µm. Experimental schematic and representative images (f) of J774A.1 mouse macrophages ± anti-mSIRPα clone P84 incubated with IgG-opsonized mouse RBCs for 30 min prior to immunostaining for SIRPα and phospho-tyrosine (p-Tyr). Quantification (g) of SIRPα and p-Tyr immunofluorescence at the phagocytic synapse between the macrophage and target RBC. Scale bar: 10 µm. Fluorescence intensity in the phagocytic synapse was normalized to fluorescence from macrophages that were not undergoing phagocytosis (mean ± s.e.m., n = 7). Statistical significance of each imaging bin along the phagocytic synapse was assessed by Welch’s t-test (two-tailed, unpaired with Holm-Šídák correction for multiple comparisons). h, Summary of macrophage number versus therapeutic outcome for tumouroids in vitro and tumours in vivo. Growth rate is controlled by tumour cell proliferation and macrophage phagocytosis. CD47 signalling dominates pro-phagocytic signalling at low and high macrophage numbers, but CD47 depletion, IgG opsonization, and macrophage infiltration can eliminate tumours.

Source data

Extended Data Fig. 7 Conditionally immortalized macrophage progenitor engineering and opsonization of various B16 lines with convalescent serum.

a,b, Flow cytometry analysis of surface markers on conditionally immortalized macrophage (CIM) progenitors. Staining for myeloid markers (a) on CIM progenitors and freshly harvested bone marrow cells for comparison. Flow histograms of SIRPα staining (b) on SIRPα KO CIM progenitors and unedited wild-type CIM progenitors. Anti-SIRPα binding was detected with AF546 anti-rat IgG. c–f, Convalescent serum IgG from CD47 KO tumour complete responders opsonizes CD47 KO, DKO, and WT B16. (c) Representative fluorescence image of serum-opsonized CD47 KO B16 engulfed by BMDMs (yellow arrows). Scale bar: 50 μm. (d) Quantification of the phagocytic index for macrophages engulfing CD47 KO B16 opsonized with convalescent sera, naïve serum, or mAb controls. Convalescent sera were collected from second challenge survivors 70 days after tumour challenge (mean ± s.e.m., n = 3 wells for antibody conditions, n = 3 wells per serum sample for one naïve serum and three convalescent sera). (e) Percent phagocytic macrophages engulfing DKO B16 (lacking CD47 and Tyrp1) opsonized with convalescent sera collected from first challenge survivors 99 days after tumour challenge, naïve sera, or mAb controls (mean ± s.e.m., n = 3 wells for antibody conditions, n = 3 wells per serum sample for four naïve or convalescent sera). (f) Percent phagocytic macrophages engulfing WT B16 opsonized with convalescent serum from a third challenge survivor, naïve serum, or anti-Tyrp1 (mean ± s.e.m., n = 3 wells per condition). In panels (d-f), statistical significance was assessed by one-way ANOVA and Šídák’s multiple comparisons test between selected groups.

Extended Data Fig. 8 Convalescent sera from CD47 KO tumour complete responders contain multiple subclasses of IgG that bind CD47 KO and DKO B16, but not syngeneic YUMM2.1 melanoma cells.

a, Kinetics of de novo IgG generation revealed by flow cytometry of CD47 KO (filled symbols) and DKO B16 cells (open symbols) stained with serum followed by a polyclonal anti-mouse IgG [H+L] secondary antibody (row 1) or secondary antibodies specific for mouse IgG1 and IgG3 (rows 2 and 3). The corresponding data for IgG2a/c and IgG2b are shown in Fig. 6b. The median fluorescence intensity was corrected by subtracting the background fluorescence equal to the signal of cells incubated with secondary antibodies only. Symbols and error bars depict the mean ± s.e.m. of each group. The range of staining observed with 10 µg ml−1 anti-Tyrp1 is denoted by the pink boxes or lines. The gating strategy and representative histograms are shown in Supplementary Fig. 4. b, Median fluorescence intensity of B16 WT or Yumm2.1 cells incubated with 5% (v/v) serum collected from a mouse that survived the third challenge, with naïve serum, or with anti-Tyrp1 followed by staining with monoclonal secondary antibodies as in Fig. 6b and panel (a) of this figure. c, Growth curves for CD47 KO tumouroids treated with convalescent serum, naïve serum, anti-Tyrp1, or mIgG2a isotype control Ab without macrophages. Solid lines are non-linear regression of the data to a simple exponential of the form A(t) = A1ek(t-1) (mean ± s.e.m., n = 3 or 4 tumouroids each for anti-Tyrp1, mIgG2a, 9 convalescent sera, and 9 naïve sera).

Extended Data Fig. 9 Generalization of CD47 disruption and macrophage phagocytosis in prostate, lung, and glioblastoma tumour models.

a, Targeting RM-9 prostate tumours for phagocytosis. Immunostaining of RM-9 (a-i) with anti-GD2 and mIgG2a isotype control Ab. Bound primary antibodies were detected with a secondary antibody conjugated with AF647. Clusters of anti-GD2 are consistent with ‘cap formation’ as should occur due to GD2 association with rafts77 or due to the polyclonal secondary antibody. Scale bar: 100 µm. Flow cytometry histograms (a-ii) of anti-GD2 (left) and anti-CD47 (right) and their isotype (iso.) controls binding to RM-9. Bound primary antibodies were detected with secondary antibodies conjugated with PE and AF647, respectively. Phagocytosis of RM-9 cells (a-iii) by BMDMs with anti-GD2 opsonization and/or CD47 antibody blockade. Statistical significance was assessed by one-way ANOVA with Tukey’s multiple comparisons test (mean ± s.e.m., n = 3 wells per condition). WT RM-9 tumour growth curves (a-iv). Mice were treated with 4 × 106 SIRPα KO CIM progenitors i.v. on day 4 and 250 µg anti-GD2 i.v. on days 4, 5, 7, 9, 11, 13, and 15 (n = 10) or left untreated (n = 5). Each symbol represents a separate tumour, and complete responses (‘survivor’, 1) in which a tumour was never palpable are depicted with the half-filled squares. b, Targeting TC-1 lung tumours for phagocytosis. (b-i) Phagocytosis of CD47 KO TC-1 by BMDMs with anti-mouse RBC IgG opsonization (mean ± s.e.m., n = 3 per condition). Statistical significance was assessed by Welch’s t-test (two-tailed, unpaired). (b-ii) Tumour growth curves at early timepoints following injection of TC-1 cells pre-opsonized with anti-mouse RBC and anti-CD47 or unopsonized controls (mean ± s.e.m., n = 10 mice per group). Statistical significance at each timepoint between control (squares) and pre-opsonized (circles) was assessed by Student’s t-test (two-tailed, unpaired). The dashed lines are linear fits, which were compared by the extra sum-of-squares F test (p = 0.002, two-tailed). c, Binding and functional analysis of convalescent serum IgG from vaccine-enhanced CAR-T therapy against CT-2A glioma54. Median fluorescence intensity (c-i) (corrected by subtracting the background signal measured on unstained cells) of CT-2A stained with 5% (v/v) convalescent or naïve serum followed by the same secondary Ab panel used for Fig. 6b and Extended Data Fig. 8a, b (median with all data points, n = 5 convalescent and 3 naïve sera). (c-ii) Lack of phagocytosis of CT-2A cells treated with convalescent sera even when CD47 was blocked. As a positive control, CT-2A were opsonized with 5% (v/v) anti-mRBC polyclonal antiserum (mean ± s.d., n = 3 wells per condition).

Source data

Extended Data Fig. 10 Potential for combination immunotherapy in human melanoma and rationale for CD47 combination therapy versus monotherapy in syngeneic mouse models.

a, TCGA analysis of TYRP1 and CD47 in human melanoma. Survival analysis of metastatic melanoma patients in The Cancer Genome Atlas (TCGA SKCM) based on expression of TYRP1 (left) and CD47 (right) above or below the median. Significance was determined by the log-rank (Mantel-Cox) test. Expression levels of TYRP1 (left, one-way ANOVA, p = 0.68) and CD47 (right, p = 0.36) across different treatment modalities reported in TCGA. In the box plots, the centreline depicts the median, the boxes span the IQR, and the whiskers span 1.5 × IQR with outlier points included. b, Meta-analysis of tumour growth from data reported by Mosely et al.78 and Lechner et al.79 across different syngeneic mouse tumour models plotted against an immunogenicity score defined as the summation of log2(expression) of MHC class I-related genes reported by Mosely et al. (Pearson correlation coefficients r and two-tailed p values are reported in the plot legend). Graphical growth data were digitized using WebPlotDigitizer version 4.2, and growth curves were fit to the exponential growth equation V(t) = V0ekt to determine the tumour growth rate k (± standard error). Tumour types are: melanoma (B16), lung (LLC), kidney (RENCA), breast (4T1), and colon (CT26 and MC38). Note that B16 doubling rates in our studies are consistent with the data here, even with CD47 knockout.

Supplementary information

Supplementary Information

Supplementary Figures, Tables and Code.

Reporting Summary

Peer Review File

Source data

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dooling, L.J., Andrechak, J.C., Hayes, B.H. et al. Cooperative phagocytosis of solid tumours by macrophages triggers durable anti-tumour responses. Nat. Biomed. Eng 7, 1081–1096 (2023). https://doi.org/10.1038/s41551-023-01031-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41551-023-01031-3

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