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:

The macrophage tetraspan MS4A4A enhances dectin-1-dependent NK cell–mediated resistance to metastasis

Abstract

The plasma membrane tetraspan molecule MS4A4A is selectively expressed by macrophage-lineage cells, but its function is unknown. Here we report that MS4A4A was restricted to murine and human mononuclear phagocytes and was induced during monocyte-to-macrophage differentiation in the presence of interleukin 4 or dexamethasone. Human MS4A4A was co-expressed with M2/M2-like molecules in subsets of normal tissue-resident macrophages, infiltrating macrophages from inflamed synovium and tumor-associated macrophages. MS4A4A interacted and colocalized with the β-glucan receptor dectin-1 in lipid rafts. In response to dectin-1 ligands, Ms4a4a-deficient macrophages showed defective signaling and defective production of effector molecules. In experimental models of tumor progression and metastasis, Ms4a4a deficiency in macrophages had no impact on primary tumor growth, but was essential for dectin-1-mediated activation of macrophages and natural killer (NK) cell–mediated metastasis control. Thus, MS4A4A is a tetraspan molecule selectively expressed in macrophages during differentiation and polarization, essential for dectin-1-dependent activation of NK cell–mediated resistance to metastasis.

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: MS4A4A expression and its regulation in human macrophages.
Fig. 2: Macrophage expression of MS4A4A in human tissues.
Fig. 3: MS4A4A expression is restricted to CD163+ macrophages in inflamed synovium.
Fig. 4: MS4A4A is expressed in murine TAMs but is dispensable for primary tumor growth.
Fig. 5: MS4A4A and dectin-1 associate in lipid rafts after zymosan engagement.
Fig. 6: Ms4a4a is required for dectin-1 signaling.
Fig. 7: Ms4a4a regulates dectin-1 recognition and control of B16F1 metastasis.
Fig. 8: Expression of Ms4a4a in macrophages is required for NK-mediated, dectin-1-triggered protection against metastasis.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request. Ms4a4afl/fl animals have been developed by the research team and are available under a material transfer agreement after evaluation of potential conflict of interest.

References

  1. Biswas, S. K. & Mantovani, A. Macrophage plasticity and interaction with lymphocyte subsets: cancer as a paradigm. Nat. Immunol. 11, 889–896 (2010).

    Article  CAS  Google Scholar 

  2. Murray, P. J. et al. Macrophage activation and polarization: nomenclature and experimental guidelines. Immunity 41, 14–20 (2014).

    Article  CAS  Google Scholar 

  3. Mantovani, A., Marchesi, F., Malesci, A., Laghi, L. & Allavena, P. Tumour-associated macrophages as treatment targets in oncology. Nat. Rev. Clin. Oncol. 14, 399–416 (2017).

    Article  CAS  Google Scholar 

  4. Ruffell, B., Affara, N. I. & Coussens, L. M. Differential macrophage programming in the tumor microenvironment. Trends Immunol. 33, 119–126 (2012).

    Article  CAS  Google Scholar 

  5. Gabrilovich, D. I., Ostrand-Rosenberg, S. & Bronte, V. Coordinated regulation of myeloid cells by tumours. Nat. Rev. Immunol. 12, 253–268 (2012).

    Article  CAS  Google Scholar 

  6. Martinez, F. O., Gordon, S., Locati, M. & Mantovani, A. Transcriptional profiling of the human monocyte-to-macrophage differentiation and polarization: new molecules and patterns of gene expression. J. Immunol. 177, 7303–7311 (2006).

    Article  CAS  Google Scholar 

  7. Martinez, F. O. et al. Genetic programs expressed in resting and IL-4 alternatively activated mouse and human macrophages: similarities and differences. Blood 121, e57–e69 (2013).

    Article  CAS  Google Scholar 

  8. Varol, C., Mildner, A. & Jung, S. Macrophages: development and tissue specialization. Annu. Rev. Immunol. 33, 643–675 (2015).

    Article  CAS  Google Scholar 

  9. Sanyal, R. et al. MS4A4A: a novel cell surface marker for M2 macrophages and plasma cells. Immunol. Cell Biol. 95, 611–619 (2017).

    Article  CAS  Google Scholar 

  10. Charrin, S. et al. Lateral organization of membrane proteins: tetraspanins spin their web. Biochem. J. 420, 133–154 (2009).

    Article  CAS  Google Scholar 

  11. Levy, S. & Shoham, T. The tetraspanin web modulates immune-signalling complexes. Nat. Rev. Immunol. 5, 136–148 (2005).

    Article  CAS  Google Scholar 

  12. Howie, D. et al. MS4A4B is a GITR-associated membrane adapter, expressed by regulatory T cells, which modulates T cell activation. J. Immunol. 183, 4197–4204 (2009).

    Article  CAS  Google Scholar 

  13. Chiba, S. et al. Recognition of tumor cells by Dectin-1 orchestrates innate immune cells for anti-tumor responses. eLife 3, e04177 (2014).

    Article  Google Scholar 

  14. Adra, C. N. et al. Cloning of the cDNA for a hematopoietic cell-specific protein related to CD20 and the beta subunit of the high-affinity IgE receptor: evidence for a family of proteins with four membrane-spanning regions. Proc. Natl Acad. Sci. USA 91, 10178–10182 (1994).

    Article  CAS  Google Scholar 

  15. Hupp, K., Siwarski, D., Mock, B. A. & Kinet, J. P. Gene mapping of the three subunits of the high affinity FcR for IgE to mouse chromosomes 1 and 19. J. Immunol. 143, 3787–3791 (1989).

    CAS  PubMed  Google Scholar 

  16. Hollingworth, P. et al. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s disease. Nat. Genet. 43, 429–435 (2011).

    Article  CAS  Google Scholar 

  17. Greer, P. L. et al. A family of non-GPCR chemosensors defines an alternative logic for mammalian olfaction. Cell 165, 1734–1748 (2016).

    Article  CAS  Google Scholar 

  18. Cruse, G. et al. The CD20 homologue MS4A4 directs trafficking of KIT toward clathrin-independent endocytosis pathways and thus regulates receptor signaling and recycling. Mol. Biol. Cell 26, 1711–1727 (2015).

    Article  CAS  Google Scholar 

  19. Ishibashi, K., Suzuki, M., Sasaki, S. & Imai, M. Identification of a new multigene four-transmembrane family (MS4A) related to CD20, HTm4 and beta subunit of the high-affinity IgE receptor. Gene 264, 87–93 (2001).

    Article  CAS  Google Scholar 

  20. Boucheix, C. & Rubinstein, E. Tetraspanins. Cell Mol. Life Sci. 58, 1189–1205 (2001).

    Article  CAS  Google Scholar 

  21. Hemler, M. E. Tetraspanin functions and associated microdomains. Nat. Rev. Mol. Cell Biol. 6, 801–811 (2005).

    Article  CAS  Google Scholar 

  22. Tarrant, J. M., Robb, L., van Spriel, A. B. & Wright, M. D. Tetraspanins: molecular organisers of the leukocyte surface. Trends Immunol. 24, 610–617 (2003).

    Article  CAS  Google Scholar 

  23. Dombrowicz, D. et al. Allergy-associated FcRbeta is a molecular amplifier of IgE- and IgG-mediated in vivo responses. Immunity 8, 517–529 (1998).

    Article  CAS  Google Scholar 

  24. Polyak, M. J., Li, H., Shariat, N. & Deans, J. P. CD20 homo-oligomers physically associate with the B cell antigen receptor. Dissociation upon receptor engagement and recruitment of phosphoproteins and calmodulin-binding proteins. J. Biol. Chem. 283, 18545–18552 (2008).

    Article  CAS  Google Scholar 

  25. Meyer-Wentrup, F. et al. Dectin-1 interaction with tetraspanin CD37 inhibits IL-6 production. J. Immunol. 178, 154–162 (2007).

    Article  CAS  Google Scholar 

  26. Taylor, P. R. et al. Dectin-1 is required for beta-glucan recognition and control of fungal infection. Nat. Immunol. 8, 31–38 (2007).

    Article  CAS  Google Scholar 

  27. Wuthrich, M., Deepe, G. S. Jr. & Klein, B. Adaptive immunity to fungi. Annu. Rev. Immunol. 30, 115–148 (2012).

    Article  CAS  Google Scholar 

  28. Dambuza, I. M. & Brown, G. D. C-type lectins in immunity: recent developments. Curr. Opin. Immunol. 32, 21–27 (2015).

    Article  CAS  Google Scholar 

  29. Goodridge, H. S. et al. Differential use of CARD9 by dectin-1 in macrophages and dendritic cells. J. Immunol. 182, 1146–1154 (2009).

    Article  CAS  Google Scholar 

  30. Brown, G. D. Dectin-1: a signalling non-TLR pattern-recognition receptor. Nat. Rev. Immunol. 6, 33–43 (2006).

    Article  CAS  Google Scholar 

  31. del Fresno, C. et al. Interferon-beta production via Dectin-1-Syk-IRF5 signaling in dendritic cells is crucial for immunity to C. albicans. Immunity 38, 1176–1186 (2013).

    Article  Google Scholar 

  32. Brubaker, S. W., Bonham, K. S., Zanoni, I. & Kagan, J. C. Innate immune pattern recognition: a cell biological perspective. Annu. Rev. Immunol. 33, 257–290 (2015).

    Article  CAS  Google Scholar 

  33. Rogers, N. C. et al. Syk-dependent cytokine induction by Dectin-1 reveals a novel pattern recognition pathway for C type lectins. Immunity 22, 507–517 (2005).

    Article  CAS  Google Scholar 

  34. Zanoni, I. et al. CD14 controls the LPS-induced endocytosis of Toll-like receptor 4. Cell 147, 868–880 (2011).

    Article  CAS  Google Scholar 

  35. Dennehy, K. M. et al. Syk kinase is required for collaborative cytokine production induced through Dectin-1 and Toll-like receptors. Eur. J. Immunol. 38, 500–506 (2008).

    Article  CAS  Google Scholar 

  36. Herre, J. et al. Dectin-1 uses novel mechanisms for yeast phagocytosis in macrophages. Blood 104, 4038–4045 (2004).

    Article  CAS  Google Scholar 

  37. Underhill, D. M., Rossnagle, E., Lowell, C. A. & Simmons, R. M. Dectin-1 activates Syk tyrosine kinase in a dynamic subset of macrophages for reactive oxygen production. Blood 106, 2543–2550 (2005).

    Article  CAS  Google Scholar 

  38. Cain, D. W. & Cidlowski, J. A. Immune regulation by glucocorticoids. Nat. Rev. Immunol. 17, 233–247 (2017).

    Article  CAS  Google Scholar 

  39. Colotta, F. et al. Interleukin-1 type II receptor: a decoy target for IL-1 that is regulated by IL-4. Science 261, 472–475 (1993).

    Article  CAS  Google Scholar 

  40. Kodelja, V. et al. Alternative macrophage activation-associated CC-chemokine-1, a novel structural homologue of macrophage inflammatory protein-1 alpha with a Th2-associated expression pattern. J. Immunol. 160, 1411–1418 (1998).

    CAS  PubMed  Google Scholar 

  41. Bellora, F. et al. The interaction of human natural killer cells with either unpolarized or polarized macrophages results in different functional outcomes. Proc. Natl Acad. Sci. USA 107, 21659–21664 (2010).

    Article  CAS  Google Scholar 

  42. Mattiola, I. et al. Priming of human resting NK cells by autologous M1 macrophages via the engagement of IL-1β, IFN-β, and IL-15 pathways. J. Immunol. 195, 2818–2828 (2015).

    Article  CAS  Google Scholar 

  43. Gorelik, E., Wiltrout, R. H., Okumura, K., Habu, S. & Herberman, R. B. Role of NK cells in the control of metastatic spread and growth of tumor cells in mice. Int. J. Cancer 30, 107–112 (1982).

    Article  CAS  Google Scholar 

  44. Molgora, M., Barajon, I., Mantovani, A. & Garlanda, C. Regulatory role of IL-1R8 in immunity and disease. Front. Immunol. 7, 149 (2016).

    Article  Google Scholar 

  45. Molgora, M. et al. IL-1R8 is a checkpoint in NK cells regulating anti-tumour and anti-viral activity. Nature 551, 110–114 (2017).

    Article  Google Scholar 

  46. Mantovani, A., Giavazzi, R., Polentarutti, N., Spreafico, F. & Garattini, S. Divergent effects of macrophage toxins on growth of primary tumors and lung metastases in mice. Int. J. Cancer 25, 617–620 (1980).

    Article  CAS  Google Scholar 

  47. Massara, M. et al. ACKR2 in hematopoietic precursors as a checkpoint of neutrophil release and anti-metastatic activity. Nat. Commun. 9, 676 (2018).

    Article  Google Scholar 

  48. Noy, R. & Pollard, J. W. Tumor-associated macrophages: from mechanisms to therapy. Immunity 41, 49–61 (2014).

    Article  CAS  Google Scholar 

  49. Majety, M., Runza, V., Lehmann, C., Hoves, S. & Ries, C. H. A drug development perspective on targeting tumor-associated myeloid cells. FEBS J. 285, 763–776 (2018).

    Article  CAS  Google Scholar 

  50. Franks, S. E., Getahun, A., Hogarth, P. M. & Cambier, J. C. Targeting B cells in treatment of autoimmunity. Curr. Opin. Immunol. 43, 39–45 (2016).

    Article  CAS  Google Scholar 

  51. Aletaha, D. et al. 2010 Rheumatoid arthritis classification criteria: an American College of Rheumatology/European League Against Rheumatism collaborative initiative. Arthritis Rheum. 62, 2569–2581 (2010).

    Article  Google Scholar 

  52. Kelly, S. et al. Ultrasound-guided synovial biopsy: a safe, well-tolerated and reliable technique for obtaining high-quality synovial tissue from both large and small joints in early arthritis patients. Ann. Rheum. Dis. 74, 611–617 (2015).

    Article  CAS  Google Scholar 

  53. Bonavita, E. et al. PTX3 is an extrinsic oncosuppressor regulating complement-dependent inflammation in cancer. Cell 160, 700–714 (2015).

    Article  CAS  Google Scholar 

  54. Millington, M. et al. High-precision FLIM–FRET in fixed and living cells reveals heterogeneity in a simple CFP–YFP fusion protein. Biophys. Chem. 127, 155–164 (2007).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The authors thank C. Garlanda and M. Kallikourdis (Humanitas Clinical and Research Center, Scientific Institute for Research and Healthcare (IRCCS)) for providing cDNAs of murine leukocyte subsets, and T. Irimura (Juntendo University School of Medicine, Tokyo, Japan) and R. Giavazzi (Mario Negri Institute, Milan, Italy) for providing MC38 and SL4 cells. Technical assistance from A. Fontanini, C. Perrucchini, T. Schorn, R. Porte and F. Pasqualini is acknowledged. A. Inforzato (Humanitas Clinical and Research Center, IRCCS), A. Diefenbach (Charité – Universitätsmedizin Berlin, Germany) and L. Florin (University Medical Centre of the Johannes Gutenberg University, Mainz, Germany) are gratefully acknowledged for their support and discussion. Financial support came from Fondazione Cariplo (grant no. 2015–0564 to A.M.), Cluster Alisei (grant no. MEDINTECH CTN01_00177_962865 to A.M.), the European Research Council (grant no. 669415-PHII to A.M.), the Italian Association for Cancer Research (AIRC IG-2016 grant no. 19014 to A.M.; AIRC 5 × 1000 grant no. 21147 to A.M.; AIRC IG-2016 grant no. 19213 to M.L.), Medical Research Council (Pathobiology of Early Arthritis Cohort grant no. 36661 to C.P.) and Arthritis Research UK Experimental Treatment Centre (grant no. 20022 to C.P.). I.M. was supported by a Mario and Valeria Rindi fellowship and a Fellowship for abroad from the Italian Foundation for Cancer Research, and by a European Federation of Immunological Societies-IL short-term fellowship. B.S. was supported by Ministero della Salute (progetto Finalizzata GR-2013-02356522). R.S.G. was supported by a PhD studentship (PD/BD/114138/2016) from Fundação para a Ciência e Tecnologia, Portugal. S.K.B. was supported by core funding from Singapore Immunology Network—Agency for Science, Technology & Research (A*STAR), Singapore. F.T. was supported by a fellowship from the A*STAR Research Attachment Program, Singapore. S.L. was supported by Fondazione Beretta, Italy.

Author information

Authors and Affiliations

Authors

Contributions

I.M., F.T. and M.D.P. performed in vitro experiments. I.M., B.S. and F.T. performed in vivo experiments. R.S.-G., M.M. and M. Sironi provided support for in vivo and in vitro experiments. A.D. conducted the imaging analysis. D.M. and M. Stravalaci provided support for the analysis of molecular interactions. R.C. performed bioinformatics analysis. I.N.S. and Y.L. performed gene expression experiments. S.L., W.V., M.A.B., A.N. and T.G. performed histology. G.M. and C.P. provided access to patients’ samples. S.K.B., B.B., C.P., A.M. and M.L. contributed to the experimental design and the supervision of the study.

Corresponding authors

Correspondence to Alberto Mantovani or Massimo Locati.

Ethics declarations

Competing interests

A.M. is a recipient of commercial research grants from Novartis; a consultant/advisory board member for Novartis, Roche, Ventana, Pierre Fabre, Verily, AbbVie, Compugen, Macrophage Therapeutics, AstraZeneca, Biovelocita, BG Fund, Third Rock and Verseau; an inventor of patents related to PTX3 and other innate immunity molecules; and also receives royalties for reagents related to innate immunity.

Additional information

Peer Review Information: Laurie Dempsey was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Integrated supplementary information

Supplementary Figure 1 Expression levels of MS4A4A, MS4A6A and MS4A7 in human monocyte-derived macrophages and DCs.

Representative dot plots and Mean Fluorescence Intensity (MFI) of the expression of MS4A4A (a, b), MS4A6A (c, d), and MS4A7 (e, f) in monocyte-derived macrophages (Mɸ) and dendritic cells (DC) treated (Dex, open symbol) or not (NT, closed symbol) with 10−6 M Dex for 24 h. Ctrl represents isotype control staining. Results are shown as mean ± SEM. Three independent experiments were performed for panels a to d (n = 5 donors) and four independent experiments were performed for panels e to h (n = 6 donors). Statistical analysis by one-way ANOVA.

Supplementary Figure 2 Transcript levels of MS4A4A in human tissues.

MS4A4A transcript levels in human tissues (a) and cell types (b) based on the Laboratory for Systems Biology and Medicine database RefExA. EnC, endothelial cells, EpC, epithelial cells, SMC, smooth muscle cells.

Supplementary Figure 3 MS4A4A transcript levels in human tumors and its correlation with TAM markers.

a) Boxplot of MS4A4A mRNA expression in human tumors (closed boxes) and corresponding normal tissues (open boxes) among different TCGA cancer datasets. b) Heatmap correlation between mRNAs encoding MS4A4A and TAM markers in tumor tissues among different TCGA cancer datasets. Genes are represented according to their average Pearson correlation with MS4A4A among all cancer datasets. Number of samples included in the analysis of both panels (normal tissue, corresponding tumor tissue) are: bladder urothelial cancer: 19, 408; breast cancer: 112, 1100; cervical adenocarcinoma: 3, 306; cholangiocarcinoma: 9, 36; colon adenocarcinoma: 41, 287; colorectal adenocarcinoma: 51, 382; esophageal cancer: 11, 185; esophageal carcinoma: 46, 600; glioblastoma: 5, 166; glioma: 5, 696; head and neck cancer: 44, 522; kidney cancer: 25, 66; kidney clear cell cancer: 72, 534; kidney renal papillary cell cancer: 32, 291; liver cancer: 50, 373; lung cancer: 59, 517; lung squamous cell cancer: 51, 501; pan-kidney cohort: 129, 891; pancreatic cancer: 4, 179; paraganglioma: 3, 184; prostate cancer: 52, 498; rectum cancer: 10, 95; sarcoma: 2, 263; stomach adenocarcinoma: 35, 415; thymoma: 2, 120; thyroid cancer: 59, 509; uterine cancer: 24, 177.

Supplementary Figure 4 Ms4a4a expression in murine leukocyte subsets and polarized BMDMs.

a) Ms4a4a mRNA levels in murine leukocyte subsets sorted form the spleen of wild-type animals (NK cells: CD45+/CD3-/NK1.1+; CD4+ T cells: CD45+/CD3+/CD4+; CD8+ T cells: CD45+/CD3+/CD8+; B cells: CD45+/B220+; PMN: CD45+/CD11bhi/Gr1hi; mDC: CD45+/CD11b+/CD11c+; pDC: CD45+/CD11b-/CD11c+; macrophages: CD45+/CD11b+/F4/80hi). Results are shown as mean ± SEM of Ms4a4a relative expression normalized to Gapdh (n = 3 mice). Statistical analysis by one-way ANOVA. b) MS4A4A mRNA levels in wild-type BMDMs exposed or not (M0) to 20 ng/ml IFN-γ plus 100 ng/ml LPS, 20 ng/ml IL-4, 10−6 M Dex, or 20 ng/ml IL-4 plus 10−6 M Dex for 18 h. Results are shown as fold change over resting macrophages (M0; n = 5). Statistical analysis by one-way ANOVA. c, Ms4a4a mRNA levels in BMDMs from wild-type (closed symbols) or Ms4a4a-/-(open symbols) mice treated (Dex) or not (NT) with 10−6 M Dex for 24 h. Results are shown as mean ± SEM of Ms4a4a relative expression normalized to Gapdh (n = 3 mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test and one-way ANOVA. d) Percentage of NK cells (CD45+/CD3-/NK1.1+), T cells (CD45+/CD3+), B cells (CD45+/B220+), PMN (CD45+/CD11bhi/Gr1hi), mDC (CD45+/CD11b+/CD11c+), pDC (CD45+/CD11b-/CD11c+), and macrophages (CD45+/CD11b+/F4/80hi) in the spleen of wild-type (closed symbols) or Ms4a4a-/- (open symbols) mice. Results are shown as mean ± SEM (n = 4). No statistical difference by two-tailed unpaired Student’s t test.

Supplementary Figure 5 Expression levels of MS4A4A and Dectin-1 in human monocyte-derived macrophages and DCs.

a, b) Flow cytometry analysis of Dectin-1 expression in monocyte-derived macrophages (Mɸ) and dendritic cells (DC) treated (Dex, open symbol in b) or not (NT, closed symbol in b) with 10−6 M Dex for 24 h. Ctrl represents isotype control staining. Results are shown as mean ± SEM. Three independent experiments were performed (n = 5 donors). Statistical analysis by one-way ANOVA. c, d) Flow cytometry analysis of Dectin-1 and MS4A4A expression in monocyte-derived macrophages (Mɸ) and dendritic cells (DC) treated (Dex, open symbol in d) or not (NT, closed symbol in d) with 10−6 M Dex for 24 h. Ctrl represents isotype control staining. Percentage of double-positive cells is shown in d. Results are shown as mean ± SEM. Three independent experiments were performed (n = 5 donors). Statistical analysis by one-way ANOVA.

Supplementary Figure 6 Ms4a4a impact on Dectin-1 signalling in BMDMs.

ac) Syk phosphorylation in BMDMs from Ms4a4a-/- (open symbol) and wild-type (closed symbol) mice primed for 18 h with 10 ng/ml GM-CSF and stimulated or not with 100 μg/ml depleted zymosan (a), 100 μg/ml curdlan (b), or 100 ng/ml PMA (d). Results are shown as mean ± SEM of relative MFI (fold on untreated). Three independent experiments were performed. a) time 5: 4 wild-type and 4 Ms4a4a-/- mice; time 15: 5 wild-type and 8 Ms4a4a-/- mice. b) time 5: 6 wild-type and 5 Ms4a4a-/- mice; time 15: 5 wild-type and 6 Ms4a4a-/- mice. c) time 5: 4 wild-type and 5 Ms4a4a-/- mice; time 15: 4 wild-type and 4 Ms4a4a-/- mice. Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test. d, e) Phosphorylation of ERK (d) and p38 (e) in BMDMs from Ms4a4a-/- (open symbols) and wild-type (closed symbols) mice primed for 18 h with 10 ng/ml GM-CSF and stimulated or not (NT) with 100 μg/ml depleted zymosan, 100 μg/ml curdlan, 100 ng/ml LPS, and 100 ng/ml PMA. Results are shown as mean ± SEM of relative MFI (fold on untreated). Two independent experiments were performed. One dot represents one mouse (2–4 for e, 2–6 for f). fk) Secretion of IL-6 (fh) and TNF (ik) by BMDMs from Ms4a4a-/- (open symbol) and wild-type (closed symbol) mice primed for 18 h with 10 ng/ml GM-CSF and stimulated for 24 h (f, g and i, j) or 6 h (hk) with 100 μg/ml depleted zymosan (fi), 100 μg/ml curdlan (gj) or 100 ng/ml Pam3Cys (hk). Cytokine levels in untreated cells were below detection limit. Results are shown as mean ± SEM. Two independent experiments for f and i (9 wild-type and 11 Ms4a4a-/- mice for f; 8 wild-type and 12 Ms4a4a-/- mice for i); five independent experiments for g and j (14 wild-type and 16 Ms4a4a-/- mice for g; 13 wild-type and 16 Ms4a4a-/- mice for j), three independent experiments for h and k (7 wild-type and 10 Ms4a4a-/- mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test.

Supplementary Figure 7 Role of Ms4a4a in B16F1 metastatic spreading and macrophage infiltration.

a) Representative images of lungs from Ms4a4a-/- and wild-type mice upon i.v. injection of B16F1 or B16F10 cells. b) Number of metastatic foci in lungs from Ms4a4a-/-Clec7a-/- (open symbol) and wild-type (closed symbol) mice upon B16F1 i.v. injection. Results are shown as mean ± SEM. Four independent experiments were performed (11 wild-type and 19 Ms4a4a-/-Clec7a-/-mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test. c) Representative immunofluorescence images of the distribution of NK cells (NKp46+), macrophages (F4/80+), and proliferating tumor cells (Ki-67+) in the lung of wild-type mice injected i.v. with B16F1 cells. Nuclei were counterstained with DAPI and visualized in blue. Images refer to one experiment out of two analyzed. Scale bars: 150 μm in the confocal images (left panels), 15 μm in the 3D bland images of representative areas (right panels). dg) Frequency (d, f) and numbers (e, g) of interstitial (d, e) and alveolar (f, g) macrophages in the lungs of Ms4a4a-/- (open symbol) and wild-type (closed symbol) mice injected i.v. with B16F1 cells. Resuls are shown as percentage or absolute numbers (n) of F4/80+/CD11c- (interstitial) or F4/80+/CD11c+ (alveolar) macrophages pregated on Live/CD45+/SSC-Ahi/CD11blo/hi cells that were Ly6C and Ly6G negative. Results are expressed as mean ± SEM. Three independent experiments were performed (9 wild-type and 6 Ms4a4a-/- mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test. h) NK cell degranulation towards B16F10 cells. Results are shown as mean ± SEM of percentage of CD107a+ cells gated on Live/CD45+/CD3-/NK1.1+ splenocytes. Two independent experiments were performed (5 wild-type and 4 Ms4a4a-/- mice). Statistical analysis by two tailed unpaired (Mann-Whitney), Student’s t test. ik) Calhm6 mRNA levels in Ms4a4a-/- (open symbol) and wild-type (closed symbol) BMDMs primed for 18 h with 10 ng/ml GM-CSF and stimulated for 24 h with zymosan (i), depleted zymosan (j), or LPS (k). Results are shown as mean ± SEM of fold of induction. Two independent experiments were performed for i (3 wild-type and 5 Ms4a4a-/- mice), j (7 wild-type and 9 Ms4a4a-/- mice), and k (10 wild-type and 5 Ms4a4a-/- mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test. l) IL-18 secretion by Ms4a4a-/- (open symbol) and wild-type (closed symbol) BMDMs primed for 18 h with 10 ng/ml GM-CSF and stimulated for 24 h with zymosan. Resuts are shown as mean ± SEM. Two independent experiments were performed (6 wild-type and 5 Ms4a4a-/- mice). Statistical analysis by two-tailed unpaired (Mann-Whitney), Student’s t test. IL-18 levels in untreated cells and IL-12p70 levels in untreated and stimulated cells were below detection limit and are not shown.

Supplementary Figure 8 Ms4a4a contribution to Dectin-1-mediated control of metastatic spread.

a) Representative dot plot (left) and Mean Fluorescence Intensity (right) of WGA binding on the surface of MC38 and SL4 cells assessed by flow cytometry. Results are shown as mean ± SEM of relative MFI (fold on unstained). Three independent experiments were performed (5 independent samples for SL4, 7 independent samples for MC38). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test. b, c) Number of metastatic foci in liver (b) and lungs (c) of Clec7a-/- (open symbol) and wild-type (closed symbol) mice upon intrasplenic injection of MC38 (b) or i.v. injection of SL4 (c) cells. Results are shown as mean ± SEM. One experiment was performed (5 wild-type and 6 Ms4a4a-/- mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test. d) Number of metastatic foci in liver of Ms4a4a-/- (open symbol) and wild-type (closed symbol) mice injected intrasplenic with MC38 cells upon NK cell depletion. Results are shown as mean ± SEM. Two independent experiments were performed (5 wild-type and 9 Ms4a4a-/- mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test and one-way ANOVA. e) Number of metastatic foci in lungs from Ms4a4a-/- (open symbol) and wild-type (closed symbol) mice upon i.v. injection of SL4 cells. Results are shown as mean ± SEM. One experiment was performed (6 mice). Statistical analysis by two-tailed unpaired (Mann-Whitney) Student’s t test.

Supplementary information

Supplementary Information

Supplementary Figs. 1–8 and Supplementary Tables 1–3

Reporting Summary

Supplementary Note

Gating strategies applied to Figures and Supplementary Figures are reported.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mattiola, I., Tomay, F., De Pizzol, M. et al. The macrophage tetraspan MS4A4A enhances dectin-1-dependent NK cell–mediated resistance to metastasis. Nat Immunol 20, 1012–1022 (2019). https://doi.org/10.1038/s41590-019-0417-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-019-0417-y

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