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:

Selective multi-kinase inhibition sensitizes mesenchymal pancreatic cancer to immune checkpoint blockade by remodeling the tumor microenvironment

Abstract

KRAS-mutant pancreatic ductal adenocarcinoma (PDAC) is highly immunosuppressive and resistant to targeted and immunotherapies. Among the different PDAC subtypes, basal-like mesenchymal PDAC, which is driven by allelic imbalance, increased gene dosage and subsequent high expression levels of oncogenic KRAS, shows the most aggressive phenotype and strongest therapy resistance. In the present study, we performed a systematic high-throughput combination drug screen and identified a synergistic interaction between the MEK inhibitor trametinib and the multi-kinase inhibitor nintedanib, which targets KRAS-directed oncogenic signaling in mesenchymal PDAC. This combination treatment induces cell-cycle arrest and cell death, and initiates a context-dependent remodeling of the immunosuppressive cancer cell secretome. Using a combination of single-cell RNA-sequencing, CRISPR screens and immunophenotyping, we show that this combination therapy promotes intratumor infiltration of cytotoxic and effector T cells, which sensitizes mesenchymal PDAC to PD-L1 immune checkpoint inhibition. Overall, our results open new avenues to target this aggressive and therapy-refractory mesenchymal PDAC subtype.

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: Resistance to MEK inhibition in vitro and in vivo.
Fig. 2: Systematic combination drug screens identify new therapies for nonglandular mesenchymal PDAC.
Fig. 3: Genetic screens uncover nintedanib targets that sensitize mesenchymal PDAC toward trametinib.
Fig. 4: The combination treatment prolongs survival and reprograms the TME in vivo.
Fig. 5: The combination treatment enhances tumor immune infiltration specifically in the mesenchymal subtype.
Fig. 6: The T/N combination sensitizes mesenchymal PDAC toward anti-PD-L1 ICB.
Fig. 7: ScRNA-seq analysis reveals context-specific responses of tumor cells and their microenvironment on combination drug treatment.
Fig. 8: The combination therapy induces a T-cell-mediated anti-tumor immune response in mesenchymal PDAC.

Similar content being viewed by others

Data availability

The RNA-seq dataset has been deposited in the EBIArrayExpress repository with accession no. E-MTAB-11187. The MS kinobead pulldown and the MS secretomics data have been deposited in the ProteomeXchange Consortium (http://proteomecentral.proteomexchange.org) via the PRIDE partner repository with accession nos. PXD023267 and PXD027877, respectively. The scRNA-seq data have been deposited in the EBIArrayExpress repository with accession no. E-MTAB-9954. The human pancreatic cancer data were derived from previous studies and are available in the supplementary information of the respective publications6,7,8. All other data have been provided as supplementary tables or source data files. Mice and cell lines are available from the corresponding author on reasonable request. Key resources are listed in Supplementary Table 8. Source data are provided with this paper.

Code availability

Analyses were performed using open-source software, and in-house scripts in R v.3.6.2 and Python v.3.8.3, which are available from the corresponding author on reasonable request. Codes are provided via the GitHub repository at the following link: https://github.com/stefanie-baerthel/combinatorial_treatment_analysis.

References

  1. Quaresma, M., Coleman, M. P. & Rachet, B. 40-year trends in an index of survival for all cancers combined and survival adjusted for age and sex for each cancer in England and Wales, 1971–2011: a population-based study. Lancet 385, 1206–1218 (2015).

    Article  PubMed  Google Scholar 

  2. Siegel, R. L., Miller, K. D. & Jemal, A. Cancer statistics, 2020. CA Cancer J. Clin. 70, 7–30 (2020).

    Article  PubMed  Google Scholar 

  3. Aung, K. L. et al. Genomics-driven precision medicine for advanced pancreatic cancer: early results from the COMPASS trial. Clin. Cancer Res. 24, 1344–1354 (2018).

    Article  CAS  PubMed  Google Scholar 

  4. Kalimuthu, S. N. et al. Morphological classification of pancreatic ductal adenocarcinoma that predicts molecular subtypes and correlates with clinical outcome. Gut 69, 317–328 (2020).

    Article  Google Scholar 

  5. Chan-Seng-Yue, M. et al. Transcription phenotypes of pancreatic cancer are driven by genomic events during tumor evolution. Nat. Genet. 52, 231–240 (2020).

    Article  CAS  PubMed  Google Scholar 

  6. Bailey, P. et al. Genomic analyses identify molecular subtypes of pancreatic cancer. Nature 531, 47–52 (2016).

    Article  CAS  PubMed  Google Scholar 

  7. Dijk, F. et al. Unsupervised class discovery in pancreatic ductal adenocarcinoma reveals cell-intrinsic mesenchymal features and high concordance between existing classification systems. Sci. Rep. 10, 337 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Puleo, F. et al. Stratification of pancreatic ductal adenocarcinomas based on tumor and microenvironment features. Gastroenterology 155, 1999–2013.e1993 (2018).

    Article  PubMed  Google Scholar 

  9. Morrison, A. H., Byrne, K. T. & Vonderheide, R. H. Immunotherapy and prevention of pancreatic cancer. Trends Cancer 4, 418–428 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Balachandran, V. P. et al. Identification of unique neoantigen qualities in long-term survivors of pancreatic cancer. Nature 551, 512–516 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Chen, D. S. & Mellman, I. Elements of cancer immunity and the cancer-immune set point. Nature 541, 321–330 (2017).

    Article  CAS  PubMed  Google Scholar 

  12. Ino, Y. et al. Immune cell infiltration as an indicator of the immune microenvironment of pancreatic cancer. Br. J. Cancer 108, 914–923 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Collisson, E. A. et al. A central role for RAF–>MEK–>ERK signaling in the genesis of pancreatic ductal adenocarcinoma. Cancer Discov. 2, 685–693 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Blumenschein, G. R. Jr et al. A randomized phase II study of the MEK1/MEK2 inhibitor trametinib (GSK1120212) compared with docetaxel in KRAS-mutant advanced non-small-cell lung cancer (NSCLC). Ann. Oncol. 26, 894–901 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Caunt, C. J., Sale, M. J., Smith, P. D. & Cook, S. J. MEK1 and MEK2 inhibitors and cancer therapy: the long and winding road. Nat. Rev. Cancer 15, 577–592 (2015).

    Article  CAS  PubMed  Google Scholar 

  16. Mueller, S. et al. Evolutionary routes and KRAS dosage define pancreatic cancer phenotypes. Nature 554, 62–68 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Miyabayashi, K. et al. Intraductal transplantation models of human pancreatic ductal adenocarcinoma reveal progressive transition of molecular subtypes. Cancer Discov.10, 1566–1589 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Hänzelmann, S., Castelo, R. & Guinney, J. GSVA: gene set variation analysis for microarray and RNA-Seq data. BMC Bioinformatics 14, 7 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Liberzon, A. et al. The molecular signatures database Hallmark gene set collection. Cell Systems 1, 417–425 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Schönhuber, N. et al. A next-generation dual-recombinase system for time- and host-specific targeting of pancreatic cancer. Nat. Med. 20, 1340–1347 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Ruscetti, M. et al. Senescence-induced vascular remodeling creates therapeutic vulnerabilities in pancreas cancer. Cell 181, 424–441.e421 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. DeJarnette, J. B. et al. Specific requirement for CD3epsilon in T cell development. Proc. Natl Acad. Sci. USA 95, 14909–14914 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Pathria, P., Louis, T. L. & Varner, J. A. Targeting tumor-associated macrophages in cancer. Trends Immunol. 40, 310–327 (2019).

    Article  CAS  PubMed  Google Scholar 

  24. Nywening, T. M. et al. Targeting both tumour-associated CXCR2+ neutrophils and CCR2+ macrophages disrupts myeloid recruitment and improves chemotherapeutic responses in pancreatic ductal adenocarcinoma. Gut 67, 1112–1123 (2018).

    Article  CAS  PubMed  Google Scholar 

  25. Steele, C. W. et al. CXCR2 inhibition profoundly suppresses metastases and augments immunotherapy in pancreatic ductal adenocarcinoma. Cancer Cell 29, 832–845 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Gerrard, T. L., Cohen, D. J. & Kaplan, A. M. Human neutrophil-mediated cytotoxicity to tumor cells. J. Natl Cancer Instit. 66, 483–488 (1981).

    CAS  Google Scholar 

  27. Bakhoum, S. F. & Cantley, L. C. The multifaceted role of chromosomal instability in cancer and its microenvironment. Cell 174, 1347–1360 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Mackenzie, K. J. et al. cGAS surveillance of micronuclei links genome instability to innate immunity. Nature 548, 461–465 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Faget, D. V., Ren, Q. & Stewart, S. A. Unmasking senescence: context-dependent effects of SASP in cancer. Nat. Rev. Cancer 19, 439–453 (2019).

    Article  CAS  PubMed  Google Scholar 

  30. Meissner, F., Scheltema, R. A., Mollenkopf, H.-J. & Mann, M. Direct proteomic quantification of the secretome of activated immune cells. Science 340, 475–478 (2013).

    Article  CAS  PubMed  Google Scholar 

  31. Matsumura, S. et al. Radiation-induced CXCL16 release by breast cancer cells attracts effector T cells. J. Immunol. 181, 3099–3107 (2008).

    Article  CAS  PubMed  Google Scholar 

  32. Hojo, S. et al. High-level expression of chemokine CXCL16 by tumor cells correlates with a good prognosis and increased tumor-infiltrating lymphocytes in colorectal cancer. Cancer Res. 67, 4725–4731 (2007).

    Article  CAS  PubMed  Google Scholar 

  33. Li, B. et al. Epigenetic regulation of CXCL12 plays a critical role in mediating tumor progression and the immune response in osteosarcoma. Cancer Res. 78, 3938–3953 (2018).

    Article  CAS  PubMed  Google Scholar 

  34. Nagarsheth, N., Wicha, M. S. & Zou, W. Chemokines in the cancer microenvironment and their relevance in cancer immunotherapy. Nat. Rev. Immunol. 17, 559–572 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Mehta, A. K. et al. Targeting immunosuppressive macrophages overcomes PARP inhibitor resistance in BRCA1-associated triple-negative breast cancer. Nat. Cancer 2, 66–82 (2021).

    Article  CAS  PubMed  Google Scholar 

  36. Peng, D. H. et al. Th17 cells contribute to combination MEK inhibitor and anti-PD-L1 therapy resistance in KRAS/p53 mutant lung cancers. Nat. Commun. 12, 2606 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Di Pilato, M. et al. CXCR6 positions cytotoxic T cells to receive critical survival signals in the tumor microenvironment. Cell 184, 4512–4530.e4522 (2021).

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  39. Steins, A. et al. High-grade mesenchymal pancreatic ductal adenocarcinoma drives stromal deactivation through CSF-1. EMBO Rep. 21, e48780 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Sahai, E. et al. A framework for advancing our understanding of cancer-associated fibroblasts. Nat. Rev. Cancer 20, 174–186 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Schneider, G., Schmidt-Supprian, M., Rad, R. & Saur, D. Tissue-specific tumorigenesis: context matters. Nat. Rev. Cancer 17, 239–253 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Elyada, E. et al. Cross-species single-cell analysis of pancreatic ductal adenocarcinoma reveals antigen-presenting cancer-associated fibroblasts. Cancer Discov. 9, 1102–1123 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Hosein, A. N., Brekken, R. A. & Maitra, A. Pancreatic cancer stroma: an update on therapeutic targeting strategies. Nat. Rev. Gastroenterol. Hepatol. 17, 487–505 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Ligorio, M. et al. Stromal microenvironment shapes the intratumoral architecture of pancreatic cancer. Cell 178, 160–175.e127 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Batlle, E. & Massagué, J. Transforming growth factor-β signaling in immunity and cancer. Immunity 50, 924–940 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Olive, K. P. et al. Inhibition of Hedgehog signaling enhances delivery of chemotherapy in a mouse model of pancreatic cancer. Science 324, 1457–1461 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Sherman, M. H. et al. Vitamin D receptor-mediated stromal reprogramming suppresses pancreatitis and enhances pancreatic cancer therapy. Cell 159, 80–93 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Hayashi, A. et al. A unifying paradigm for transcriptional heterogeneity and squamous features in pancreatic ductal adenocarcinoma. Nat. Cancer 1, 59–74 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Brzostek-Racine, S., Gordon, C., Van Scoy, S. & Reich, N. C. The DNA damage response induces IFN. J. Immunol. 187, 5336–5345 (2011).

    Article  CAS  PubMed  Google Scholar 

  50. Zhou, F. Molecular mechanisms of IFN-γ to up-regulate MHC class I antigen processing and presentation. Int. Rev. Immunol. 28, 239–260 (2009).

    Article  CAS  PubMed  Google Scholar 

  51. Respa, A. et al. Association of IFN-γ signal transduction defects with impaired HLA class I antigen processing in melanoma cell lines. Clin. Cancer Res. 17, 2668–2678 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kearney, C. J. et al. Tumor immune evasion arises through loss of TNF sensitivity. Sci. Immunol. 3, eaar3451 (2018).

    Article  PubMed  Google Scholar 

  53. Liao, W. et al. KRAS-IRF2 axis drives immune suppression and immune therapy resistance in colorectal cancer. Cancer Cell 35, 559–572.e557 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Tape, C. J. et al. Oncogenic KRAS regulates tumor cell signaling via stromal reciprocation. Cell 165, 910–920 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Kuilman, T. & Peeper, D. S. Senescence-messaging secretome: SMS-ing cellular stress. Nat. Rev. Cancer 9, 81–94 (2009).

    Article  CAS  PubMed  Google Scholar 

  56. Lederer, D. J. & Martinez, F. J. Idiopathic pulmonary fibrosis. N. Engl. J. Med. 378, 1811–1823 (2018).

    Article  CAS  PubMed  Google Scholar 

  57. Richeldi, L. et al. Efficacy and safety of nintedanib in idiopathic pulmonary fibrosis. N. Engl. J. Med. 370, 2071–2082 (2014).

    Article  PubMed  Google Scholar 

  58. von Burstin, J. et al. E-cadherin regulates metastasis of pancreatic cancer in vivo and is suppressed by a SNAIL/HDAC1/HDAC2 repressor complex. Gastroenterology 137, 361–371.e361-365 (2009).

    Article  Google Scholar 

  59. Eser, S. et al. Selective requirement of PI3K/PDK1 signaling for Kras oncogene-driven pancreatic cell plasticity and cancer. Cancer Cell 23, 406–420 (2013).

    Article  CAS  PubMed  Google Scholar 

  60. Ianevski, A., Giri, A. K. & Aittokallio, T. SynergyFinder 2.0: visual analytics of multi-drug combination synergies. Nucleic Acids Res. 48, W488–W493 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Klaeger, S. et al. The target landscape of clinical kinase drugs. Science https://doi.org/10.1126/science.aan4368 (2017).

  62. Shevchenko, A., Tomas, H., Havli, J., Olsen, J. V. & Mann, M. In-gel digestion for mass spectrometric characterization of proteins and proteomes. Nat. Protoc. 1, 2856–2860 (2006).

    Article  CAS  PubMed  Google Scholar 

  63. Vizcaíno, J. A. et al. The PRoteomics IDEntifications (PRIDE) database and associated tools: status in 2013. Nucleic Acids Res. 41, D1063–D1069 (2013).

    Article  PubMed  Google Scholar 

  64. Hafner, M., Niepel, M., Chung, M. & Sorger, P. K. Growth rate inhibition metrics correct for confounders in measuring sensitivity to cancer drugs. Nat. Methods 13, 521–527 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Clark, N. A. et al. GRcalculator: an online tool for calculating and mining dose-response data. BMC Cancer 17, 698 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Bindea, G. et al. ClueGO: a Cytoscape plug-in to decipher functionally grouped gene ontology and pathway annotation networks. Bioinformatics 25, 1091–1093 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. DeWeirdt, P. C. et al. Optimization of AsCas12a for combinatorial genetic screens in human cells. Nat. Biotechnol. 39, 94–104 (2021).

    Article  CAS  PubMed  Google Scholar 

  68. Joung, J. et al. Genome-scale CRISPR–Cas9 knockout and transcriptional activation screening. Nat. Protoc. 12, 828–863 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Li, W. et al. MAGeCK enables robust identification of essential genes from genome-scale CRISPR/Cas9 knockout screens. Genome Biol. 15, 554 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Doench, J. G. et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR–Cas9. Nat. Biotechnol. 34, 184–191 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Jackson, E. L. et al. Analysis of lung tumor initiation and progression using conditional expression of oncogenic K-ras. Genes Dev. 15, 3243–3248 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Hingorani, S. R. et al. Preinvasive and invasive ductal pancreatic cancer and its early detection in the mouse. Cancer Cell 4, 437–450 (2003).

    Article  CAS  PubMed  Google Scholar 

  73. Nakhai, H. et al. Ptf1a is essential for the differentiation of GABAergic and glycinergic amacrine cells and horizontal cells in the mouse retina. Development 134, 1151–1160 (2007).

    Article  CAS  PubMed  Google Scholar 

  74. Catalanotti, F. et al. A Mek1–Mek2 heterodimer determines the strength and duration of the Erk signal. Nat. Struct. Mol. Biol. 16, 294–303 (2009).

    Article  CAS  PubMed  Google Scholar 

  75. Wolf, F. A., Angerer, P. & Theis, F. J. SCANPY: large-scale single-cell gene expression data analysis. Genome Biol. 19, 15 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Efremova, M., Vento-Tormo, M., Teichmann, S. A. & Vento-Tormo, R. CellPhoneDB: inferring cell–cell communication from combined expression of multi-subunit ligand–receptor complexes. Nat. Protoc. 15, 1484–1506 (2020).

    Article  CAS  PubMed  Google Scholar 

  77. Phulphagar, K. et al. Proteomics reveals distinct mechanisms regulating the release of cytokines and alarmins during pyroptosis. Cell Rep. 34, 108826 (2021).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank H. Nakhei, T. Jacks, D. Tuveson, M. Baccarini, R. Schmid and A. Bradley for providing transgenic animals, and the TUM animal facility and imaging core facility of the Department of Nuclear Medicine, Klinikum rechts der Isar, for excellent technical support. The present study was supported by the German Cancer Consortium (DKTK), Deutsche Forschungsgemeinschaft (DFG SA 1374/4-2, SFB 824 C09 to G.S. and D.S., SFB 1321 Project-ID 329628492 P06 to D.S. and M.S.S, P11 to M.S.R. and D.S., and S01 to D.S., M.R., R.R. and G.S., and SFB 1371 Project-ID 395357507 P12 to D.S.), the Wilhelm Sander-Stiftung (2020.174.1), and the European Research Council (ERC CoG No. 648521, to D.S.).

Author information

Authors and Affiliations

Authors

Contributions

C.F., S.B., S.A.W., C.S., J.J.M., A.T., J.M., T.K., J.H., J.J.S., B.T., O.R., C.S., K.G., A.C., C.V., M.Z., A.A.V., W.H.P., R.M., R.Ö., T.A. and J.R. performed the research. M.S.R., B.K., K.S., F.M., M.R., M.F., R.R., M.S.S. and D.S. contributed new reagents and analytical tools. C.F., S.B., S.A.W., C.S., J.M., J.J.S., J.C., A.C., F.B., M.J., K.S., O.B., J.R, F.M., M.F., R.R., M.S.S., G.S. and D.S. analyzed the data. C.F. and D.S. wrote the paper.

Corresponding author

Correspondence to Dieter Saur.

Ethics declarations

Competing interests

The authors declare that they have no competing interests. Correspondence and requests for materials should be addressed to D.S.

Peer review

Peer review information

Nature Cancer thanks Michael Hemann and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 Assessment of differential pharmacologic and genetic dependencies and signaling pathway activities in PDAC subtypes.

a, Clonogenic assay of two hPDAC cell lines (top) and two mPDAC cell cultures (bottom) treated with the MEK inhibitor trametinib. The shown cell lines represent the drug-response of the epithelial and mesenchymal subtypes to trametinib treatment. b, c, β-score distribution of CRISPR/Cas9 genome wide negative-selection (viability) screens performed in one classical (6075, panel (b)) and one mesenchymal (9091, panel (c)) mPDAC cell line. Highlighted in yellow, for the classical line, and blue, for the mesenchymal line, are the β-scores of KRAS and the core genes involved in direct KRAS downstream signaling. d, CRISPR/Cas9 dependency scores of KRAS and core genes involved in direct KRAS downstream signaling. The dependency scores of all hPDAC cell lines were obtained from the DepMap database and are shown in grey. Dependency scores corresponding to classical and mesenchymal cell lines included in the T/N drug screen are represented in the yellow and blue violin plots. Data were obtained from the CRISPR dataset and analyzed through the DepMap release DepMap 21Q2 Public (https://depmap.org/portal/download/). e, f, Mesenchymal (9091) and classical (8661) PDAC cell cultures were used to generate site-specific phospho-array datasets (Phospho Explorer antibody microarray, Full Moon Biosystems). Phospho-array data (supplementary table 1) were used to test for the enrichment of differentially phosphorylated sites between classical and mesenchymal mPDAC cell lines. Functionally grouped networks with reactome terms as nodes, showing pathways overrepresented in classical (e) and mesenchymal (f) cells are represented with the ClueGO plugin of Cytoscape. Only the pathways with an adjusted p value (calculated by two-sided hypergeometric test, Bonferroni corrected) ≤ 0.05 are depicted. The node size represents the term enrichment significance.

Source data

Extended Data Fig. 2 Genetic depletion of Mek1/2 in established PDAC.

a, Genetic strategy to delete Mek1 and Mek2 by 4-hydroxytamoxifen (4OHT)-mediated CreERT2 activation. Pdx1-Flp;FSF-KrasG12D/+;FSF-R26CAG-CreERT2/+ mice were crossed with mice harboring loxP-flanked Mek1 and Mek2 alleles. This allowed MEK1/2 deletion in established PDAC by tamoxifen administration in vitro and in vivo after orthotopic transplantation. b, Genotyping PCR of PDAC cells with indicated genotypes to analyze recombination of the floxed Mek1 allele. Non-recombined mutant, recombined mutant and wild-type PCR products are indicated on the right side. Representative gel of three independent experiments. c, Western blot analysis of MEK1 and MEK2 expression in primary PDAC cell cultures with indicated genotypes after 4 days of tamoxifen (4-OHT) and vehicle (ethanol, EtOH) treatment. HSP90 served as loading control. Representative gel of three independent experiments. d, Clonogenic assays of mPDAC cells with indicated genotypes. Control cells treated with vehicle (ethanol; EtOH) are shown in the upper row, 4-OHT treated cells in the lower row. e, Schematic representation of the experimental set-up to test the effect of Mek1/2 knockout in vivo by tamoxifen administration using syngeneic immunocompetent PDAC models. mPDAC cells with conditional Mek1lox/lox;Mek2lox/lox alleles were used for the orthotopic transplantation experiments. f, Waterfall plot showing tumor response of vehicle and tamoxifen treated animals after one week of treatment (fold-change compared to baseline before treatment based on MRI-volumetric measurements, y-axis). Each bar represents one mouse. P values calculated with two-tailed unpaired t test. g, Kaplan-Meier survival curve of vehicle and tamoxifen treated PDAC models. Number of mice is indicated in the corresponding panels. P value was calculated with log-rank (Mantel-Cox) test. h, Representative images of HE and IHC for MEK1, MEK2 and KI67 of tissue sections of tumors from orthotopically transplanted Mek1lox/lox;Mek2lox/lox models treated with vehicle or tamoxifen. Representative pictures of three independent experiments. Scale bars, 100 µm.

Source data

Extended Data Fig. 3 Pharmacologic and genetic modulation of drug sensitivity in classical and mesenchymal PDAC cell cultures.

a, b, Clonogenic assays of a representative human (left) and mouse (right) PDAC cell culture showing antagonism to the trametinib/nintedanib (T/N) combination. Cell cultures were treated with indicated concentrations of T/N. c, Western blot of phospho-ERK and ERK in T/N (10 nM trametinib + 2 µM nintedanib) and vehicle treated classical and mesenchymal primary mPDAC cell lines. HSP90 served as loading control. Classical cell lines are marked in yellow, mesenchymal in blue. Representative gels of three independent experiments. d, Clonogenic assays using increased drug concentrations of the T/N combination of three of the most antagonistic cell lines, as depicted in Fig. 2, panel (g). e, Doxycycline-induced overexpression of KRASG12D in mouse PDAC cells. 2259 mPDAC cells representative of the classical subtype was transduced with lentivirus carrying doxycycline-inducible KRASG12D or GFP-control expression constructs. KRASG12D or GFP expression were induced by doxycycline (100 ng/ml) for one or 14 days. f, Western blot of phospho-ERK and total ERK in cells overexpressing KRASG12D or GFP for one day. HSP90 served as loading control. g, Expression of the marker gene Cdh1 for epithelial cell differentiation was evaluated by qRT-PCR (normalized to Cyclophilin B). Data are shown as mean ± SD; n = 3 biological replicates. P value was calculated with two-tailed unpaired t test. h, Representative picture of three independent experiments of the morphological changes of PDAC cells upon KRASG12D induction for one or 14 days of doxycycline treatment. Scale bars, 200 µm. i, Representative clonogenic assays of mPDAC cells treated with the indicated concentrations of trametinib and nintedanib upon KRASG12D (right panel) or GFP (left panel) overexpression. j, Bliss synergy scores for the mPDAC cell line treated with the combination of trametinib and nintedanib upon KRASG12D or GFP overexpression.

Source data

Extended Data Fig. 4 Kinobead-based proteomic identification of the trametinib and nintedanib targets and treatment-induced changes in the phosphoproteome of classical and mesenchymal PDAC.

a, b, Representative pictures of the target space of trametinib (a) and nintedanib (b) for 2259 PDAC cells. A phylogenetic tree of all kinases for the 2259 primary mouse PDAC cell culture is shown. The indicated circle sizes indicate potency (apparent dissociation constants (Kdapp)), the color code specifies protein-drug interaction with the designated or other targets. Arrows highlight the identified targets. c, Radar plot showing the overlay of the pKd (−log10Kd) for the targets of nintedanib in the 6 PDAC cell cultures tested. PDAC cells of the classical (n = 4) and mesenchymal (n = 2) subtype are indicated with the color code. d, Heatmap showing the differentially expressed genes between epithelial and mesenchymal cell cultures identified as targets of nintedanib. The color code indicates the Z score. e, Volcano plots representing the change in gene expression of the nintedanib targets (in blue) upon trametinib treatment. The x-axis log2 fold change (treated/control), the y-axis shows the per test adjusted p values, which were calculated by differential expression test (using the DESeq2 package). A gene was considered to be differentially expressed with a Benjamini-Hochberg adjusted p-value of 0.05 and an absolute fold change >1. f, g, Mesenchymal (9091) and classical (8661) PDAC cell cultures were used to generate site-specific phospho-array datasets (Phospho Explorer antibody microarray, Full Moon Biosystems). The cell lines were analyzed at basal condition and in presence of T/N (trametinib 10 nM + nintedanib 2 μM). Phospho-array data (Supplementary Table 3) were used to test for the decrease of differentially phosphorylated sites between T/N and vehicle (DMSO) treated classical and mesenchymal mPDAC cells. Functionally grouped maps, obtained with the ClueGO plugin of Cytoscape, representing pathways overrepresented in mesenchymal (f) and classical (g) mPDAC upon T/N treatment are shown. Only the pathways with an adjusted p value (calculated by two-sided hypergeometric test, Bonferroni corrected) ≤ 0.05 are represented. The node size represents the term enrichment significance.

Extended Data Fig. 5 Pharmacologic assessment of nintedanib targets.

a, Combinatorial drug screen on mesenchymal hPDAC cell line MiaPaca2 and mPDAC cell line 9091, as shown in Fig. 2, panel (b). The MEK inhibitor trametinib was used in fixed concentration and combined with 418 additional drugs under preclinical and clinical investigation. Highlighted in orange are the drugs in the high-throughput drug screen showing overlapping targets with nintedanib. b, Venn diagrams showing the target overlap between the drugs identified in (a) and nintedanib (see Supplementary Table 4) as reported from the ProteomicsDB database (https://www.proteomicsdb.org). c, Venn diagrams showing the target overlap between nintedanib and additional drugs with an overlapping target profile chosen for further target assessment. The overlapping targets are listed below each figure. The target information was downloaded from the ProteomicsDB database (https://www.proteomicsdb.org). d, Representative clonogenic assays of mesenchymal mPDAC cell cultures treated with trametinib in combination with the drugs shown in (c) as compared to nintedanib. The cell lines were treated with the indicated concentrations of trametinib and the indicated experimental drug. e, Bliss synergy scores of clonogenic assays shown in (d) integrated with cell morphology for the treated mPDAC cell cultures (classical subtype depicted in yellow and mesenchymal in blue).

Source data

Extended Data Fig. 6 Genetic screens to identify relevant nintedanib targets sensitizing mesenchymal PDAC towards trametinib.

a, Western blot of Cas9 expression in the clonal cell lines used for CRISPR/Cas9 screens. β-Actin served as loading control. b, Editing efficiency at the Hprt locus. c, Relative viability upon 6-Thioguanine treatment to validate Cas9 function in Hprt proficient and deficient Cas9 cells (mean ± SD; n = 3 biological replicates). d, Relative cell growth (y-axis), assessed by cell counting, in the presence of different concentrations of trametinib (mean ± SD; n = 3 biological replicates). The pink line indicates the trametinib concentration used for the CRISPR/Cas9 screens. e, Phospho-ERK, ERK and Cas9 Western blots of clones used for CRISPR/Cas9 screens. Cells were treated with DMSO or trametinib (1.25 nM, 2.5 nM, 5 nM, 10 nM and 20 nM) for 4 days. HSP90 served as loading control. f, Focused CRISPR/Cas9-based genetic screening in mesenchymal mPDAC cells 8248 and 8570. Trametinib sensitivity (x-axis) represents β-scores calculated as sgRNA representation difference between trametinib-treated cells and their initial representation. Differential sensitivity (y-axis) indicates β-score differences between trametinib- and DMSO-treated arms. In red, genes presenting differential sensitivity ≤ −0.25. g, Network visualization of normalized gene expression (assessed by RNA-seq) of nintedanib targets shown in Fig. 3 (d). h, Lentiviral CRISPR/Cas9-mediated deletion of selected top-scoring nintedanib targets in 8248 and 8570 cells. Knock-out cells were treated with trametinib (5 nM) or DMSO and viability was assessed through clonogenic assays. i, Quantification of panel (h). Data are normalized to DMSO-treated non-targeting controls (mean ± SD; n = 3 biological replicates). The dashed line represents the mean of trametinib-treated non-targeting controls. j, Editing efficiency of each sgRNA used in Fig. 2 (f and g) and in panels (h) and (i) of this figure. k, Combinatorial deletion of nintedanib targets via ribonucleoprotein (RNP) electroporation. Mesenchymal 8248 and 8570 knock-out cells were treated with trametinib (5 nM) or DMSO and viability was assessed through clonogenic assays. l, Quantification of panel (k). Data are normalized to DMSO-treated non-targeting controls (mean ± SD; n = 3 biological replicates). Dashed line represents the mean of trametinib-treated non-targeting controls. The shown gels are representative of three independent experiments.

Source data

Extended Data Fig. 7 Characterization of context-specific changes of the tumor vasculature and the adaptive immune system in classical and mesenchymal PDAC subtypes upon therapy.

a, Orthotopically transplanted tumors of the indicated subtypes were treated with vehicle (control) and the T/N combination. Representative images of immunofluorescence stainings of tissue sections for P-selectin (upper panel) and α-SMA (lower panel) (magenta). CD31 was used to detect endothelial cells (green). DAPI was used for nuclear staining (blue). Scale bars, 25 µm. b, c, Quantification of the P-selectin+ vessels (b) and α-SMA + vessels (c) of the immunofluorescence stainings depicted in (a). Individual tumors are shown as single dots in the graph (classical: control n = 3, T/N n = 4; mesenchymal: control n = 4, T/N n = 5). d, Orthotopically transplanted tumors of the indicated subtypes were treated with vehicle (control) and the indicated drugs and drug combinations, explanted, single cell suspended and analyzed by flow cytometry. Panel (d) shows the staining for CD45 + cells. Individual tumors are shown as single points in the graph. e, f, Graphs representing the percentage of CD4 + (e) and CD8 + (f) cells in the PDAC control cohort and in the different treatment conditions as analyzed by flow cytometry. Single points represent individual tumors. g, Left, scheme of the in vivo experimental strategy using orthotopic PDAC cell transplantations into T cell deficient CD3ε knockout (KO) mice. Right, representative FACS plot of immunodeficient CD3ε-KO and wild-type C57BL/6 mice, highlighting the lack of T cells in the CD3ε-KO animals. h, Representative MRI picture of vehicle (control) and T/N treated PDAC bearing CD3ε-KO mice before (week 2) and after 1 week treatment (week 3). P values in (b), (c), (d) and (f) were calculated with two-tailed unpaired t test. T: trametinib, N: nintedanib, T/N: trametinib+nintedanib.

Source data

Extended Data Fig. 8 Characterization of context-specific changes of the innate immune system in classical and mesenchymal PDAC subtypes upon therapy.

a, Pie charts representative of the fraction of innate immune cell populations in PDAC from vehicle control and mice that received the combination (T/N) for both classical and mesenchymal orthotopically transplanted tumors as analyzed by flow cytometry. The number of tumors per condition analyzed is depicted in the corresponding panel. b, Graphs representing the percentage of Ly6G- CD11b + F4/80+ macrophages in PDAC of the control cohort and in the treatment conditions as analyzed by flow cytometry. Single points represent individual tumors. c, Representative immunofluorescence staining for CD80/CD68 and ARG1/CD68 cells in both classical and mesenchymal tumors treated with the T/N combination therapy or vehicle as control. Scale bars, 50 µm. d, Quantification of the M1-like polarization macrophage markers CD80/CD68 + and INOS/CD68 + from the immunofluorescence stainings depicted in (c). Individual tumors are shown as single points in the graph (classical: control n = 5, T/N n = 5; mesenchymal: control n = 4, T/N n = 5). e, Quantification of the M2-like polarization macrophage markers ARG1/CD68 + and MRC1/CD68 + from the staining depicted in (c). Individual tumors are shown as single points in the graph (classical: control n = 5, T/N n = 5; mesenchymal: control n = 4, T/N n = 5). f, Graphs representing the percentage of Ly6G + CD11b + neutrophils in the control cohort and in the treatment conditions as analyzed by flow cytometry. Single points represent individual tumors. P values in (d) and (e) were calculated with two-tailed unpaired t test. T: trametinib, N: nintedanib, T/N: trametinib+nintedanib.

Source data

Extended Data Fig. 9 scRNA-seq reveals treatment-induced changes in TME cell subpopulations and activation of the DNA damage pathway in cancer cells.

a, Dotplot displaying marker gene expression across each identified cluster of cancer cells and corresponding tumor microenvironment for both classical and mesenchymal tumors. The clusters are indicated on the y axis and the main markers for each identified population are indicated on the x axis. b, Left, UMAP plot showing all identified cell populations within the scRNA-seq experiment. Middle, UMAP plot showing classical (yellow) and mesenchymal (blue) tumors from all treatment and vehicle groups. Right, UMAP plot showing the treatment-induced changes in cell type composition among the identified cell populations across subtypes. Lower part, UMAP density plots showing distribution of annotated clusters upon treatment, cell numbers for each condition are integrated below. c, UMAP plot showing the identified tumor cell clusters. The expression of Cdh1 and Krt18, epithelial markers, and of Cdh2 and Vim, mesenchymal markers, across treatment conditions are shown below. d, Heatmap of the most differentially expressed genes from the gene expression signature in Fig. 7 across subtypes and treatment conditions. e, Gene set enrichment analysis (GSEA) of scRNA-seq data of cancer cells reveals enrichment of DNA damage in both classical and mesenchymal tumors upon treatment with the T/N combination. NES and FDR-q values are indicated. f, Representative images of immunohistochemical staining for γH2AX of control and T/N treated tumor sections for both classical and mesenchymal subtypes. Scale bar, 70 µm. g, Quantification of γH2AX positive cells in (f). Individual tumors are shown as single points in the graph (classical: control n = 6, T/N n = 5; mesenchymal: control n = 8, T/N n = 7). P values were calculated with two-tailed unpaired t test. Endo cells: endothelial cells. T/N: trametinib+nintedanib. T/N + aPD-L1: trametinib+nintedanib+anti PD-L1 antibody.

Source data

Extended Data Fig. 10 Context-dependent reprogramming of the cancer cell derived secretome and cancer associated fibroblasts (CAFs) by the T/N combination therapy.

a, Volcano plots highlighting the changes in secreted factors upon T/N treatment in classical (left) and mesenchymal (right) PDAC cells. The x-axis shows log2 fold change (treated/control), the y-axis the per test adjusted p values, which were calculated by differential expression test (two-sided t test). b, Circos plot showing the key communication signals from tumor cells to T cell subtypes, tumor cells and acinar cells in classical mPDAC. The ligand protein expression fold change, identified from secretome experiments, between T/N and control is shown in the middle. Normalized receptor expression levels obtained from scRNA-seq data are shown in the outer concentric circles. c, UMAP plot highlighting the whole population of CAF cells identified in classical and mesenchymal tumors. d, Left, UMAP plot showing the CAF population across different treatment conditions in classical tumors. Right, UMAP plots displaying the identified CAF clusters and resulting subpopulations for classical tumors. e, UMAP plots of the CAF cluster displaying selected marker gene expression. f, Heatmap displaying expression of selected genes in CAFs across the identified clusters. The y axis shows the selected marker genes, the x axis represents each of the identified clusters in (d). g, Violin plot showing Tgfb1 expression by myoCAFs across the different treatment conditions. h, Proportion of CAF subtypes in the indicated different treatment conditions. CAF subpopulations were identified in the fibroblast cell clusters and annotated with the markers described in (f). T/N: trametinib+nintedanib, T/N + aPD-L1: trametinib+nintedanib+anti PD-L1 antibody.

Source data

Supplementary information

Supplementary Information

Supplementary Fig. 1 FACS gating strategy.

Reporting Summary

Supplementary Table

Supplementary Tables 1–8.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 2

Unprocessed gels.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 3

Unprocessed gels.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed gels.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Falcomatà, C., Bärthel, S., Widholz, S.A. et al. Selective multi-kinase inhibition sensitizes mesenchymal pancreatic cancer to immune checkpoint blockade by remodeling the tumor microenvironment. Nat Cancer 3, 318–336 (2022). https://doi.org/10.1038/s43018-021-00326-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43018-021-00326-1

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