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:

Cellular adaptation to cancer therapy along a resistance continuum

Abstract

Advancements in precision oncology over the past decades have led to new therapeutic interventions, but the efficacy of such treatments is generally limited by an adaptive process that fosters drug resistance1. In addition to genetic mutations2, recent research has identified a role for non-genetic plasticity in transient drug tolerance3 and the acquisition of stable resistance4,5. However, the dynamics of cell-state transitions that occur in the adaptation to cancer therapies remain unknown and require a systems-level longitudinal framework. Here we demonstrate that resistance develops through trajectories of cell-state transitions accompanied by a progressive increase in cell fitness, which we denote as the ‘resistance continuum’. This cellular adaptation involves a stepwise assembly of gene expression programmes and epigenetically reinforced cell states underpinned by phenotypic plasticity, adaptation to stress and metabolic reprogramming. Our results support the notion that epithelial-to-mesenchymal transition or stemness programmes—often considered a proxy for phenotypic plasticity—enable adaptation, rather than a full resistance mechanism. Through systematic genetic perturbations, we identify the acquisition of metabolic dependencies, exposing vulnerabilities that can potentially be exploited therapeutically. The concept of the resistance continuum highlights the dynamic nature of cellular adaptation and calls for complementary therapies directed at the mechanisms underlying adaptive cell-state transitions.

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: Progressive cell-state transitions along a resistance continuum.
Fig. 2: Assembly of gene module expression accompanies drug adaptation.
Fig. 3: CNAs and chromatin reprogramming during drug adaptation.
Fig. 4: In vivo models recapitulate persister and adapted states.
Fig. 5: CRISPR-based negative-selection screen reveals systems-level metabolic dependencies.

Similar content being viewed by others

Data availability

Raw and processed data from this manuscript have been deposited in the NCBI Genome Expression Omnibus (GEO) under the accession number GSE247691. Metabolomics data have been deposited in the Metabolomics Workbench (study ID: ST003102; https://doi.org/10.21228/M8WM7T).

Code availability

Relevant code used in this study is available at https://github.com/yanailab/drug-resistance.

References

  1. Vasan, N., Baselga, J. & Hyman, D. M. A view on drug resistance in cancer. Nature 575, 299–309 (2019).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  2. McGranahan, N. & Swanton, C. Clonal heterogeneity and tumor evolution: past, present, and the future. Cell 168, 613–628 (2017).

    CAS  PubMed  Google Scholar 

  3. Sharma, S. V. et al. A chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations. Cell 141, 69–80 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Hugo, W. et al. Non-genomic and immune evolution of melanoma acquiring MAPKi resistance. Cell 162, 1271–1285 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Marine, J.-C., Dawson, S.-J. & Dawson, M. A. Non-genetic mechanisms of therapeutic resistance in cancer. Nat. Rev. Cancer 20, 743–756 (2020).

    CAS  PubMed  Google Scholar 

  6. Shaffer, S. M. et al. Rare cell variability and drug-induced reprogramming as a mode of cancer drug resistance. Nature 546, 431–435 (2017).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  7. Oren, Y. et al. Cycling cancer persister cells arise from lineages with distinct programs. Nature 596, 576–582 (2021).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  8. Brock, A., Chang, H. & Huang, S. Non-genetic heterogeneity–a mutation-independent driving force for the somatic evolution of tumours. Nat. Rev. Genet. 10, 336–342 (2009).

    CAS  PubMed  Google Scholar 

  9. Pisco, A. O. et al. Non-Darwinian dynamics in therapy-induced cancer drug resistance. Nat. Commun. 4, 2467 (2013).

    ADS  PubMed  Google Scholar 

  10. Tsoi, J. et al. Multi-stage differentiation defines melanoma subtypes with differential vulnerability to drug-induced iron-dependent oxidative stress. Cancer Cell 33, 890–904.e5 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Vander Velde, R. et al. Resistance to targeted therapies as a multifactorial, gradual adaptation to inhibitor specific selective pressures. Nat. Commun. 11, 2393 (2020).

    ADS  Google Scholar 

  12. Domcke, S., Sinha, R., Levine, D. A., Sander, C. & Schultz, N. Evaluating cell lines as tumour models by comparison of genomic profiles. Nat. Commun. 4, 2126 (2013).

    ADS  PubMed  Google Scholar 

  13. Lord, C. J. & Ashworth, A. PARP inhibitors: synthetic lethality in the clinic. Science 355, 1152–1158 (2017).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Goyal, Y. et al. Diverse clonal fates emerge upon drug treatment of homogeneous cancer cells. Nature 620, 651–659 (2023).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. Shen, S., Vagner, S. & Robert, C. Persistent cancer cells: the deadly survivors. Cell 183, 860–874 (2020).

    CAS  PubMed  Google Scholar 

  16. Lin, L. et al. SOX17 and PAX8 constitute an actionable lineage-survival transcriptional complex in ovarian cancer. Oncogene 41, 1767–1779 (2022).

    CAS  PubMed  Google Scholar 

  17. Kim, C., Wang, X.-D. & Yu, Y. PARP1 inhibitors trigger innate immunity via PARP1 trapping-induced DNA damage response. eLife 9, e60637 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Kreß, J. K. C. et al. The integrated stress response effector ATF4 is an obligatory metabolic activator of NRF2. Cell Rep. 42, 112724 (2023).

    PubMed  Google Scholar 

  19. Shibue, T. & Weinberg, R. A. EMT, CSCs, and drug resistance: the mechanistic link and clinical implications. Nat. Rev. Clin. Oncol. 14, 611–629 (2017).

    PubMed  PubMed Central  Google Scholar 

  20. Guo, W. et al. Slug and Sox9 cooperatively determine the mammary stem cell state. Cell 148, 1015–1028 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Arlt, M. F., Wilson, T. E. & Glover, T. W. Replication stress and mechanisms of CNV formation. Curr. Opin. Genet. Dev. 22, 204–210 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Quintanal-Villalonga, Á. et al. Lineage plasticity in cancer: a shared pathway of therapeutic resistance. Nat. Rev. Clin. Oncol. 17, 360–371 (2020).

    PubMed  PubMed Central  Google Scholar 

  23. Schmitt, C. A. et al. A senescence program controlled by p53 and p16INK4a contributes to the outcome of cancer therapy. Cell 109, 335–346 (2002).

    CAS  PubMed  Google Scholar 

  24. Birsoy, K. et al. An essential role of the mitochondrial electron transport chain in cell proliferation is to enable aspartate synthesis. Cell 162, 540–551 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Sayin, V. I. et al. Activation of the NRF2 antioxidant program generates an imbalance in central carbon metabolism in cancer. eLife 6, e28083 (2017).

    PubMed  PubMed Central  Google Scholar 

  26. Shen, Y.-A. et al. Inhibition of the MYC-regulated glutaminase metabolic axis is an effective synthetic lethal approach for treating chemoresistant ovarian cancers. Cancer Res. 80, 4514–4526 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Debaugnies, M. et al. RHOJ controls EMT-associated resistance to chemotherapy. Nature 616, 168–175 (2023).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Seo, J. et al. AP-1 subunits converge promiscuously at enhancers to potentiate transcription. Genome Res. 31, 538–550 (2021).

    PubMed  PubMed Central  Google Scholar 

  29. Martínez-Zamudio, R. I. et al. AP-1 imprints a reversible transcriptional programme of senescent cells. Nat. Cell Biol. 22, 842–855 (2020).

    PubMed  PubMed Central  Google Scholar 

  30. Larsen, S. B. et al. Establishment, maintenance, and recall of inflammatory memory. Cell Stem Cell 28, 1758–1774.e8 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Freddolino, P. L., Yang, J., Momen-Roknabadi, A. & Tavazoie, S. Stochastic tuning of gene expression enables cellular adaptation in the absence of pre-existing regulatory circuitry. eLife 7, e31867 (2018).

    PubMed  PubMed Central  Google Scholar 

  32. Gallaher, J. A., Enriquez-Navas, P. M., Luddy, K. A., Gatenby, R. A. & Anderson, A. R. A. Spatial heterogeneity and evolutionary dynamics modulate time to recurrence in continuous and adaptive cancer therapies. Cancer Res. 78, 2127–2139 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Rehman, S. K. et al. Colorectal cancer cells enter a diapause-like DTP state to survive chemotherapy. Cell 184, 226–242.e21 (2021).

    CAS  PubMed  Google Scholar 

  34. Marsolier, J. et al. H3K27me3 conditions chemotolerance in triple-negative breast cancer. Nat. Genet. 54, 459–468 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Hata, A. N. et al. Tumor cells can follow distinct evolutionary paths to become resistant to epidermal growth factor receptor inhibition. Nat. Med. 22, 262–269 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Blount, Z. D., Lenski, R. E. & Losos, J. B. Contingency and determinism in evolution: replaying life’s tape. Science 362, eaam5979 (2018).

    PubMed  Google Scholar 

  37. Huang, S. Reconciling non-genetic plasticity with somatic evolution in cancer. Trends Cancer Res. 7, 309–322 (2021).

    CAS  Google Scholar 

  38. Zhang, K. et al. Longitudinal single-cell RNA-seq analysis reveals stress-promoted chemoresistance in metastatic ovarian cancer. Sci. Adv. 8, eabm1831 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Rukhlenko, O. S. et al. Control of cell state transitions. Nature 609, 975–985 (2022).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Boumahdi, S. & de Sauvage, F. J. The great escape: tumour cell plasticity in resistance to targeted therapy. Nat. Rev. Drug Discov. 19, 39–56 (2020).

    CAS  PubMed  Google Scholar 

  41. Zhao, W. et al. A new bliss independence model to analyze drug combination data. J. Biomol. Screen. 19, 817–821 (2014).

    PubMed  Google Scholar 

  42. Hangauer, M. J. et al. Drug-tolerant persister cancer cells are vulnerable to GPX4 inhibition. Nature 551, 247–250 (2017).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  43. Cybulska, P. et al. A genomically characterized collection of high-grade serous ovarian cancer xenografts for preclinical testing. Am. J. Pathol. 188, 1120–1131 (2018).

    CAS  PubMed  Google Scholar 

  44. Shen, Y. et al. BMN 673, a novel and highly potent PARP1/2 inhibitor for the treatment of human cancers with DNA repair deficiency. Clin. Cancer Res. 19, 5003–5015 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Klein, A. M. et al. Droplet barcoding for single-cell transcriptomics applied to embryonic stem cells. Cell 161, 1187–1201 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Baron, M. et al. The stress-like cancer cell state is a consistent component of tumorigenesis. Cell Syst. 11, 536–546.e7 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Stoeckius, M. et al. Cell hashing with barcoded antibodies enables multiplexing and doublet detection for single cell genomics. Genome Biol. 19, 224 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Corces, M. R. et al. An improved ATAC-seq protocol reduces background and enables interrogation of frozen tissues. Nat. Methods 14, 959–962 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Gong, W., Kwak, I.-Y., Pota, P., Koyano-Nakagawa, N. & Garry, D. J. DrImpute: imputing dropout events in single cell RNA sequencing data. BMC Bioinformatics 19, 220 (2018).

    PubMed  PubMed Central  Google Scholar 

  51. Haghverdi, L., Lun, A. T. L., Morgan, M. D. & Marioni, J. C. Batch effects in single-cell RNA-sequencing data are corrected by matching mutual nearest neighbors. Nat. Biotechnol. 36, 421–427 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Neftel, C. et al. An integrative model of cellular states, plasticity, and genetics for glioblastoma. Cell 178, 835–849.e21 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Gel, B. et al. regioneR: an R/Bioconductor package for the association analysis of genomic regions based on permutation tests. Bioinformatics 32, 289–291 (2016).

    CAS  PubMed  Google Scholar 

  54. Kumar, L. & E Futschik, M. Mfuzz: a software package for soft clustering of microarray data. Bioinformation 2, 5–7 (2007).

    PubMed  PubMed Central  Google Scholar 

  55. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  57. Zhang, Y. et al. Model-based analysis of ChIP-seq (MACS). Genome Biol. 9, R137 (2008).

    PubMed  PubMed Central  Google Scholar 

  58. Amemiya, H. M., Kundaje, A. & Boyle, A. P. The ENCODE Blacklist: Identification of problematic regions of the genome. Sci. Rep. 9, 9354 (2019).

    ADS  PubMed  PubMed Central  Google Scholar 

  59. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    PubMed  PubMed Central  Google Scholar 

  61. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    PubMed  PubMed Central  Google Scholar 

  62. Levin, M. et al. The mid-developmental transition and the evolution of animal body plans. Nature 531, 637–641 (2016).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  63. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Tarasov, A., Vilella, A. J., Cuppen, E., Nijman, I. J. & Prins, P. Sambamba: fast processing of NGS alignment formats. Bioinformatics 31, 2032–2034 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. McKenna, A. et al. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20, 1297–1303 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Boeva, V. et al. Control-FREEC: a tool for assessing copy number and allelic content using next-generation sequencing data. Bioinformatics 28, 423–425 (2012).

    CAS  PubMed  Google Scholar 

  68. Wang, K., Li, M. & Hakonarson, H. ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 38, e164 (2010).

    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).

    PubMed  PubMed Central  Google Scholar 

  70. Pacold, M. E. et al. A PHGDH inhibitor reveals coordination of serine synthesis and one-carbon unit fate. Nat. Chem. Biol. 12, 452–458 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Simón-Manso, Y. et al. Metabolite profiling of a NIST Standard Reference Material for human plasma (SRM 1950): GC–MS, LC–MS, NMR, and clinical laboratory analyses, libraries, and web-based resources. Anal. Chem. 85, 11725–11731 (2013).

    PubMed  Google Scholar 

  72. Chen, W. W., Freinkman, E., Wang, T., Birsoy, K. & Sabatini, D. M. Absolute quantification of matrix metabolites reveals the dynamics of mitochondrial metabolism. Cell 166, 1324–1337.e11 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Smith, C. A. et al. METLIN: a metabolite mass spectral database. Ther. Drug Monit. 27, 747–751 (2005).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank R. White, S. Frank, M. Schober, M. D. Vibranovski, A. Rozanski and the members of the Yanai laboratory for helpful discussions; the laboratories of B. Neel, K.-K. Wong and A. Lund for providing cell lines; J. A. Ferry for providing histology annotations; and the laboratory of M. Maurano for assistance with sequencing. We are grateful to NYU Langone Metabolomics Core Resource Laboratory for their assistance with metabolomics experiments and NYU Langone Experimental Pathology Laboratory for their assistance with sectioning, staining and imaging services. This work was supported by the following NIH grants: P50 CA225450, R01 LM013522, R21 CA264361, U01 CA260432 and U54 CA263001 (I.Y.); R01AG075272 (T.L.); 1U54CA274507-01A1 (A.M.); R37CA222504 and R01CA227649 (T.P.); F32CA239394 (B.R.K.); and F30 CA257400 (D.B.). This work was also supported by grants from the Leon Lowenstein Foundation and the Mary Kay Ash Foundation (I.Y.).

Author information

Authors and Affiliations

Authors

Contributions

G.S.F. conceived the study with I.Y., and performed the experiments and all data analyses. M.B. assisted with scRNA-seq experiments. B.R.K., S.M., D.A. Levine and T.L. designed the in vivo resistance experiments and B.R.K. performed scRNA-seq for these experiments. B.R.K., S.M., G.S.F., T.L. and I.Y. designed the persistence in vivo experiments. B.R.K., G.S.F. and M.P. performed scRNA-seq for these experiments with assistance from D.B. S.M. performed all of the mouse work. M.P. performed and assisted with in vitro experiments. G.S.F., J.P.B., T.P. and I.Y. designed the CRISPR screen experiments. G.S.F. and J.P.B. performed the CRISPR screen experiments under the supervision of T.P. and I.Y. G.S.F., A.B., A.M. and I.Y. designed the SOX17 and Slug functional experiments. A.B. performed these experiments under the supervision of A.M. G.S.F. and A.R. designed and performed the ATAC-seq, scRNA-seq and WES experiments. I.D. contributed with WES data processing and analysis. G.S.F. and A.S.P. performed additional scRNA-seq and ATAC-seq experiments. K.H.T., M.C. and A.S.P. assisted with in vitro experiments. D.A. Liberman assisted with data analysis and staining. G.A. and F.K. assisted with scRNA-seq. G.S.F and I.Y. wrote the manuscript with inputs from A.M. I.Y. oversaw the project. All authors commented on the manuscript.

Corresponding author

Correspondence to Itai Yanai.

Ethics declarations

Competing interests

D.A. Levine is a founder of Resident Diagnostics and is a full-time employee of Merck & Co. T.P. received funding from Dracen Pharmaceuticals, Kymera Therapeutics, Bristol Myers Squibb and Agios Pharmaceuticals not related to the submitted work. T.P. consulted for Vividion Therapeutics, Tohoku University and Faeth Therapeutics.

Peer review

Peer review information

Nature thanks Isaac Harris and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer review 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 figures and tables

Extended Data Fig. 1 Generation of drug-adapted populations in additional resistance models.

a, Drug-adapted lines generated for Ovsaho, a BRCA-mutant ovarian cancer cell line, using dose-escalation from 1 to 40 μM of olaparib. b, Cell viability assays (9 days of treatment) for the Ovsaho adapted lines showing progressive increase in resistance. Bar plots show the average of 6 technical replicates with standard deviation (SD) error bars. c, Same as in a for COV362, also a BRCA-mutant ovarian cancer cell line. d, Same as in b for COV362. e, Drug-adapted lines generated for A375, a melanoma cell line, using dose-escalation from 30 nM to 4 μM of dabrafenib. f, Cell viability assays (4 days of treatment) for the A375 adapted lines. g, Drug-adapted lines generated for PC9, a lung adenocarcinoma cell line, using dose-escalation from 8 nM to 1.2 μM of osimertinib. h, Cell viability assays (5 days of treatment) for the PC9 adapted lines. Cell viability assays were performed using CellTiter Glo 2.0 and values are normalized by DMSO control. Data are shown as means and SDs of 6 technical replicates.

Extended Data Fig. 2 Drug-adaptation in acute versus gradual dose exposure and treatment duration.

a, Generation of drug-tolerant persisters for Kuramochi cell line by treatment with 320 μM of olaparib for 9 days. The surviving cells were allowed to recover for 9 days without treatment. Two rounds of treatment-recovery experiments were performed in triplicates and the viability of the recovering cells was tested. Cell viability curves are shown as the average of three replicates (n = 6 technical replicates) and error bars as SD. The bar plot shows no significant IC50 increase for persisters relative to the parental population (C vs P320.1: P = 0.03; ns, not significant; two-tailed t-tests). b–d, Extended treatments at constant doses. Representative Kuramochi adapted lines (T5, T20 and T80) were maintained on olaparib treatment at the final concentrations to which they had adapted (5, 20 and 80 μM, respectively) for the number of days that they had taken to adapt to their subsequent populations at increased concentrations (T10, T40, T160, respectively) (see Fig. 1a), generating the extended T5-E, T10-E and T80-E populations (three independent replicates). Prolonged exposure at constant doses did not generate similarly resistant populations for the T5 and T20 extended treatments when compared to their dose-escalated counterparts (T10 and T40, respectively). The extended treatments did however lead to higher resistance relative to their parental populations. The cells from the T80 extended treatments were similarly resistant to T160, highlighting the role of treatment duration. Together, these observations indicate that both dose and time influence the fitness outcomes along the adaptive path. Cell viability was measured by CellTiter Glo 2.0 (each biological replicate with 6 technical replicates) and by cell counting (bar plots show the average of three biological replicates and error bars as SD) after 9 days of treatment at their respective escalated doses (10, 40 and 160 μM). IC50 (left bar plots) and relative cell counts (right bar plots) were compared with their parental and dose-escalated populations (significant P-values are shown; ns, not significant; two-tailed t-tests). Dose response curves for the original T5, T10, T20, T40, T80 and T160 are from Fig. 1b as performed in 3 independent experiments. e–h, To test whether dose escalation facilitates drug adaptation in our model, we compared the resistance phenotypes between cells that were dose-escalated from 1 to 10 μM with drug-naive cells directly exposed to the same doses for the same duration. Dose-escalated cells were able to proliferate and reach confluency, contrasting with drug-naive cells directly exposed to higher doses (5 and 10 μM). The continuous long-term 10 μM treatment led to very few viable cells, which prevented us from testing their fitness with viability assays (h). The populations that were dose-escalated up to 2.5 μM were significantly more resistant than their continuously treated counterparts. Plots indicate dose-response curves for three biological replicates (R1, R2, R3; 6 technical replicates each) and bar plots of average IC50 values (3 biological replicates with SD error bars). Bar plots (right) showing the average cell counts in f-h for the 3 biological replicates with SD error bars. Cell viability assays and cell counts (treated with 2.5, 5 and 10��μM of olaparib, respectively) after the treatment periods are shown and compared with the dose-escalated populations (P-values are indicated; ns, not significant; two-tailed t-tests). Microscopic photos show that T10 (dose-escalated) cells reached confluency while T10.C (continuous) did not (scale bar 50 μm).

Extended Data Fig. 3 scRNA-Seq analysis for two additional ovarian cancer adapted cell lines.

a, Spearman’s correlations across the averaged transcriptomes of the Ovsaho olaparib-adapted lines. Untreated control (C) and T5 to T40 were collected for scRNA-Seq. b-c, UMAP of Ovsaho scRNA-seq data on the adapted lines colored by treatment (b) and by cluster as determined by Louvain clustering (c), highlighting the emergence of distinct cellular states. d, Bar plot representing the percentage of cells from each condition in each cluster, indicating cell state frequency changes across the adapted lines. e, Differentially expressed genes across Ovsaho clusters, highlighting the expression of shared genes with Kuramochi (Fig. 2a), COV362 (Extended Data Fig. 3j) and PDX resistance models (Fig. 4f). Color bar represents the scaled average expression for a particular cluster and circle sizes are the fraction of cells expressing the gene. Overall, genes highlight dedifferentiation and metabolic reprogramming along the resistance continuum. f, Same as in a for COV362. g-h, Same as in b and c for COV362, respectively. i-j, Same as in d and e for COV362, respectively. k, Drug-tolerant persisters generation by treatment of drug-naive cells with 10 and 320 μM of olaparib for 9 days. The adapted T10 and T320 lines were maintained at the same doses, respectively (left, scale bar 50 μm, representative of two biological replicates). UMAP of the scRNA-Seq data for the persister and adapted cell populations (right). Phenotypic features of persisters are shown in Supplementary Fig. 2. l, Differentially expressed genes across conditions (relative to C) highlighting shared and distinct marker gene expression among the persister and adapted lines.

Extended Data Fig. 4 scRNA-Seq analysis for the melanoma and lung adenocarcinoma adapted lines.

a, Spearman’s correlations across the averaged transcriptomes of the A375 (melanoma) adapted lines that evolved resistance through dose-escalation from 30 nM to 4 μM of dabrafenib (Extended Data Fig. 1e,f). b, UMAP of scRNA-seq data on the individual lines. Colors and numbers indicate subpopulations determined by Louvain clustering. c-d, UMAP for all cells colored by treatment (c) and by cluster as determined by Louvain clustering (d), highlighting distinct transcriptional states across the adapted lines. e, Bar plot representing the percentage of cells from each condition in each cluster, indicating the emergence of cell states and frequency changes that suggest state selection along the resistance continuum. f, Differentially expressed genes across A375 clusters, highlighting the expression of lineage plasticity and metabolic reprogramming markers across the resistance continuum. Differentiation markers indicate multiple intermediate states, showing a progression from SOX10high (clusters 0-3) to SOX10low expression (clusters 4-7) and loss of MHC class II genes (CD74, HLDA-DRA/B) (Supplementary Table 4). Diverse lineage states arise along the resistance path, such as cells expressing both melanocytic (DCT, MITF) and neural-crest (NGFR) lineage markers (cluster 3); a population expressing highly MITF, RAP1B and TMSB4X (cluster 4); cells expressing neuronal markers (ROBO1, BMP5, SYT4) (cluster 5); and finally a mesenchymal-like invasive state (SOX9, FN1) (clusters 6-7, Supplementary Table 4). A gradual increase in the expression of ATF4 and NRF2 regulated genes toward the most resistant states (clusters 4 to 7), indicating extensive metabolic rewiring. Color bar represents the scaled average expression for a particular cluster and circle sizes are the fraction of cells expressing the gene. g-h, Same as in a and b for PC9 (lung adenocarcinoma), respectively. Adapted lines evolved resistance through dose-escalation from 8 nM to 1.2 μM of osimertinib (Extended Data Fig. 1g,h). A persister population (P1.2) was generated by treatment with 1.2 μM of osimertinib for 9 days. i-l, Same as in c-f for PC9, respectively, including a persister population (P1.2, i). PC9 cells transitioned to a KRT6Alow low expression and coexist in a combination of CD44highCD24low and CD44lowCD24high states (l) (Supplementary Table 5).

Extended Data Fig. 5 Expression of CD24 and CD44 as a proxy for adapted states in Kuramochi cells.

a, Scatter plot showing the average expression of CD24 and CD44 markers in each of the identified subpopulations colored by treatment condition (left) and state (right). b, Representative plots of the flow cytometry analysis of CD24 and CD44 staining across the Kuramochi adapted populations. Bottom left quadrants represent double negative staining for both markers. A transition from CD24highCD44low (top right) to CD24lowCD44high (bottom right quadrants) is observed. Numbers indicate the percentage of cells in each quadrant. Colored squares indicate the subpopulations sorted for the cell viability assays shown in (c), ATAC-Seq (Fig. 3) and CRISPR-based screen (Fig. 5). c, Cell viability assays of sorted subpopulations indicating differential resistance within (T10) and between (T10, T40, T320) treatment conditions. Sorted subpopulations were labeled according to their CD24/CD44 profiles. Bar plots show the IC50 averages of two independent experiments (6 technical replicates each).

Extended Data Fig. 6 EMT signatures and perturbation of lineage plasticity in drug resistance.

a, Western blot of the induced SOX17 in the T320 line compared to the parental control (C) cells (three technical replicates). Bar plot shows the quantification signal. Full western blot shown in Supplementary Fig. 6. b, Schematic of the constitutive SOX17 overexpression construct (top). Quantification by live cell imaging of the percent area covered by cells over a 20 day time period of T320 parental and T320 with SOX17 overexpression on DMSO (left) and on olaparib (5 μM) treatment (right). Average and SD of three biological replicates (three technical replicates each) shown. c, Western blot of SOX17 shRNA knockdown in parental control (C) cells compared to non-targeting scrambled shRNA (three technical replicates). Bar plot shows the quantification signal. Full western blot shown in Supplementary Fig. 7. d, Quantification of SOX17 shRNA knockdown in parental control (C) cells and adapted T5, T40 and T320 by live cell imaging (three biological replicates, three technical replicates each) of the percent area covered by cells normalized by DMSO controls on day five of treatment of 2.5 μM olaparib. Standard two-sided t-test was performed by comparing with the negative control cells (Neg. Ctrl) (SOX17 #1: P = 0.01; SOX17 #2: P = 0.02; SOX17 #3: P = 0.05; T40: P = 0.004; T320: P = 0.008; ns: not significant). e-i, EMT signature scores were calculated across treatment conditions for each resistance model. The set of genes annotated as “Hallmark epithelial-mesenchymal transition” from the MSigDB was used to calculate the scores with the UCell algorithm. Violin plots represent the score distributions with the average plus and minus SD. Numbers of cells are: (e) C (N = 453), T1 (N = 1223), T2.5 (N = 760), T5 (N = 398), T5 (N = 398), T10 (N = 1030), T20 (N = 640), T40 (N = 918), T80 (N = 910), T160 (N = 1348), T320 (N = 568); (f) C (N = 840), T5 (N = 713), T10 (N = 704), T20 (N = 731), T40 (N = 650); (g) C (N = 1148), T5 (N = 180), T10 (N = 278), T20 (N = 447), T40 (N = 272); (h) C (N = 442), T003 (N = 478), T006 (N = 906), T0125 (N = 306), T0250 (N = 733), T05 (N = 493), T1 (N = 605), T2 (N = 641), T4 (N = 455); (i) C (N = 169), P1.2 (N = 673), T0008 (N = 329), T0016 (N = 528), T0032 (N = 414), T0075 (N = 505), T0150 (N = 329), T0300 (N = 442), T0600 (N = 424), T1.2 (N = 395). j, Representative images (biological duplicates) of stained membranes from a transwell cell invasion assay. Colors represent the untreated cells (C), adapted lines T5 and T320 and the Slug-lines on different doxycycline (dox) concentrations (C-Slug0 as negative and C-Slug1 on 1 μg/ml of dox). Quantification of transwell cell invasion assay data of three images per cell line (right). One way ANOVA comparing Slug-induced (SNAI2) lines to the drug-naive (C) cell line. k, Quantification of induced SNAI2 gene expression from qRT-PCR (two independent runs in triplicates). One way ANOVA comparing Slug-induced lines to the drug-naive (C) cell line. l, Representative western blot (three technical replicates) of the induced Slug lines with different doxycycline concentrations ranging from 0.01 μg/ml to 1 μg/ml. We observed that C-Slug1 produced a relevant phenotype, which was used for further experiments. Bar plot shows the quantification signal. Full western blot shown in Supplementary Fig. 8. m, Quantification by live cell imaging of DMSO controls For Slug-induced lines. Data represents percent area covered by cells over a 54 day time period. Average and SD of three biological replicates (three technical replicates each) are shown.

Extended Data Fig. 7 Copy number inference from scRNA-seq data for the adapted and persister lines.

a, Heat maps depicting scRNA-Seq inferred copy number alterations (CNAs) for Kuramochi adapted and persister lines relative to untreated control (C). Gene expression signal is scaled along the chromosomes as calculated by inferCNV. Red represents genomic regions with inferred copy number gain and blue represents regions with inferred copy number loss. Colors and labels on the left represent the subpopulations identified by transcriptome clustering (Fig. 1d) showing correspondence between inferred CNAs and transcriptional states. b, Enrichment of genes within CNA regions identified by whole-exome sequencing overlapping with the genes present in the transcriptional modules detected for the adapted lines (Fig. 2a) in respect to random distributions (P-values determined by 10,000 permutations, two-sided permutation tests). Blue and red lines indicate the observed overlap of copy number loss and gain, respectively, and the gray lines indicate the significance threshold in an expected random distribution. c, Inferred CNAs defined for the adapted Ovsaho and COV362 lines relative to their untreated controls. Color annotations represent sample and transcriptional clusters identified in Extended Data Fig. 3c,h.

Extended Data Fig. 8 Chromatin accessibility and stability of drug resistance.

a, Correlation between the number of differentially accessible peaks and total number of detected peaks per sample. The lack of a positive correlation indicates that the number of differentially accessible peaks are not explained by peak detection bias (Spearman’s rho = −0.62, P = 0.07, two-sided Spearman Rank Correlation test). b, Heat maps showing the hierarchical clustering (rows) of the enriched TF motifs in the differentially more (green) and less (purple) accessible peaks (relative to C). Color bar represents the scaled p-values for TF motif enrichment (only transcription factors with P-value < 1e-5 were included, one-sided hypergeometric test, Benjamini-Hochberg corrected). The analysis controlled by peaks not overlapping CNAs (“no-CNA”) reveals the same patterns when compared to all peaks. c, Enrichment of differentially accessible ATAC-Seq peaks overlapping CNA regions (relative to the untreated control C) identified by whole-exome sequencing for representative adaptive lines (T10, T40 and T320) (P-values determined by 1,000 permutations, two-sided permutation tests). Blue and red lines indicate the observed overlap of copy number loss and gain, respectively, and the gray lines indicate the significance threshold from the expected random distribution. The proportion of peaks not overlapping CNAs ranged from 70% for less, and 80–90% for more accessible peaks, highlighting widespread global chromatin changes. d, Transcription factor (TF) motif enrichment analysis controlling for peaks not overlapping with CNAs, showing similar frequency of peaks containing TFs motifs involved with regulation of stress response (AP1, NRF2 and ATF4) and ovarian cancer TFs (SOX17, PAX8 and WT1). Bar plots represent the percentage of differentially accessible peaks for each condition (relative to C). Color bars represent the motif enrichment P-values (one-sided hypergeometric test, Benjamini-Hochberg corrected). e, Proliferation of C, T320 and T320W (which is the T320 population after drug removal for 120 days) on DMSO, 10, 40 and 320 μM of olaparib for 8 days. The T320W exhibits partial reversibility in resistance, especially at 320 μM. Proliferation was measured every 2 days using Cell TiterGlo 2.0 assay and values are normalized from day 0 (24 h after cell seeding). Data represents the average (dots) of 6 technical replicates and error bars indicate the SD. f, Box plots of the signal for less (purple) (N = 6771) and more (green) (N = 7332) accessible peaks shown in Fig. 3f. This shows that the signal across the less accessible regions (relative to C, Fig. 3d) both with and without treatment are overall stable when comparing T320 and T320W. For the more accessible regions, the signal decreases in T320W without treatment and recovers its accessibility during drug treatment. P-values were determined from two-sided Wilcoxon tests and the effect sizes are Cohen’s d. Boxes represent the interquartile range (IQR) and centre lines represent the median. Whiskers extend to the smallest and largest values within 1.5 times the IQR from the lower and upper quartiles, respectively. Data points beyond the whiskers are outliers.

Extended Data Fig. 9 Persistence and resistance model of high-grade serous ovarian cancer patient-derived xenografts.

a, Tumor growth curves for PDXs treated with four doses of talazoparib (shades of green) and vehicle controls (gray) in the persistence experiment. Each condition was collected in duplicates for scRNA-Seq. Two replicates of each condition were generated (indicated by “.1” and “.2”). b-d, Tumor growth curves for PDXs in first (b), second (c) and third (d) rounds of talazoparib treatment in the resistance experiment. A drug holiday in the first round is indicated in red. Two replicates for each condition were collected for scRNA-Seq: vehicles (gray), round 1 (solid pink) and round 3 (blue, with an additional vehicle treated R3.V). A replicate tumor from round 1 (dashed pink) was seeded in a new cohort to generate the round 2 treatment. e, Hematoxylin and eosin (H + E) staining of the entire section of V1, D3.1 (persistence) and R3.2 (resistance) (column 1) and an enlargement (column 2). Histology was annotated by a gynecologic pathologist. The V1 tumor has the classic appearance of serous carcinoma with nests of neoplastic cells with “slit-like spaces” separating strands of tumor cells. The D3.1 tumor still is recognizable as serous carcinoma while more poorly differentiated than V1. The R3.2 tumor has a biphasic pattern with some areas that retain a recognizable papillary pattern, while it is dominated by poorly differentiated/undifferentiated tumor structures with discohesive tumor cells. DAPI staining (blue), CD44 (green) and SOX17 (red) as single channels (columns 3-5), merged (column 6) and an enlargement (column 7). Consistent with scRNA-Seq data, the well-differentiated regions in V1 highly express SOX17 and lack expression of CD44, which is also observed in D3.1. Regions of similar histology in R3.2 have weaker expression of SOX17 and remain to lack expression of CD44 (columns 4-5). In contrast, the poorly differentiated regions in R3.2 strongly express CD44 and weakly express SOX17 (columns 6-7), thus indicating heterogeneous populations in different resistant states. Scale bars represent 100 μm except in column 1 (1000 μm). f, Heat map of the scaled gene average expression across samples of the dose-response experiment (persistence), highlighting differentially expressed genes that show a dose-dependent trend. g, Inferred CNAs defined for the persistence and resistance PDX experiments relative to their respective vehicle treated tumors. Color annotations represent sample and transcriptional clusters identified in Fig. 4b,d.

Extended Data Fig. 10 Metabolic vulnerabilities along the resistance continuum.

a, Rank plots showing the score differences for CD, T10, T320 and T320D. Higher absolute score differences indicate increased gene depletion relative to C. Highlighted genes are significantly depleted (FDR < 0.05, two-sided likelihood ratio test (LTR)) at the endpoint. Only genes with negative score differences are shown, indicating gene depletion. b, Overlap between the significantly depleted genes with high absolute score differences (>1 SD) relative to C. P-values were determined by one-sided hypergeometric tests. c, Heat maps of the differentially abundant metabolites (P < 0.05, two-sided t-tests) comparing olaparib sensitive (C) and resistant (T320) populations. Conditions represent cells on olaparib treatment (320 μM for 24 h) or cells without treatment. The experiment was performed in triplicates for each condition. Color scale represents scaled normalized metabolite abundance. Metabolites involved in nucleotide biosynthesis are highlighted in red. d, Dose-response curves of olaparib treatment or in combination with CB-839 after 6 days for representative populations within the resistance continuum (C, T5, T10, T40 and T320). Data points are normalized relative to vehicle-treated controls (for each respective line) and represent the average of 3 independent experiments (6 technical replicates per experiment) and SD. e, Bar plots showing the IC50 determined from dose-response curves (d) after 6 days of treatment with olaparib, or the combination of olaparib with CB-839. Dots represent three independent experiments (average of 6 technical replicates). Cells were sorted based on CD24low/CD44high profiles (as in Extended Data Fig. 5b,c). f, Bar plot showing the average cell viability of cells treated with CB-839 (50 nM) normalized to vehicle controls (3 independent experiments, 6 technical replicates each). g, Heat map of the bliss synergy scores41 of the olaparib and CB-839 treatment combination. h, Representative western blot (biological duplicates) showing the expression of NRF2 and its target HO1 (HMOX1) with and without olaparib treatment. Full western blot is shown in Supplementary Fig. 10.

Supplementary information

Supplementary Information

This file contains Supplementary Figures 1–10, the Supplementary Discussion and Supplementary References.

Reporting Summary

Supplementary Tables

This file contains Supplementary Tables 1-16.

Peer Review File

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

França, G.S., Baron, M., King, B.R. et al. Cellular adaptation to cancer therapy along a resistance continuum. Nature 631, 876–883 (2024). https://doi.org/10.1038/s41586-024-07690-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-024-07690-9

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing