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:

Correlating chemical sensitivity and basal gene expression reveals mechanism of action

Abstract

Changes in cellular gene expression in response to small-molecule or genetic perturbations have yielded signatures that can connect unknown mechanisms of action (MoA) to ones previously established. We hypothesized that differential basal gene expression could be correlated with patterns of small-molecule sensitivity across many cell lines to illuminate the actions of compounds whose MoA are unknown. To test this idea, we correlated the sensitivity patterns of 481 compounds with 19,000 basal transcript levels across 823 different human cancer cell lines and identified selective outlier transcripts. This process yielded many novel mechanistic insights, including the identification of activation mechanisms, cellular transporters and direct protein targets. We found that ML239, originally identified in a phenotypic screen for selective cytotoxicity in breast cancer stem-like cells, most likely acts through activation of fatty acid desaturase 2 (FADS2). These data and analytical tools are available to the research community through the Cancer Therapeutics Response Portal.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Correlating gene expression and CCL sensitivity data illuminates known small-molecule mechanisms of action.
Figure 2: MoA analysis reveals new mechanisms of small-molecule metabolism.
Figure 3: MoA analysis reveals small-molecule transport mechanisms.
Figure 4: MoA analysis identifies and illuminates the basis for a requirement for FADS2 activity in ML239 cytotoxicity.
Figure 5: Large numbers of CCLs are required to identify MoA.

Similar content being viewed by others

References

  1. Hughes, J.P., Rees, S., Kalindjian, S.B. & Philpott, K.L. Principles of early drug discovery. Br. J. Pharmacol. 162, 1239–1249 (2011).

    Article  CAS  Google Scholar 

  2. Schenone, M., Dančík, V., Wagner, B.K. & Clemons, P.A. Target identification and mechanism of action in chemical biology and drug discovery. Nat. Chem. Biol. 9, 232–240 (2013).

    Article  CAS  Google Scholar 

  3. Swinney, D.C. & Anthony, J. How were new medicines discovered? Nat. Rev. Drug Discov. 10, 507–519 (2011).

    Article  CAS  Google Scholar 

  4. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl. Acad. Sci. USA 102, 15545–15550 (2005).

    Article  CAS  Google Scholar 

  5. Hughes, T.R. et al. Functional discovery via a compendium of expression profiles. Cell 102, 109–126 (2000).

    Article  CAS  Google Scholar 

  6. Lamb, J. et al. The Connectivity Map: using gene-expression signatures to connect small molecules, genes, and disease. Science 313, 1929–1935 (2006).

    Article  CAS  Google Scholar 

  7. Barretina, J. et al. The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensitivity. Nature 483, 603–607 (2012).

    Article  CAS  Google Scholar 

  8. Basu, A. et al. An interactive resource to identify cancer genetic and lineage dependencies targeted by small molecules. Cell 154, 1151–1161 (2013).

    Article  CAS  Google Scholar 

  9. Garnett, M.J. et al. Systematic identification of genomic markers of drug sensitivity in cancer cells. Nature 483, 570–575 (2012).

    Article  CAS  Google Scholar 

  10. Guo, W. et al. Formation of 17-allylamino-demethoxygeldanamycin (17-AAG) hydroquinone by NAD(P)H:quinone oxidoreductase 1: role of 17-AAG hydroquinone in heat shock protein 90 inhibition. Cancer Res. 65, 10006–10015 (2005).

    Article  CAS  Google Scholar 

  11. Papillon-Cavanagh, S. et al. Comparison and validation of genomic predictors for anticancer drug sensitivity. J. Am. Med. Inform. Assoc. 20, 597–602 (2013).

    Article  Google Scholar 

  12. Shoemaker, R.H. The NCI60 human tumour cell line anticancer drug screen. Nat. Rev. Cancer 6, 813–823 (2006).

    Article  CAS  Google Scholar 

  13. Haibe-Kains, B. et al. Inconsistency in large pharmacogenomic studies. Nature 504, 389–393 (2013).

    Article  CAS  Google Scholar 

  14. Dančík, V. et al. Connecting small molecules with similar assay performance profiles leads to new biological hypotheses. J. Biomol. Screen. 19, 771–781 (2014).

    Article  Google Scholar 

  15. Adams, D.J. et al. NAMPT is the cellular target of STF-31-like small-molecule probes. ACS Chem. Biol. 9, 2247–2254 (2014).

    Article  CAS  Google Scholar 

  16. Zhai, D., Jin, C., Satterthwait, A.C. & Reed, J.C. Comparison of chemical inhibitors of antiapoptotic Bcl-2-family proteins. Cell Death Differ. 13, 1419–1421 (2006).

    Article  CAS  Google Scholar 

  17. Franceschini, A. et al. STRING v9.1: protein-protein interaction networks, with increased coverage and integration. Nucleic Acids Res. 41, D808–D815 (2013).

    Article  CAS  Google Scholar 

  18. Benetatos, C.A. et al. Birinapant (TL32711), a bivalent SMAC mimetic, targets TRAF2-associated cIAPs, abrogates TNF-induced NF-κB activation, and is active in patient-derived xenograft models. Mol. Cancer Ther. 13, 867–879 (2014).

    Article  CAS  Google Scholar 

  19. Zoppoli, G. et al. Putative DNA/RNA helicase Schlafen-11 (SLFN11) sensitizes cancer cells to DNA-damaging agents. Proc. Natl. Acad. Sci. USA 109, 15030–15035 (2012).

    Article  CAS  Google Scholar 

  20. Marks, K.M. et al. The selectivity of austocystin D arises from cell-line-specific drug activation by cytochrome P450 enzymes. J. Nat. Prod. 74, 567–573 (2011).

    Article  CAS  Google Scholar 

  21. Lafite, P., Dijols, S., Zeldin, D.C., Dansette, P.M. & Mansuy, D. Selective, competitive and mechanism-based inhibitors of human cytochrome P450 2J2. Arch. Biochem. Biophys. 464, 155–168 (2007).

    Article  CAS  Google Scholar 

  22. Javaid, S. et al. Dynamic chromatin modification sustains epithelial-mesenchymal transition following inducible expression of Snail-1. Cell Rep. 5, 1679–1689 (2013).

    Article  CAS  Google Scholar 

  23. Rivera, M.I. et al. Selective toxicity of the tricyclic thiophene NSC 652287 in renal carcinoma cell lines: differential accumulation and metabolism. Biochem. Pharmacol. 57, 1283–1295 (1999).

    Article  CAS  Google Scholar 

  24. Gamage, N. et al. Human sulfotransferases and their role in chemical metabolism. Toxicol. Sci. 90, 5–22 (2006).

    Article  CAS  Google Scholar 

  25. Glaros, T.G. et al. The “survivin suppressants” NSC 80467 and YM155 induce a DNA damage response. Cancer Chemother. Pharmacol. 70, 207–212 (2012).

    Article  CAS  Google Scholar 

  26. Nakahara, T. et al. YM155, a novel small-molecule survivin suppressant, induces regression of established human hormone-refractory prostate tumor xenografts. Cancer Res. 67, 8014–8021 (2007).

    Article  CAS  Google Scholar 

  27. Winter, G.E. et al. The solute carrier SLC35F2 enables YM155-mediated DNA damage toxicity. Nat. Chem. Biol. 10, 768–773 (2014).

    Article  CAS  Google Scholar 

  28. Kamath, A.V., Chong, S., Chang, M. & Marathe, P.H. P-glycoprotein plays a role in the oral absorption of BMS-387032, a potent cyclin-dependent kinase 2 inhibitor, in rats. Cancer Chemother. Pharmacol. 55, 110–116 (2005).

    Article  CAS  Google Scholar 

  29. Young, L.C. et al. Expression of multidrug resistance protein-related genes in lung cancer: correlation with drug response. Clin. Cancer Res. 5, 673–680 (1999).

    CAS  PubMed  Google Scholar 

  30. Adams, D.J. et al. Discovery of small-molecule enhancers of reactive oxygen species that are nontoxic or cause genotype-selective cell death. ACS Chem. Biol. 8, 923–929 (2013).

    Article  CAS  Google Scholar 

  31. Long, J.Z. et al. Selective blockade of 2-arachidonoylglycerol hydrolysis produces cannabinoid behavioral effects. Nat. Chem. Biol. 5, 37–44 (2009).

    Article  CAS  Google Scholar 

  32. Nomura, D.K. et al. Monoacylglycerol lipase regulates a fatty acid network that promotes cancer pathogenesis. Cell 140, 49–61 (2010).

    Article  CAS  Google Scholar 

  33. Germain, A.R. et al. Identification of a selective small molecule inhibitor of breast cancer stem cells. Bioorg. Med. Chem. Lett. 22, 3571–3574 (2012).

    Article  CAS  Google Scholar 

  34. Polyak, K. & Weinberg, R.A. Transitions between epithelial and mesenchymal states: acquisition of malignant and stem cell traits. Nat. Rev. Cancer 9, 265–273 (2009).

    Article  CAS  Google Scholar 

  35. Taube, J.H. et al. Core epithelial-to-mesenchymal transition interactome gene-expression signature is associated with claudin-low and metaplastic breast cancer subtypes. Proc. Natl. Acad. Sci. USA 107, 15449–15454 (2010).

    Article  CAS  Google Scholar 

  36. Sprecher, H. Metabolism of highly unsaturated n-3 and n-6 fatty acids. Biochim. Biophys. Acta 1486, 219–231 (2000).

    Article  CAS  Google Scholar 

  37. Obukowicz, M.G. et al. Identification and characterization of a novel delta6/delta5 fatty acid desaturase inhibitor as a potential anti-inflammatory agent. Biochem. Pharmacol. 55, 1045–1058 (1998).

    Article  CAS  Google Scholar 

  38. Catalá, A. Lipid peroxidation of membrane phospholipids generates hydroxy-alkenals and oxidized phospholipids active in physiological and/or pathological conditions. Chem. Phys. Lipids 157, 1–11 (2009).

    Article  Google Scholar 

  39. Snyder, F., Lee, T. & Wykle, R.L. Ether-linked lipids and their bioactive species. in Biochemistry of Lipids, Lipoproteins, and Membranes edn. 4 (eds. Vance, D.E. & Vance, J.E.) 233–262 (Elsevier, 2002).

  40. Yang, W.S. et al. Regulation of ferroptotic cancer cell death by GPX4. Cell 156, 317–331 (2014).

    Article  CAS  Google Scholar 

  41. Park, W.J., Kothapalli, K.S., Lawrence, P. & Brenna, J.T. FADS2 function loss at the cancer hotspot 11q13 locus diverts lipid signaling precursor synthesis to unusual eicosanoid fatty acids. PLoS One 6, e28186 (2011).

    Article  CAS  Google Scholar 

  42. Salt, M.B., Bandyopadhyay, S. & McCormick, F. Epithelial-to-mesenchymal transition rewires the molecular path to PI3K-dependent proliferation. Cancer Discov. 4, 186–199 (2014).

    Article  CAS  Google Scholar 

  43. Byers, L.A. et al. An epithelial-mesenchymal transition gene signature predicts resistance to EGFR and PI3K inhibitors and identifies Axl as a therapeutic target for overcoming EGFR inhibitor resistance. Clin. Cancer Res. 19, 279–290 (2013).

    Article  CAS  Google Scholar 

  44. Fischer, E.S. et al. Structure of the DDB1-CRBN E3 ubiquitin ligase in complex with thalidomide. Nature 512, 49–53 (2014).

    Article  CAS  Google Scholar 

  45. Palmer, A.C. & Kishony, R. Opposing effects of target overexpression reveal drug mechanisms. Nat. Commun. 5, 4296 (2014).

    Article  CAS  Google Scholar 

  46. Yu, C. & Golub, T.R. Multiplex methods to assay mixed cell populations simultaneously. US Patent WO 2013138585 A1 (2013).

  47. Cowley, G.S. et al. Parallel genome-scale loss of function screens in 216 cancer cell lines for the identification of context-specific genetic dependencies. Sci. Data 1, 140035 (2014).

    Article  CAS  Google Scholar 

  48. Chen, C. et al. Selective inhibitors of CYP2J2 related to terfenadine exhibit strong activity against human cancers in vitro and in vivo. J. Pharmacol. Exp. Ther. 329, 908–918 (2009).

    Article  CAS  Google Scholar 

  49. Yang, X. et al. A public genome-scale lentiviral expression library of human ORFs. Nat. Methods 8, 659–661 (2011).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the US National Cancer Institute's Cancer Target Discovery and Development (CTD2) Network (grant number U01CA176152, awarded to S.L.S.). We acknowledge the following colleagues and centers for contributing compounds and for valuable critique: Boston University, J. Bradner, P. Brown, C. Chen, J. Clardy, CNIO (Spanish National Cancer Research Center), E.J. Corey, the Drug Synthesis and Chemistry Branch (Developmental Therapeutics Program, Division of Cancer Treatment and Diagnosis, National Cancer Institute), Eutropics, J. Gutterman, E. Holson, Karyopharm, M. Meyerson, A. Myers, J. Porco, J. Qi, Sanford-Burnham, M. Serrano-Wu, M. Shair, B. Stockwell, L. Walensky, X. Wang and D. Zaharevitz. We thank A. Deik for conducting lipid profiling measurements; S. Chattopadhay, J. Law, G. Schaefer, M. Stewart, V. Viswanathan and other members of the Schreiber laboratory for advice and helpful discussions; S. Wang for helping curate the Informer Set; K. Emmith, J. Aseidu and the CSofT informatics group for development and support of cell-line and data-tracking software; A. Vrcic and the Broad Compound Management team for handling the Informer Set; J. Boehm, A. Tsherniak, A. Aguirre and the Broad Cancer Program for training and advice; the Broad Biological Samples Platform for providing CCLs; and L. Garraway and the Broad-Novartis Cancer Cell Line Encyclopedia (CCLE) team. S.L.S. is an Investigator of the Howard Hughes Medical Institute.

Author information

Authors and Affiliations

Authors

Contributions

M.G.R. designed and executed validation experiments, analyzed and interpreted data, made the figures, and wrote the manuscript; B.S.-L. designed and executed validation experiments, analyzed and interpreted data, and wrote the manuscript; J.H.C. designed and executed the primary cell-line screen; D.J.A. designed and executed validation experiments and analyzed data; E.V.P. executed the primary cell-line screen; S.G. executed validation experiments; S.J. executed validation experiments and analyzed data; M.E.C. executed the primary cell-line screen; V.L.J. executed the primary cell-line screen; N.E.B. analyzed and interpreted primary cell-line screen data; C.K.S. executed the primary cell-line screen; B.A. developed the CTRP; A.L. analyzed and interpreted data; P.M. developed the CTRP; J.D.K. wrote the manuscript; C.S.-Y.H. designed the primary cell-line screen; B.M. designed the primary cell-line screen; T.L. developed the CTRP; V.D. analyzed and interpreted primary cell-line screen data; D.A.H. designed validation experiments; C.B.C. designed and executed lipid profiling experiments and analyzed data; J.A.B. designed the primary cell-line screen; M.P. designed the primary cell-line screen; B.K.W. executed validation experiments and wrote the manuscript; P.A.C. designed, analyzed and interpreted primary cell-line screening data, directed the study and wrote the manuscript; A.F.S. designed the primary cell-line screen, interpreted data, directed the study and wrote the manuscript; S.L.S. designed the primary cell-line screen, directed the study and wrote the manuscript.

Corresponding authors

Correspondence to Paul A Clemons, Alykhan F Shamji or Stuart L Schreiber.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary Text and Figures

Supplementary Results, Supplementary Table 1 and Supplementary Figures 1–8. (PDF 15128 kb)

Supplementary Data Set 1

The cancer cell-line panel. Description of the cancer cell lines (CCLs) profiled in this experiment, including annotations from the CCLE1 for lineage and histology and experimental growth conditions (including media composition). (XLSX 78 kb)

Supplementary Data Set 2

The small-molecule informer set. Description of the small-molecule informer set profiled in this experiment, including concentrations used, protein target or activity, source and vendor information, and structure. (XLSX 100 kb)

Supplementary Data Set 3

Area-under-curve values for 481 small molecules across 823 unique CCLs. (XLSX 9176 kb)

Supplementary Data Set 4

All transcripts significantly correlated with small-molecule sensitivity across all CCLs, non-HL CCLs, and HL CCLs. A Bonferroni-corrected significance cutoff of |z| > 5.83, representing a two-tailed distribution with family-wise error rate α < 0.025 in each tail, was used. (XLSX 21297 kb)

Supplementary Data Set 5

Significant lineage-specific correlation values. All significant values where the absolute value of the z-scored Pearson correlation was greater than the value across all CCLs and non-HL CCLs are included. (XLSX 192 kb)

Supplementary Data Set 6

Correlations for 660 annotated compound-target pairs across all CCLs, non-HL CCLs, and HL CCLs. For compound-target pairs, the transcript rank (of 18,543 transcripts), z-scored Pearson correlation, and significance (|z| > 3.96; Bonferroni-corrected, two-tailed distribution with family-wise error-rate α < 0.025 in each tail) are included. (XLSX 104 kb)

Supplementary Data Set 7

Expression-sensitivity correlations for TNFRSF12A with 481 small molecules across all CCLs. Included are the number of CCLs tested per small molecule; TNFRSFS12A transcript rank (see Supplementary Fig. 1b); the Pearson expression-sensitivity correlation; the z-scored Pearson correlation; minimum, median, mean, maximum, and standard deviation of z-scored correlations from permutation testing (65,536 permutations); results of Kolmogorov-Smirnov test for normality of permutation results; empirical P-value from permutation testing; and estimated P-value from the normal cumulative distribution function. (XLSX 111 kb)

Supplementary Data Set 8

Transcript loadings from principal component analyses (PCA). Results are from PC_A1 (first PC from PCA of the entire 18,543 transcript × 481 small molecule correlation matrix across all CCLs), PC_B1 (first PC from PCA of the correlation matrix excluding HL CCLs), and PC_B2 (second PC from PCA of the correlation matrix excluding HL CCLs). The number of small molecules to which each gene was significantly correlated (excluding HL CCLs) was calculated using a Bonferroni-corrected, two-tailed distribution with family-wise error-rate α < 0.025 in each tail (|z| > 5.83). (XLSX 881 kb)

Supplementary Data Set 9

Small-molecule scores from principal component analyses. (XLSX 67 kb)

Supplementary Data Set 10

GSEA results with PC_B2 gene loadings as the input using the C2 and C5 gene sets. See www.broadinstitute.org/gsea/doc/GSEAUserGuideTEXT.htm#_Viewing_Analysis_results for analysis description. (XLSX 465 kb)

Supplementary Data Set 11

Correlations across all 481 small molecules for 654 individual transcripts most correlated with response to at least one small molecule. Included are the number of CCLs tested per small molecule; transcript rank; the Pearson expression-sensitivity correlation; the z-scored Pearson correlation; minimum, median, mean, maximum, and standard deviation of z-scored correlations from permutation testing (≥16,384 permutations); results of Kolmogorov-Smirnov test for normality of permutation results; empirical P-value from permutation testing; and estimated P -value from the normal cumulative distribution function. (XLSX 29036 kb)

Supplementary Data Set 12

GSEA results with austocystin D expression-sensitivity correlations (excluding HL CCLs) as the input using the C2 and C5 gene sets. See www.broadinstitute.org/gsea/doc/GSEAUserGuideTEXT.htm#_Viewing_Analysis_results for analysis description. (XLSX 467 kb)

Supplementary Data Set 13

Raw results from profiling cellular lipids in NCIH661 cells. Results are from 24-hour treatment with 2 μM ML239, 2 μM SC-26196, co-treatment, or DMSO, with individual replicates numbered across columns. Metabolite IDs match those from The Human Metabolome Database (www.hmdb.ca). (XLSX 53 kb)

Supplementary Data Set 14

Mapping of accession numbers for the file CCLE_Expression_Entrez_2012-10-18.res to the gene symbols used in this analysis. (XLSX 782 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rees, M., Seashore-Ludlow, B., Cheah, J. et al. Correlating chemical sensitivity and basal gene expression reveals mechanism of action. Nat Chem Biol 12, 109–116 (2016). https://doi.org/10.1038/nchembio.1986

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nchembio.1986

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