Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

The MCL1 inhibitor S63845 is tolerable and effective in diverse cancer models

Abstract

Avoidance of apoptosis is critical for the development and sustained growth of tumours. The pro-survival protein myeloid cell leukemia 1 (MCL1) is overexpressed in many cancers, but the development of small molecules targeting this protein that are amenable for clinical testing has been challenging. Here we describe S63845, a small molecule that specifically binds with high affinity to the BH3-binding groove of MCL1. Our mechanistic studies demonstrate that S63845 potently kills MCL1-dependent cancer cells, including multiple myeloma, leukaemia and lymphoma cells, by activating the BAX/BAK-dependent mitochondrial apoptotic pathway. In vivo, S63845 shows potent anti-tumour activity with an acceptable safety margin as a single agent in several cancers. Moreover, MCL1 inhibition, either alone or in combination with other anti-cancer drugs, proved effective against several solid cancer-derived cell lines. These results point towards MCL1 as a target for the treatment of a wide range of tumours.

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: S63845 binds to the BH3-binding groove of MCL1 and kills tumour cell lines by inducing BAX/BAK-dependent apoptosis.
Figure 2: S63845 is effective against haematological cancer-derived cell lines in vitro and in vivo.
Figure 3: S63845 is effective against AML samples in vitro and in vivo, but does not readily kill normal human haematopoietic progenitor cells.
Figure 4: Activity of S63845 against solid-tumour-derived cell lines in vitro with or without other targeted agents.
Figure 5: S63845 is well tolerated in mice at doses that efficiently kill tumour cells.

Similar content being viewed by others

References

  1. Green, D. R. & Llambi, F. Cell death signaling. Cold Spring Harb. Perspect. Biol. 7, 1–24 (2015)

    Article  CAS  Google Scholar 

  2. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011)

    Article  CAS  PubMed  Google Scholar 

  3. Czabotar, P. E., Lessene, G., Strasser, A. & Adams, J. M. Control of apoptosis by the BCL-2 protein family: implications for physiology and therapy. Nat. Rev. Mol. Cell Biol. 15, 49–63 (2014)

    Article  CAS  PubMed  Google Scholar 

  4. Petros, A. M. et al. Rationale for Bcl-xL/Bad peptide complex formation from structure, mutagenesis, and biophysical studies. Protein Sci. 9, 2528–2534 (2000)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Delbridge, A. R. D., Grabow, S., Strasser, A. & Vaux, D. L. Thirty years of BCL-2: translating cell death discoveries into novel cancer therapies. Nat. Rev. Cancer 16, 99–109 (2016)

    Article  CAS  PubMed  Google Scholar 

  6. Wilson, W. H. et al. Navitoclax, a targeted high-affinity inhibitor of BCL-2, in lymphoid malignancies: a phase 1 dose-escalation study of safety, pharmacokinetics, pharmacodynamics, and antitumour activity. Lancet Oncol. 11, 1149–1159 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Roberts, A. W. et al. Substantial susceptibility of chronic lymphocytic leukemia to BCL2 inhibition: results of a phase I study of navitoclax in patients with relapsed or refractory disease. J. Clin. Oncol. 30, 488–496 (2012)

    Article  CAS  PubMed  Google Scholar 

  8. Roberts, A. W. et al. Targeting BCL2 with venetoclax in relapsed chronic lymphocytic leukemia. N. Engl. J. Med. 374, 311–322 (2016)

    Article  CAS  PubMed  Google Scholar 

  9. Juin, P., Geneste, O., Gautier, F., Depil, S. & Campone, M. Decoding and unlocking the BCL-2 dependency of cancer cells. Nat. Rev. Cancer 13, 455–465 (2013)

    Article  CAS  PubMed  Google Scholar 

  10. Beroukhim, R. et al. The landscape of somatic copy-number alteration across human cancers. Nature 463, 899–905 (2010)

    Article  CAS  ADS  PubMed  PubMed Central  Google Scholar 

  11. Akgul, C. Mcl-1 is a potential therapeutic target in multiple types of cancer. Cell. Mol. Life Sci. 66, 1326–1336 (2009)

    Article  CAS  PubMed  Google Scholar 

  12. Glaser, S. P. et al. Anti-apoptotic Mcl-1 is essential for the development and sustained growth of acute myeloid leukemia. Genes Dev. 26, 120–125 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Kelly, G. L. et al. Targeting of MCL-1 kills MYC-driven mouse and human lymphomas even when they bear mutations in p53. Genes Dev. 28, 58–70 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Koss, B. et al. Requirement for antiapoptotic MCL-1 in the survival of BCR–ABL B-lineage acute lymphoblastic leukemia. Blood 122, 1587–1598 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Grabow, S., Delbridge, A. R., Valente, L. J. & Strasser, A. MCL-1 but not BCL-XL is critical for the development and sustained expansion of thymic lymphoma in p53-deficient mice. Blood 124, 3939–3946 (2014)

    Article  CAS  PubMed  Google Scholar 

  16. Spinner, S. et al. Re-activation of mitochondrial apoptosis inhibits T-cell lymphoma survival and treatment resistance. Leukemia 30, 1520–1530 (2016)

    Article  CAS  PubMed  Google Scholar 

  17. Gong, J.-N. et al. Hierarchy for targeting pro-survival BCL2 family proteins in multiple myeloma: pivotal role of MCL1. Blood blood-2016- 03–704908 (2016)

  18. Wei, G. et al. Chemical genomics identifies small-molecule MCL1 repressors and BCL-XL as a predictor of MCL1 dependency. Cancer Cell 21, 547–562 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Leverson, J. D. et al. Potent and selective small-molecule MCL-1 inhibitors demonstrate on-target cancer cell killing activity as single agents and in combination with ABT-263 (navitoclax). Cell Death Dis. 6, e1590 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Pelz, N. F. et al. Discovery of 2-indole-acylsulfonamide myeloid cell leukemia 1 (Mcl-1) inhibitors using fragment-based methods. J. Med. Chem. 59, 2054–2066 (2016)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Nhu, D., Lessene, G., Huang, D. & Burns, C. J. Small molecules targeting Mcl-1: the search for a silver bullet in cancer therapy. MedChemComm 7, 778–787 (2016)

    Article  CAS  Google Scholar 

  22. Opferman, J. T. et al. Obligate role of anti-apoptotic MCL-1 in the survival of hematopoietic stem cells. Science 307, 1101–1104 (2005)

    Article  CAS  ADS  PubMed  Google Scholar 

  23. Wang, X. et al. Deletion of MCL-1 causes lethal cardiac failure and mitochondrial dysfunction. Genes Dev. 27, 1351–1364 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Thomas, R. L. et al. Loss of MCL-1 leads to impaired autophagy and rapid development of heart failure. Genes Dev. 27, 1365–1377 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Lessene, G., Czabotar, P. E. & Colman, P. M. BCL-2 family antagonists for cancer therapy. Nat. Rev. Drug Discov. 7, 989–1000 (2008)

    Article  CAS  PubMed  Google Scholar 

  26. Bruncko, M. et al. Structure-guided design of a series of MCL-1 inhibitors with high affinity and selectivity. J. Med. Chem. 58, 2180–2194 (2015)

    Article  CAS  PubMed  Google Scholar 

  27. Tunquist, B. J., Woessner, R. D. & Walker, D. H. Mcl-1 stability determines mitotic cell fate of human multiple myeloma tumor cells treated with the kinesin spindle protein inhibitor ARRY-520. Mol. Cancer Ther. 9, 2046–2056 (2010)

    Article  CAS  PubMed  Google Scholar 

  28. Oltersdorf, T. et al. An inhibitor of Bcl-2 family proteins induces regression of solid tumours. Nature 435, 677–681 (2005)

    Article  CAS  ADS  PubMed  Google Scholar 

  29. Souers, A. J. et al. ABT-199, a potent and selective BCL-2 inhibitor, achieves antitumor activity while sparing platelets. Nat. Med. 19, 202–208 (2013)

    Article  CAS  PubMed  Google Scholar 

  30. Aubrey, B. J. et al. An inducible lentiviral guide RNA platform enables the identification of tumor-essential genes and tumor-promoting mutations in vivo. Cell Reports 10, 1422–1432 (2015)

    Article  CAS  PubMed  Google Scholar 

  31. Derenne, S. et al. Antisense strategy shows that Mcl-1 rather than Bcl-2 or Bcl-x(L) is an essential survival protein of human myeloma cells. Blood 100, 194–199 (2002)

    Article  CAS  PubMed  Google Scholar 

  32. Zhang, B., Gojo, I. & Fenton, R. G. Myeloid cell factor-1 is a critical survival factor for multiple myeloma. Blood 99, 1885–1893 (2002)

    Article  CAS  PubMed  Google Scholar 

  33. Lee, E. F. et al. A novel BH3 ligand that selectively targets Mcl-1 reveals that apoptosis can proceed without Mcl-1 degradation. J. Cell Biol. 180, 341–355 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Chen, L. et al. Differential targeting of prosurvival Bcl-2 proteins by their BH3-only ligands allows complementary apoptotic function. Mol. Cell 17, 393–403 (2005)

    Article  CAS  PubMed  Google Scholar 

  35. Touzeau, C. et al. The Bcl-2 specific BH3 mimetic ABT-199: a promising targeted therapy for t(11;14) multiple myeloma. Leukemia 28, 210–212 (2014)

    Article  CAS  PubMed  Google Scholar 

  36. Lazareth, A. et al. MB4-2 breakpoint in MMSET combined with del(17p) defines a subset of t(4;14) multiple myeloma with very poor prognosis. Haematologica 100, e471–e474 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Zhang, H. et al. Mcl-1 is critical for survival in a subgroup of non-small-cell lung cancer cell lines. Oncogene 30, 1963–1968 (2011)

    Article  CAS  PubMed  Google Scholar 

  38. Xiao, Y. et al. MCL-1 Is a key determinant of breast cancer cell survival: validation of MCL-1 dependency utilizing a highly selective small molecule inhibitor. Mol. Cancer Ther. 14, 1837–1847 (2015)

    Article  CAS  PubMed  Google Scholar 

  39. Cragg, M. S., Harris, C., Strasser, A. & Scott, C. L. Unleashing the power of inhibitors of oncogenic kinases through BH3 mimetics. Nat. Rev. Cancer 9, 321–326 (2009)

    Article  CAS  PubMed  Google Scholar 

  40. Cragg, M. S. et al. Treatment of B-RAF mutant human tumor cells with a MEK inhibitor requires Bim and is enhanced by a BH3 mimetic. J. Clin. Invest. 118, 3651–3659 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Corcoran, R. B. et al. Synthetic lethal interaction of combined BCL-XL and MEK inhibition promotes tumor regressions in KRAS mutant cancer models. Cancer Cell 23, 121–128 (2013)

    Article  CAS  PubMed  Google Scholar 

  42. Perciavalle, R. M. et al. Anti-apoptotic MCL-1 localizes to the mitochondrial matrix and couples mitochondrial fusion to respiration. Nat. Cell Biol. 14, 575–583 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Fang, C. et al. Single diastereomer of a macrolactam core binds specifically to myeloid cell leukemia 1 (MCL1). ACS Med. Chem. Lett. 5, 1308–1312 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Moon, A. F., Mueller, G. A., Zhong, X. & Pedersen, L. C. A synergistic approach to protein crystallization: combination of a fixed-arm carrier with surface entropy reduction. Protein Sci. 19, 901–913 (2010)

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Kabsch, W. XDS. Acta Crystallogr. D Biol. Crystallogr. 66, 125–132 (2010)

    Article  CAS  Google Scholar 

  46. Vagin, A. & Teplyakov, A. MOLREP: an automated program for molecular replacement. J. Appl. Cryst. 30, 1022–1025 (1997)

    Article  CAS  Google Scholar 

  47. Murshudov, G. N., Vagin, A. A. & Dodson, E. J. Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr. D Biol. Crystallogr. 53, 240–255 (1997)

    Article  CAS  PubMed  Google Scholar 

  48. Nikolovska-Coleska, Z. et al. Development and optimization of a binding assay for the XIAP BIR3 domain using fluorescence polarization. Anal. Biochem. 332, 261–273 (2004)

    Article  CAS  PubMed  Google Scholar 

  49. Murray, J. B., Roughley, S. D., Matassova, N. & Brough, P. A. Off-rate screening (ORS) by surface plasmon resonance. An efficient method to kinetically sample hit to lead chemical space from unpurified reaction products. J. Med. Chem. 57, 2845–2850 (2014)

    Article  CAS  PubMed  Google Scholar 

  50. Meiby, E. et al. Fragment screening by weak affinity chromatography: comparison with established techniques for screening against HSP90. Anal. Chem. 85, 6756–6766 (2013)

    Article  CAS  PubMed  Google Scholar 

  51. Adams, J. M. et al. The c-myc oncogene driven by immunoglobulin enhancers induces lymphoid malignancy in transgenic mice. Nature 318, 533–538 (1985)

    Article  CAS  ADS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank S. Courtade-Gaiani and D. Valour for bioinformatics support, E. Borges for assistance on manuscript formatting, E. Schneider and C. Wagner-Legrand, H. Johnson, G. Siciliano and K. Hughes for technical help for in vivo studies, N. Whitehead for protein production support, P. Bouillet and L. A. O’Reilly for assistance with histology, M. Fallowfield, J. D’Alessandro, L. Terry, V. Lemesre, J.-P. Galizzi and C. de la Moureyre for in vitro assay support, H. Simmonite for analytical support and L. Andrieu, L. Montane and A. Schmutz for biostatistical support. Research at WEHI is supported by the National Health and Medical Research Council Australia (NHMRC, GNT1016647, GNT1016701, GNT1020363, GNT1086291, GNT1049720, GNT1057742, GNT1079560), the Leukemia and Lymphoma Society (SCOR grant 7001-03), The Cancer Council (1086157 GLK, grant in aid to A.W.R. and D.C.S.H.), The Kay Kendall Leukemia Fund Intermediate Fellowship (KKL331 to G.L.K.), the Victoria Cancer Agency, the Australian Cancer Research Foundation, a Victorian State Government Operational Infrastructure Support (OIS) grant and the estate of Anthony (Toni) Redstone OAM.

Author information

Authors and Affiliations

Authors

Contributions

A.K., Z.S., J.D., M.Cs., A.Pa., Z.B.S., S.S., G.R., A.Pr., B.B., L.O., G.B. and C.G. supervised and performed the chemistry. A.K., Z.B.S., J.M., J.D. and I.C. performed the drug design and molecular modelling. A.L.M., G.L.T.-B., G.L.K., J.-N.G., D.M.M., A.Stu., D.S., C.D.R., G.P., C.C., G.G. and N.C. performed cell based experiments. G.L.K., G.L.-A., A.-M.G., F.G., M.S.B., L.C.A.G. and M.J.H. performed the in vivo experiments. B.J.A. contributed to the histology analysis. J.M., A.R., A.Su., P.D., N.M., J.S. and C.P. produced recombinant proteins, performed biochemical assays and crystallographic studies. M.Ch., G.L.K., A.B. and M.J.H. designed the in vivo experiments. F.M., N.G.-S. and B.L. designed and performed the bioinformatic analysis. A.-L.M., J.M., J.D., G.L.K., F.C., J.A.H., A.W.R., D.C.S.H., A.H.W., A.Str., G.L. and O.G. supervised the studies, designed the experiments and interpreted the results. A.Str., G.L., D.C.S.H. and O.G. wrote the manuscript with the assistance of A.L.M., G.L.K. and the other authors.

Corresponding author

Correspondence to Olivier Geneste.

Ethics declarations

Competing interests

G.L.T.-B., A.L.M., G.G., A.Stu., N.C., M.Ch., A.B., G.L.-A., A.-M.G., F.G., O.G., J.A.H., B.L., N.G.-S., F.C., F.M., A.K., Z.Szl., M.Cs., A.Pa., Z.Sza., S.S., G.R., L.O., A.Pr., B.B. and G.B. are employees of Servier. N.M., C.P., J.S., I.C., C.G., J.D., A.R., A.Su., P.D. and J.M. are employees of Vernalis (R&D) Ltd., G.L., A.Str., G.L.K., M.J.H., J.G., D.S., C.D.R., A.W.R., C.C., M.S.B. and D.C.S.H. are employees of the Walter and Eliza Hall Institute of Medical Research, which receives research funding and milestone payments in relation to venetoclax (ABT-199). G.L., A.Str., G.L.K., M.J.H., J.-N.G., D.S., C.D.R., A.W.R., M.S.B. and D.C.S.H. receive research funding from Servier. A.H.W. serves on the advisory board for Servier and receives research funding from Servier.

Additional information

Reviewer Information Nature thanks S. Fletcher, F. Stegmeier, G. Wagner and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Extended data figures and tables

Extended Data Figure 1 Biophysical characterization of the binding of MCL1 inhibitors to human and mouse MCL1 and serum effect on their cellular potency.

a, Affinity comparison of binding of S63845 to human or mouse MCL1 (n = 4 and 3 biological replicates for human and mouse, respectively; see source data). b, Binding affinity data of S63845 and A-1210477 for MCL1 (fluorescence polarization (FP) and SPR), BCL-2 (fluorescence polarization) and BCL-XL (n = 2 for fluorescence polarization and n = 3–4 for SPR, see source data). c, SPR sensograms, KD and kinetic parameters for the binding of S63845 and A-1210477 to human MCL1 (n = 4 and 3 biological replicates for S63845 and A-1210477, respectively). d, Impact of serum concentration (FBS, fetal bovine serum) on the biological activity of the MCL1 inhibitors, S63845 and A-1210477, in H929 cells (n = 3 biological replicates, see source data).

Source data

Extended Data Figure 2 Targeting MCL1 genetically in H929 cells induces caspase-mediated cell death.

a, H929 cells rely on MCL1 for their survival. The viability of H929 cells was determined with CellTiter-Glo assays 72 h after addition of doxocycline to induce expression of sgRNAs to target BCL-2, BCL-XL, BCL-W, MCL1 or BFL1 using CRISPR/Cas9 technology30. Two sgRNAs were tested for each gene: sg#1 (no fill pattern) and sg#2 (slash pattern). Mean and individual data points of n = 2 biological replicates performed in triplicate are shown (see source data). b, Targeting MCL1 by CRISPR/Cas9 causes rapid cell death. The viability (CellTiter-Glo assays) of H929 cells stably expressing Cas9 and containing an inducible expression construct for a sgRNA to target MCL1 was determined 0–96 h after addition of doxocycline. Mean and individual data points of at least n = 2 biological repicates performed in triplicate are shown (see source data). c, Caspases mediate the killing of H929 cells when MCL1 is genetically targeted. The viability (propidium iodide uptake determined by flow cytometry) of H929 cells after switching on the expression of the indicated sgRNAs by the addition of doxocycline (+DOX) was measured after culture in the presence (+) or absence (−) of the broad-spectrum caspase inhibitor, QVD-OPh (25 μM). Data (mean ± s.d.) are derived from one experiment performed in triplicate (see source data). EV, empty vector.

Source data

Extended Data Figure 3 Impact of treatment with S63845 on the interaction of MCL1 with pro-apoptotic BCL-2 family members and on the level and stability of MCL1.

a, HeLa cells transduced with Flag–BCL-XL, Flag–BCL-2 or Flag–MCL1 expression constructs were treated for 4 h with increasing concentrations of S63845, before immunopreciptation using anti-FLAG antibody. Immunoprecipitates and total inputs were analysed by immunoblotting for the FLAG-tagged proteins as well as the associated BAK and BAX proteins. For gel source images, see Supplementary Fig. 1. b, HCT-116 cells were incubated for 16 h with increasing concentrations of the proteasome inhibitor, MG132, as a control, before assessing the levels of MCL1, BCL-XL and actin (protein loading control) by immunoblotting. c, HCT-116 cells were pre-incubated for 1 h with DMSO or 1 μM S63845 and then treated with the translation inhibitor, emetin (20 μg ml−1), for 0.5, 1, 2 or 4 h. Determination of MCL1 protein levels was performed by immunoblotting (actin served as a loading control). d, Densitometry of MCL1 immunoblot (Extended Data Fig. 3c, long exposure for DMSO-treated cells and short exposure for S63845-treated cells) and MCL1 half-life estimation (21 min for DMSO-treated cells and 79 min for S63845-treated cells). e, HCT-116 cells were incubated for 16 h with 300 nM S63845, or transfected with 3× Flag–MCL1 expressing plasmid (400 ng, for 48 h) before assessing MCL1 mRNA level by reverse transcriptase–PCR quantification (see source data).

Source data

Extended Data Figure 4 S63845 induces apoptosis of sensitive tumour derived cell lines.

a, PARP cleavage in H929 cells after treatment for 6 h with the MCL1 inhibitors, S63845 or A-1210477. Mean and individual data points from three biological replicates are shown (see source data). b, Treatment with S63845 causes mitochondrial release of cytochrome c in sensitive cells. H929 cells were incubated for 4 h with increasing concentrations of S63845 before assessing cytochrome c abundance in the cytosolic fraction by immunoblotting. LDH was used as a protein loading control. For gel source images, see Supplementary Fig. 1. c, Treatment with MCL1 inhibitors causes PARP cleavage in sensitive cells. AMO1 and MV4-11 cells were treated for 6 h with increasing concentrations of the MCL1 inhibitors, S63845 or A-1210477. Cleaved PARP (a marker of apoptotic cell death) was detected by a Meso Scale eletrochemiluminescence assay. Means and individual data points from three biological replicates are shown (see source data).

Source data

Extended Data Figure 5 Correlation between the sensitivity of multiple-myeloma-derived cell lines to S63845 and killing by the MCL1-selective ligand BIM2A. BAX/BAK dependency for S63845-induced killing in KMS-12-PE and AMO1 myeloma cells. Expression of pro-survival BCL-2 family proteins across a panel of multiple-myeloma-derived lines.

a, Correlation between the sensitivity of multiple myeloma (MM)-derived cell lines to S63845 and killing by enforced expression of the MCL1-selective ligand BIM2A. The sensitivity of 25 multiple-myeloma-derived cell lines to S63845 was plotted against the viability of the same cells infected with a lentivirus that expresses the MCL1-selective ligand BIM2A (ref. 33) (r = 0.55) or with a lentivirus expressing the BCL-2, BCL-XL and BCL-W binding ligand BIM–BAD (ref. 34) (r = −0.22) (data from Extended Fig. 5b, see source data). The r and P values were determined by Pearson correlation analysis. b, Table of IC50 values and per cent viability as defined in Extended Data Fig. 5a measured using the CellTiter-Glo assay. The IC50 values were calculated from 2–5 biological replicates (see source data, mean ± s.d.). As a comparison, the dependence of these multiple myeloma cell lines on MCL1 or BCL-2/BCL-XL/BCL-W was validated by a genetic approach. The survival of multiple myeloma cell lines 24 h after enforced expression of BIM2A (which inhibits MCL1) or BIM–BAD (which inhibits BCL-2, BCL-XL and BCL-W) was determined by CellTiter-Glo assay and data were normalized to the survival observed when the inert BIM variant (BIM4E) was expressed. These BIM variants have been previously described13,34; the data with the BIM variants, shown for comparison, are extracted from a larger data set17. The multiple myeloma cell lines are grouped based on their characteristic cytogenetic abnormality. c, BAX/BAK-dependent killing of KMS-12-PE and AMO1 myeloma cells by S63845. KMS-12-PE or AMO1 myeloma cells were engineered to lack BAX and BAK using CRISPR/Cas9 technology1 (see Methods). Cells expressing the sgRNA empty vector (sgEV) were also generated to serve as negative controls. These cells were treated with S63845 and their viability was determined after 48 h by the propidium iodide exclusion assay. Data shown are mean viability ± s.d. derived from three (KMS-12-PE) or two (AMO1) biological replicates (see source data). d, Expression of MCL1 and related pro-survival BCL-2 proteins across a selected panel of multiple-myeloma-derived lines (shown in Fig. 2a). The levels of MCL1, BCL-2 and BCL-XL were compared between the six multiple-myeloma-derived cell lines most sensitive to S63845 (left, IC50 < 10 nM; data extracted from Extended Data Table 1) and six least sensitive ones (right, IC50 > 100 nM). For gel source images, see Supplementary Fig. 1.

Source data

Extended Data Figure 6 Anti-tumour efficacy and effect of dose scheduling of S63845 in H929 and AMO1 multiple myeloma xenograft models. The in vitro activity of S63845 in human Burkitt-lymphoma-derived cell lines and in vivo activity of S63845 on individual Eμ-Myc lymphoma cell lines.

a, Anti-tumour effect of S63845 in H929 multiple myeloma xenograft models. Mean ± s.e.m. tumour volumes of eight animals per treatment group are shown (***P < 0.001 compared to vehicle). For tumour source data, see Supplementary Fig. 2. Tumour volumes of 8 individual mice per treatment group (shown in Fig. 2b) are shown. b, SCID mice were inoculated with 5 × 106 AMO1 cells and randomized nine days after grafting. After randomization, mice were either not treated (black line) or treated by i.v. administration of S63845, every day for 5 consecutive days at 25 mg per kg body weight (red line). For tumour source data, see Supplementary Fig. 2. c, Comparison of S63845 anti-tumour efficacy in the AMO1 model after administration at 25 mg per kg body weight either twice a week for two weeks (red diamonds) or once a week for two weeks (orange triangles). Mean ± s.e.m. tumour volumes of eight animals per treatment group are shown (***P < 0.001 compared to vehicle). For tumour source data, see Supplementary Fig. 2. d, Body weight evaluation of mice that were either left untreated (black squares) or treated with S63845 at 25 mg per kg body weight twice a week for two weeks (red diamonds) or once a week for two weeks (orange triangles). Mean ± s.e.m. body weights of eight animals per treatment group are shown. For source data, see Supplementary Fig. 2. e, Sensitivity of human Burkitt-lymphoma (BL)-derived cell lines to S63845. IC50 values of Burkitt-lymphoma-derived cell lines were determined following 24 h in vitro exposure to increasing doses of S63845. Average IC50 values were determined from three biological replicates (each carried out in triplicate, see source data); s.d. are shown. f, Representative haematoxylin and eosin (H&E)-stained sections of spleens and bone marrow from mice transplanted with Eμ-Myc lymphoma cells 4 days after the end of treatment with either vehicle or S63845 from Fig. 2e. Scale bars represent 553 μm for the ×5 objective, 272 μm for the ×10 objective, 133 μm for the ×20 objective and 68 μm for the ×40 objective. g, Response of different Eμ-Myc lymphoma-derived cell lines to in vivo treatment with S63845. Survival curves of C57BL/6–Ly5.1+ mice transplanted with each of the six independent Eμ-Myc lymphoma-derived cell lines that were treated on days 5–9 after transplant with either vehicle (black line) or 25 mg per kg body weight S63845 (red line), by tail vein injection (see source data). A composite of the data from all of these Eμ-Myc lymphoma-derived cell lines is presented in Fig. 2e.

Source data

Extended Data Figure 7 Spearman correlations distribution in haematological cell lines and AML data on patient samples.

a, Waterfall plots on Spearman correlations between human transcriptome (RNA-seq gene FPKM from CCLE database) and S63845 pIC50 (pIC50 = −log10(IC50)) measured on a panel of 24 haematological cancer-derived cell lines. Raw P values were adjusted for multiple testing using Benjamini-Hochberg correction. Only correlations with false discovery rate (FDR) < 0.05 were retained for the analysis. Spearman correlations were ranked and plotted as a distribution. The threshold was set at r = 0.4, which corresponded to half of the maximum correlation. b, S63845 activity against primary AML samples relative to comparable drugs. Cell viability was assessed by flow cytometric enumeration of Sytox Blue-negative cells after gating on blast cells. The concentration of each drug (in μM) causing 50% lethality for each sample (LC50) is shown. Cytogenetic risk for each sample is defined as intermediate (I) or adverse (A). NA indicates not available. Cell viability (%) after 48 h of treatment with vehicle is also shown. c, AML cell colony-forming assays showing clonogenic capacity of CD34+ sorted cells from two G-CSF mobilized normal, healthy donors (red lines and filled circles) and primary samples from 3–6 patients with AML (black symbols and lines) after 14–21 days exposure to S63845, ara-C or idarubicin, normalized to vehicle control. Error bars show ± s.d. of colony counts from duplicate plates (see source data).

Source data

Extended Data Figure 8 S63845 activity in solid tumour-derived cell lines, correlation with BCL-XL mRNA expression and impact of treatment with the MEK1/2 inhibitor, trametenib, the HER-2 inhibitor, lapatinib, the B-RAF inhibitor, PLX-4032, or the EGFR inhibitor, tarceva, on the levels of p-ERK and the pro-apoptotic protein BIM.

a, Viability of NSCLC-, melanoma- and breast tumour-derived cell lines treated with S63845. Sensitive cell lines (IC50 < 0.1 μM) are indicated in red, moderately sensitive cell lines (0.1 μM < IC50 < 1 μM) in blue and insensitive cell lines (IC50 > 1 μM) in green. Mean IC50 and individual data points from n = 2 biological replicates are shown (see source data). b, The impact of the BCL-XL mRNA expression levels on the sensitivity of a panel of 110 solid tumour-derived cell lines to S63845 (see source data). BCL-XL mRNA expression levels (RNA-seq) were extracted from the CCLE database. The Pearson correlation coefficient and P value are shown. c, Waterfall plots on Spearman correlations between human transcriptome (RNA-seq Gene FPKM from CCLE database) and S63845 pIC50 measured on a panel of 110 solid tumour cell lines. Raw P values were adjusted for multiple testing using Benjamini-Hochberg correction. Only correlations with FDR < 0.05 were retained for analysis. Spearman correlations were ranked and plotted as a distribution. The threshold was set at r = 0.4, which corresponded to half of the maximum correlation. d, Tumour cells were incubated with the indicated inhibitors for 16 h, before assessing the protein levels of BIM, ERK, phosphorylated (that is, activated) ERK (P-ERK) and actin (used as a protein loading control) by immunoblotting. For gel source images, see Supplementary Fig. 1. e, SK-MEL-28 cells were treated with S63845 with or without further addition of the MEK1/2 inhibitor, trametenib. Data are normalized to baseline cell survival at the time of initial drug treatment (day 0). Mean and individual data points from three biological replicates are shown (see source data).

Source data

Extended Data Figure 9 Tolerability of S63845 in mice and impact of treatment on healthy tissues at doses that prevent tumour expansion.

a, Impact of treatment with S63845 at doses that prevent tumour expansion on healthy tissues in mice. Healthy (not tumour-burdened) C57BL/6–Ly5.1+ mice were treated for 5 consecutive days with either vehicle or 25 mg per kg body weight S63845, and analysed 3 days after treatment had finished. Student’s unpaired t-test was carried out to compare the organ weights of spleen, thymus, lymph nodes, liver and kidney between vehicle and drug-treated mice. The organ weights were not significantly altered in response to treatment with S63845 (see source data). b, Representative haematoxylin and eosin-stained sections of the heart, liver, muscle and kidneys taken from vehicle- or S63845-treated mice at day 9 following treatment for 5 days (days 1–5). Scale bars represent 112 μm for the ×20 objective and 57 μm for the ×40 objective. c, Impact of treatment with increasing doses of S63845 on C57BL/6 mice. Kaplan–Meier survival curves are shown for mice treated with vehicle or with 25, 40, 50 or 60 mg per kg body weight S63845 for 5 consecutive days (see source data). There were 3 female and 3 male 7–8-week-old mice in each treatment group (except for the 40 mg per kg group where n = 5 because one mouse from this treatment group had to be killed owing to an injury incurred during drug administration). Using a Mantel–Cox test only the 60 mg per kg body weight dose showed significant changes relative to treatment with vehicle, ***P = 0.0006. d, Blood cell count analysis of the mice described in b (see source data). The mice were analysed as they were killed, or for the mice surviving the entire treatment, 3 days after the 5 day-treatment had been completed. Unpaired two-tailed t-test was performed and compared to vehicle. For PLT counts, compared to vehicle, **P = 0.0027 for the 25 mg kg dose, ***P = 0.0005 for the 40 mg per kg dose and *P = 0.0245 for the 50 mg per kg dose of S63845. For the RBC counts, compared to vehicle, *P = 0.0312 for the 50 mg per kg dose and *P = 0.0242 for the 60 mg per kg dose of S63845.

Source data

Extended Data Table 1 Data collection and refinement statistics

Supplementary information

Supplementary Information

This file contains Supplementary Figures 1-2, Supplementary Tables 1-2 and Supplementary Methods. (PDF 7845 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kotschy, A., Szlavik, Z., Murray, J. et al. The MCL1 inhibitor S63845 is tolerable and effective in diverse cancer models. Nature 538, 477–482 (2016). https://doi.org/10.1038/nature19830

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature19830

This article is cited by

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: 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