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:

RAS nucleotide cycling underlies the SHP2 phosphatase dependence of mutant BRAF-, NF1- and RAS-driven cancers

Abstract

Oncogenic alterations in the RAS/RAF/MEK/ERK pathway drive the growth of a wide spectrum of cancers. While BRAF and MEK inhibitors are efficacious against BRAFV600E-driven cancers, effective targeted therapies are lacking for most cancers driven by other pathway alterations, including non-V600E oncogenic BRAF, RAS GTPase-activating protein (GAP) NF1 (neurofibromin 1) loss and oncogenic KRAS. Here, we show that targeting the SHP2 phosphatase (encoded by PTPN11) with RMC-4550, a small-molecule allosteric inhibitor, is effective in human cancer models bearing RAS–GTP-dependent oncogenic BRAF (for example, class 3 BRAF mutants), NF1 loss or nucleotide-cycling oncogenic RAS (for example, KRASG12C). SHP2 inhibitor treatment decreases oncogenic RAS/RAF/MEK/ERK signalling and cancer growth by disrupting SOS1-mediated RAS–GTP loading. Our findings illuminate a critical function for SHP2 in promoting oncogenic RAS/MAPK pathway activation in cancers with RAS–GTP-dependent oncogenic BRAF, NF1 loss and nucleotide-cycling oncogenic KRAS. SHP2 inhibition is a promising molecular therapeutic strategy for patients with cancers bearing these oncogenic drivers.

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: RMC-4550 is a potent, selective, allosteric inhibitor of SHP2.
Fig. 2: SHP2 inhibition suppresses growth and RAS/MAPK signalling in cancer cell lines with class 3 BRAF mutations.
Fig. 3: SHP2 inhibition suppresses growth and RAS/MAPK signalling in cancer cell lines driven by NF1LOF mutation.
Fig. 4: SHP2 inhibition suppresses growth and RAS/MAPK signalling in a range of cancer cell lines driven by KRASG12 mutations.
Fig. 5: The mechanism of action of SHP2 inhibition is dependent on SOS1.
Fig. 6: SHP2 inhibition suppresses tumour growth and RAS/MAPK signalling in xenograft models of NF1LOF- and KRASG12C-driven cancer.
Fig. 7: SHP2 inhibition suppresses tumour growth in PDX models of class 3 BRAF-, NF1LOF- and KRASG12C-driven cancer.

Similar content being viewed by others

References

  1. Dhillon, A. S., Hagan, S., Rath, O. & Kolch, W. MAP kinase signalling pathways in cancer. Oncogene 26, 3279–3290 (2007).

    Article  PubMed  CAS  Google Scholar 

  2. Bos, J. L., Rehmann, H. & Wittinghofer, A. GEFs and GAPs: critical elements in the control of small G proteins. Cell 130, 385 (2007).

    Article  CAS  Google Scholar 

  3. The Cancer Genome Atlas Research Network Comprehensive molecular profiling of lung adenocarcinoma. Nature 511, 543–550 (2014).

  4. Krauthammer, M. et al. Exome sequencing identifies recurrent mutations in NF1 and RASopathy genes in sun-exposed melanomas. Nat. Genet. 47, 996–1002 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  5. Redig, A. J. et al. Clinical and molecular characteristics of NF1-mutant lung cancer. Clin. Cancer Res. 22, 3148–3156 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  7. Bollag, G. et al. Vemurafenib: the first drug approved for BRAF-mutant cancer. Nat. Rev. Drug Discov. 11, 873–886 (2012).

    Article  PubMed  CAS  Google Scholar 

  8. Nieto, P. et al. A Braf kinase-inactive mutant induces lung adenocarcinoma. Nature 548, 239–243 (2017).

    PubMed  PubMed Central  CAS  Google Scholar 

  9. Yao, Z. et al. Tumours with class 3 BRAF mutants are sensitive to the inhibition of activated RAS. Nature 548, 234–238 (2017).

    PubMed  PubMed Central  CAS  Google Scholar 

  10. Lito, P., Solomon, M., Li, L.-S., Hansen, R. & Rosen, N. Allele-specific inhibitors inactivate mutant KRAS G12C by a trapping mechanism. Science 351, 604–608 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Ostrem, J. M., Peters, U., Sos, M. L., Wells, J. A. & Shokat, K. M. K-Ras(G12C) inhibitors allosterically control GTP affinity and effector interactions. Nature 503, 548–551 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Patricelli, M. P. et al. Selective inhibition of oncogenic KRAS output with small molecules targeting the inactive state. Cancer Discov. 6, 316–329 (2016).

    Article  PubMed  CAS  Google Scholar 

  13. Prahallad, A. et al. PTPN11 is a central node in intrinsic and acquired resistance to targeted cancer drugs. Cell Rep. 12, 1978–1985 (2015).

    Article  PubMed  CAS  Google Scholar 

  14. Niihori, T. et al. Functional analysis of PTPN11/SHP-2 mutants identified in Noonan syndrome and childhood leukemia. J. Hum. Genet. 50, 192–202 (2005).

    Article  PubMed  CAS  Google Scholar 

  15. Tartaglia, M. & Gelb, B. D. Germ-line and somatic PTPN11 mutations in human disease. Eur. J. Med. Genet. 48, 81–96 (2005).

    Article  PubMed  Google Scholar 

  16. Chen, Y.-N. P. et al. Allosteric inhibition of SHP2 phosphatase inhibits cancers driven by receptor tyrosine kinases. Nature 535, 148–152 (2016).

    Article  PubMed  CAS  Google Scholar 

  17. Wildes, D. et al. Allosteric inhibition of SHP2 variants containing cancer-associated activating mutations. Proc. Ann. Am. Ass. Cancer Res. 109, 4877 (2018).

    Google Scholar 

  18. Garcia-Fortanet, J. et al. Allosteric inhibition of SHP2: identification of a potent, selective, and orally efficacious phosphatase inhibitor. J. Med. Chem. 59, 7773–7782 (2016).

    Article  PubMed  CAS  Google Scholar 

  19. Poulikakos, P. I. et al. RAF inhibitor resistance is mediated by dimerization of aberrantly spliced BRAF(V600E). Nature 480, 387–390 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Yao, Z. et al. BRAF mutants evade ERK-dependent feedback by different mechanisms that determine their sensitivity to pharmacologic inhibition. Cancer Cell 28, 370–383 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Lito, P. et al. Relief of profound feedback inhibition of mitogenic signaling by RAF inhibitors attenuates their activity in BRAFV600E melanomas. Cancer Cell 22, 668–682 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Corbalan-Garcia, S., Yang, S. S., Degenhardt, K. R. & Bar-Sagi, D. Identification of the mitogen-activated protein kinase phosphorylation sites on human Sos1 that regulate interaction with Grb2. Mol. Cell. Biol. 16, 5674–5682 (1996).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Kamioka, Y., Yasuda, S., Fujita, Y., Aoki, K. & Matsuda, M. Multiple decisive phosphorylation sites for the negative feedback regulation of SOS1 via ERK. J. Biol. Chem. 285, 33540–33548 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Nissan, M. H. et al. Loss of NF1 in cutaneous melanoma is associated with RAS activation and MEK dependence. Cancer Res. 74, 2340–2350 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Shapira, S., Barkan, B., Friedman, E., Kloog, Y. & Stein, R. The tumor suppressor neurofibromin confers sensitivity to apoptosis by Ras-dependent and Ras-independent pathways. Cell Death Differ. 14, 895–906 (2007).

    PubMed  CAS  Google Scholar 

  26. Hunter, J. C. et al. Biochemical and structural analysis of common cancer-associated KRAS mutations. Mol. Cancer Res. 13, 1325–1335 (2015).

    Article  PubMed  CAS  Google Scholar 

  27. Watson, I. R. et al. The RAC1 P29S hotspot mutation in melanoma confers resistance to pharmacological inhibition of RAF. Cancer Res. 74, 4845–4852 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Kobayashi, T. et al. Molecular and clinical analysis of RAF1 in Noonan syndrome and related disorders: dephosphorylation of serine 259 as the essential mechanism for mutant activation. Hum. Mutat. 31, 284–294 (2010).

    Article  PubMed  CAS  Google Scholar 

  29. Fujita-Sato, S. et al. Enhanced MET translation and signaling sustains K-Ras-driven proliferation under anchorage-independent growth conditions. Cancer Res. 75, 2851–2862 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Foster, R., Mudd, C., Wiggins, C. & Torrance, C. 3-Dimensional growth reveals KRAS dependency. Cancer Res. 73, 4289 (2013).

    Article  Google Scholar 

  31. Vartanian, S. et al. Identification of mutant K-Ras-dependent phenotypes using a panel of isogenic cell lines. J. Biol. Chem. 288, 2403–2413 (2013).

    Article  PubMed  CAS  Google Scholar 

  32. Huang, H. et al. Oncogenic K-Ras requires activation for enhanced activity. Oncogene 33, 532–535 (2014).

    Article  PubMed  CAS  Google Scholar 

  33. Tashiro, T. et al. In vivo and ex vivo cetuximab sensitivity assay using three-dimensional primary culture system to stratify KRAS mutant colorectal cancer. PLoS One 12, e0174151 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. McDonald, E. R.III et al. Project DRIVE: a compendium of cancer dependencies and synthetic lethal relationships uncovered by large-scale, deep RNAi screening. Cell 170, 577–592 (2017).

    Article  PubMed  CAS  Google Scholar 

  35. McDonald, E. R.III et al. Project DRIVE: a compendium of cancer dependencies and synthetic lethal relationships uncovered by large-scale, deep RNAi screening. Cell 170, 577–592.e10 (2017).

    Article  PubMed  CAS  Google Scholar 

  36. Drosten, M. et al. Genetic analysis of Ras signalling pathways in cell proliferation, migration and survival. EMBO J. 29, 1091–1104 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Dance, M., Montagner, A., Salles, J. P., Yart, A. & Raynal, P. The molecular functions of Shp2 in the Ras/Mitogen-activated protein kinase (ERK1/2) pathway. Cell Signal. 20, 453–459 (2008).

    Article  PubMed  CAS  Google Scholar 

  38. Hadari, Y. R., Kouhara, H., Lax, I. & Schlessinger, J. Binding of Shp2 tyrosine phosphatase to FRS2 is essential for fibroblast growth factor-induced PC12 cell differentiation. Mol. Cell. Biol. 18, 3966–3973 (1998).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Aronheim, A. et al. Membrane targeting of the nucleotide exchange factor Sos is sufficient for activating the Ras signaling pathway. Cell 78, 949–961 (1994).

    Article  PubMed  CAS  Google Scholar 

  40. Cunnick, J. M., Dorsey, J. F., Munoz-Antonia, T., Mei, L. & Wu, J. Requirement of SHP2 binding to Grb2-associated binder-1 for mitogen-activated protein kinase activation in response to lysophosphatidic acid and epidermal growth factor. J. Biol. Chem. 275, 13842–13848 (2000).

    Article  PubMed  CAS  Google Scholar 

  41. Cunnick, J. M. et al. Regulation of the mitogen-activated protein kinase signaling pathway by SHP2. J. Biol. Chem. 277, 9498–9504 (2002).

    Article  PubMed  CAS  Google Scholar 

  42. Batth, T. S. et al. Large-scale phosphoproteomics reveals Shp-2 phosphatase-dependent regulators of Pdgf receptor signaling. Cell Rep. 22, 2784–2796 (2018).

    Article  PubMed  CAS  Google Scholar 

  43. Niihori, T. et al. Germline KRAS and BRAF mutations in cardio-facio-cutaneous syndrome. Nat. Genet. 38, 294–296 (2006).

    Article  PubMed  CAS  Google Scholar 

  44. Janes, M. R. et al. Targeting KRAS mutant cancers with a covalent G12C-specific inhibitor. Cell 172, 578–589e.17 (2018).

    Article  PubMed  CAS  Google Scholar 

  45. Zehir, A. et al. Mutational landscape of metastatic cancer revealed from prospective clinical sequencing of 10,000 patients. Nat. Med. 23, 703–713 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Koltun, E. & Revolution Medicines, Inc. (WO2018013597) 2,5-Disubstituted 3-methyl pyrazines and 2,5,6-trisubstituted 3-methyl pyrazines as allosteric SHP2 inhibitors. Patent no. PCT/US2017/041577 (2017); https://go.nature.com/2mv9pBw

  47. Lee, S. J. et al. Regulation of hypoxia-inducible factor 1alpha (HIF-1alpha) by lysophosphatidic acid is dependent on interplay between p53 and Kruppel-like factor 5. J. Biol. Chem. 288, 25244–25253 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. National Cancer Institute. The NCI RAS initiative. National Institutes of Health https://www.cancer.gov/research/key-initiatives/ras (2018).

  49. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

Revolution Medicines thanks S. Kelsey for scientific guidance and critical review of the manuscript, K. Shokat for scientific guidance, N. Tibrewal for development of the phosphatase selectivity panel and J. Jasper for scientific input. F. Haderk was supported by the Mildred Scheel postdoctoral fellowship from the German Cancer Aid. Funding support from NIH/NCI U54CA224081, NIH/NCI R01CA204302, NIH/NCI R01CA211052, NIH/NCI R01CA169338 and Pew Trust and Stewart Foundation (T.G.B.) is acknowledged. We are grateful to researchers at MI Bioresearch, WuXi Apptec and Genendesign who support our in vivo pharmacology efforts.

Author information

Authors and Affiliations

Authors

Contributions

R.J.N. contributed to the design, conduct and interpretation of all experiments. F.H. contributed in vitro and in vivo experimental design, conduct and interpretation. C.S., C.J.S, G.H., D.W., C.T., K.M., A.M., J.R., T.H. and A.Z. contributed in vitro experimental design, conduct and interpretation. V.O., C.M. and Z.W. contributed design, conduct and interpretation of in vivo experiments. C.M.B. established and maintains the TH77 model. G.K., E.S.K. and A.L.G. directed the design and synthesis of RMC-4550. M.S. contributed design, conduct and interpretation of in vivo experiments. M.A.G. contributed to the interpretation of all experiments. J.A.M.S. and T.G.B. supervised the project and contributed to the design and interpretation of all experiments. R.J.N., J.A.M.S. and T.G.B. wrote the manuscript with input from all co-authors.

Corresponding authors

Correspondence to Jacqueline A. M. Smith or Trever G. Bivona.

Ethics declarations

Competing interests

T.G.B. is a consultant to Novartis, AstraZeneca, Array Biopharma (Scientific Advisory Board) and Inventiva, and is the recipient of research grants from Ignyta and Revolution Medicines. T.G.B. is a member of the Revolution Medicines Scientific Advisory Board. R.J.N., C.S., C.J.S., D.W., C.T., K.M., A.M., J.R., T.H., Z.W., G.K., E.S.K., A.L.G., M.S., M.A.G. and J.A.M.S. are employees of Revolution Medicines, Inc. Revolution Medicines holds the patent mentioned in the Methods section. The other authors declare no competing interests.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 Western blot validation of a RAS-GTP ELISA.

NCI-H358 cells were grown in 2D culture and incubated with increasing concentrations of RMC-4450 for one hour. a) Cellular lysates were prepared and levels of RAS-GTP were measured by ELISA. Shown are three independent biological replicates performed in technical duplicate. b) Assessment of RAS-GTP in samples from replicate 2 by western blot. RAS-GTP was affinity purified from the cellular lysates using the GST-RAF-RBD bound to glutathione resin. Source data is provided in Supplementary Table 9.

Supplementary Figure 2 Expression of SHP2E76K rescues Class 3 BRAF, NF1LOF and KRASG12C mutant cell lines from RMC-4550.

Stable expression of SHP2-WT and SHP2-E76K was induced upon lentiviral transfection of NCI-H508 (Class 3 BRAF), NCI-H1838 (NF1LOF), and NCI-H358 (KRASG12C) cells (n = 3 biologically independent experiments; figures show mean +/− S.D.) a, c, e) Effect of RMC-4550 on proliferation of NCI-H508 (a), NCI-H1838 cells (c), and NCI-H358 cells (e) in 3D culture. Three days after seeding, cells were treated with RMC-4550 and cell viability measured on Day 8 using CTG (data are representative of n = 2 independent experiments for NCI-H508 and NCI-H1838, and n = 3 observations for H358, in technical duplicate; figures show mean +/− S.D.). RMC-4550 exhibited concentration-dependent inhibition of growth in parental and SHP2WT-expressing cell lines representing all three genotypes but failed to elicit a reduction in proliferation ≥50% of control signal in SHP2E76K expressing cells up to the maximal test concentration of 10 μM. b, d, f) Effect of RMC-4550 on pERK levels in NCI-H508 (b), NCI-H1838 cells (d), and NCI-H358 cells (f). Cells were grown in 2D culture and incubated with 1 µM RMC-4550 for 90 minutes. Cellular lysates were prepared and levels of pERK determined by immunoblot. RMC-4550 treatment resulted in reduction of pERK levels in parental and SHP2WT expressing cells across all genotypes, but failed to suppress pERK in SHP2E76K expressing cells. Shown is one representative example of three independent experiments. Source data is provided in Supplementary Table 9.

Supplementary Figure 3 Effects of SHP2 inhibition in RAS pathway-driven cancer cells.

(a) Effect of RMC-4550 on RAS-GTP levels in BEAS-2B cells stably transfected with scrambled (shScramble) or NF1 targeting (shNF1) shRNA. Data shows a representative of three independent experiments. (b-d) (b) Cell sensitivities to the knockdowns of the indicated genes were determined using crystal violet staining assays (n = 2). The spectrophotometric intensities of the crystal violet stains were quantified and represented as a bar plot. (c) RAS-GTP levels in cellular lysates are shown by GST-RAF-RBD affinity capture. (d) Individual western blot band intensities were quantified and shown for pERK/ERK, RAS-GTP/RAS (total) level and knock down efficiency, expressed relative to a control scrambled shRNA intensity of 1.0 and the effect of knockdown on pERK. (e-h) (e), (f) Effect of RMC-4550 and the KRASG12C-specific inhibitor ARS-853 on pERK levels in KRASG12C cancer cell lines. Cells were incubated with increasing concentrations of RMC-4550 or ARS-853 for one hour. Cellular lysates were prepared and levels of pERK determined. RMC-4550 produced a concentration-dependent reduction in pERK levels. Geometric mean IC50 values for reduction in pERK by RMC-4550 in NCI-H358 and CALU-1 were 36 nM and 7 nM, respectively. A concentration-dependent reduction in pERK levels was observed for ARS-853 (geometric mean IC50 values of 4700 nM and 430 nM for NCI-H358 and CALU-1 cells, respectively); data represent n = 4 independent experiments, performed in technical duplicate; figures show mean +/− S.D.). (g), (h) Effect of RMC-4550 and ARS-853 on proliferation of two KRASG12C cancer cell lines in 3D culture. NCI-H358 and CALU-1 cells were treated with RMC-4550 or ARS-853 and cell viability measured on Day 7 using Cell Titer Glo (data represent n = 4 independent experiments, performed in duplicate for NCI-H358, and n = 3 independent experiments, performed in duplicate for CALU-1; figures show mean +/− S.D.). RMC-4550 and ARS-853 exhibited geometric mean IC50 values of 43 nM and 1020 nM, respectively, in NCI-H358 cells. Source data is provided in Supplementary Table 9.

Supplementary Figure 4 Effects of SHP2 inhibition in KRAS-driven cancer cells.

(a) Comparison of the effect of SHP2 inhibition and KRAS knockdown on the proliferation of KRAS-mutant cancer cell lines bearing different oncogenic mutations in KRAS as indicated by color coding. The negative logarithm of RMC-4550 GI50 values for cancer cells bearing activating KRAS mutations (described in Supplementary Table 5) were plotted against the negative value of their Project Drive Sensitivity Score for KRAS shRNA knockdown. Cells with high sensitivity to KRAS knockdown have larger negative Sensitivity Scores, and cells with greater sensitivity to SHP2 inhibition have larger negative log(RMC-4550 GI50) values. (b-c) Effect of RMC-4550 on RAS-GTP and pERK levels in Hras−/−, Nras−/−, Kras−/− mouse embryonic fibroblasts (MEFs) expressing solely the indicated RAS or BRAF transcript. Cells were grown in 2D culture and incubated with 1 µM RMC-4550 for 24 h. b) Cellular lysates were prepared and levels of RAS-GTP and pERK determined by immunoblot. RAS-GTP was affinity purified from the cellular lysates using GST-tagged RAF-RBD coupled to glutathione agarose beads. (c) RAS-GTP levels were quantified in whole cell lysates via RAS-GTP ELISA. Significant reduction in RAS-GTP and pERK levels is observed for KRAS WT (WT-4B) and KRAS G12 expressing MEFs, while cells expressing KRASQ61H lack response. Values above bars represent fraction of RAS-GTP in RMC-4550 vs. DMSO treated cells. Data in (b) shows a representative example of three independent experiments; data in (c) shows a representative example of two independent experiments, and bars indicates mean of two technical replicates. (d-e) (d) Expression of the wild type SOS1 and SOS-F proteins was induced by doxycycline treatment for 24 h. Cells were treated with 50 ng/mL EGF at various time intervals, lysed, and total pERK levels were measured by AlphaLisa. The SOS-F construct showed increased and sustained levels of pERK signal in the absence of EGF stimulation. Shown is one representative example performed in technical duplicate of three independent experiments. Figure indicates mean +/− S.D. (e) NCI-H358 cells transiently expressing SOS1 and SOS-F encoding cDNA for 48 h were treated with vehicle or 1 µM RMC-4550 for 1 and 6 h. Harvested lysates from the cells were probed for the indicated proteins using western blot. Data shown represent at least 3 independent biological replicates. Source data is provided in Supplementary Table 9.

Supplementary Figure 5 Xenograft models of KRASG13D and KRASQ61K are insensitive to RMC-4550.

Daily oral administration of RMC-4550 (30 mg/kg) to athymic nude mice bearing (a) HCT-116 (KRASG13D) or (b) CALU-6 (KRASQ61K) xenograft tumors did not produce significant tumor growth inhibition. Treatment with RMC-4550 was well tolerated in both models, as assessed by body weight (c,d). Source data is provided in Supplementary Table 9.

Supplementary Figure 6 Tolerability assessment in animals treated with the SHP2 inhibitor.

Repeated daily oral administration of RMC-4550 of all dose groups in the MeWo (a), NCI-H1838 (b), NCI-H358 (c), and MIA PaCa-2 (d), xenograft models and LUN023 (e), LUN037 (f), TH77 (g), LUN092 (h), and LUN156 (i) PDX models was well tolerated as assessed by body weight. Source data is provided in Supplementary Table 9.

Supplementary Figure 7 Durable response to RMC-4550 in NF1LOF model.

(a) Western blot demonstrates loss of NF1 protein in TH77 PDX model. A previously published EGFRL858R, NF1wt PDX model is shown as comparator1. (b) Tumor specimen of vehicle and RMC-4550 treated groups were analyzed regarding their general tissue morphology by H&E stain (representative of n = 3 independent samples per treatment arm, 10X magnification, size bar: 100 nm). Significant areas of necrosis (arrows) were present in tumors of RMC-4550 treated animals.

Supplementary Figure 8 Unprocessed scans of raw blots.

(a) SHP2 IP from Fig. 5c, (b) GAB1 IP from Fig. 5d, and (c) membrane fractionation from Fig. 5e. (d) Effect of RMC-4550 on RAS-GTP levels from Supplementary Figure 1b, (e) Effect of SHP2 variants on pERK in H508 cells from Supplementary Figure 2b, (f) Effect of SHP2 variants on pERK in H1838 cells from Supplementary Figure 2d, (g) Effect of SHP2 variants on pERK in H358 cells from Supplementary Figure 2f. (h) Effect of NF1 shRNA knockdown on RAS-GTP levels from Supplementary Figure 3a, and (i) RAS-GTP levels from Supplementary Figure 3c, (j) Effect of SHP2 and SOS1 shRNA knockdown in RAS-GTP levels from Supplementary Figure 3d, (k) Effect of RMC-4550 on RAS-GTP and pERK levels in RAS-less MEFs from Supplementary Figure 4b, (l) Effect of SOS or SOS-F expression on pERK from Supplementary Figure 4e, (m) NF1 levels in TH77 PDX model from Supplementary Figure 7a. Note that panels (d) through (e) are located in additional pages.

Supplementary information

Supplementary Information

Supplementary Figures 1–9 and Supplementary Table legends.

Reporting Summary

Supplementary Table 1

RMC-4550 activity against a panel of 15 phosphatases.

Supplementary Table 2

RMC-4550 activity against a panel of 468 kinases at a test concentration of 10 µM.

Supplementary Table 3

RMC-4550 activity against a panel of 44 safety targets at a test concentration of 10 μM.

Supplementary Table 4

RMC-4550 activity against a panel of NF1 mutant cell lines.

Supplementary Table 5

RMC-4550 activity against a panel of 33 KRAS mutant cell lines.

Supplementary Table 6

Pharmacokinetic parameters of RMC-4550 from single dose NCI-H358 study.

Supplementary Table 7

Pharmacokinetic parameters of RMC-4550 from end of MIA PaCa-2 efficacy study.

Supplementary Table 8

Antibodies.

Supplementary Table 9

Statistics source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nichols, R.J., Haderk, F., Stahlhut, C. et al. RAS nucleotide cycling underlies the SHP2 phosphatase dependence of mutant BRAF-, NF1- and RAS-driven cancers. Nat Cell Biol 20, 1064–1073 (2018). https://doi.org/10.1038/s41556-018-0169-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-018-0169-1

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