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.

  • Letter
  • Published:

MEK drives BRAF activation through allosteric control of KSR proteins

Abstract

RAF family kinases have prominent roles in cancer1. Their activation is dependent on dimerization of their kinase domains, which has emerged as a hindrance for drug development2,3. In mammals, RAF family kinases include three catalytically competent enzymes (ARAF, BRAF and CRAF) and two pseudokinases (KSR1 and KSR2) that have been described as scaffolds owing to their apparent ability to bridge RAF isoforms and their substrate, mitogen-activated protein kinase kinase (MEK)4. Kinase suppressor of Ras (KSR) pseudokinases were also shown to dimerize with kinase-competent RAFs to stimulate catalysis allosterically5. Although GTP-bound RAS can modulate the dimerization of RAF isoforms by engaging their RAS-binding domains, KSR1 and KSR2 lack an RAS-binding domain and therefore the regulatory principles underlying their dimerization with other RAF family members remain unknown. Here we show that the selective heterodimerization of BRAF with KSR1 is specified by direct contacts between the amino-terminal regulatory regions of each protein, comprising in part a novel domain called BRS in BRAF and the coiled-coil-sterile α motif (CC-SAM) domain in KSR1. We also discovered that MEK binding to the kinase domain of KSR1 asymmetrically drives BRAF–KSR1 heterodimerization, resulting in the concomitant stimulation of BRAF catalytic activity towards free MEK molecules. These findings demonstrate that KSR–MEK complexes allosterically activate BRAF through the action of N-terminal regulatory region and kinase domain contacts and challenge the accepted role of KSR as a scaffold for MEK recruitment to RAF.

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

Figure 1: MEK1 asymmetrically stimulates BRAF–KSR1 dimerization by interacting with KSR1.
Figure 2: KSR1-activated BRAF phosphorylates KSR1-free MEK1.
Figure 3: The NTRs of KSR1 and BRAF selectively interact.
Figure 4: MEK releases the intramolecular interaction between the NTR and kinase domain of KSR1.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

References

  1. Desideri, E., Cavallo, A. L. & Baccarini, M. Alike but different: RAF paralogs and their signaling outputs. Cell 161, 967–970 (2015)

    Article  CAS  Google Scholar 

  2. Karoulia, Z., Gavathiotis, E. & Poulikakos, P. I. New perspectives for targeting RAF kinase in human cancer. Nat. Rev. Cancer 17, 676–691 (2017)

    Article  CAS  Google Scholar 

  3. Lavoie, H. & Therrien, M. Regulation of RAF protein kinases in ERK signalling. Nat. Rev. Mol. Cell Biol. 16, 281–298 (2015)

    Article  CAS  Google Scholar 

  4. Kolch, W. Coordinating ERK/MAPK signalling through scaffolds and inhibitors. Nat. Rev. Mol. Cell Biol. 6, 827–837 (2005)

    Article  CAS  Google Scholar 

  5. Rajakulendran, T., Sahmi, M., Lefrançois, M., Sicheri, F. & Therrien, M. A dimerization-dependent mechanism drives RAF catalytic activation. Nature 461, 542–545 (2009)

    Article  CAS  ADS  Google Scholar 

  6. Lavoie, H. et al. Inhibitors that stabilize a closed RAF kinase domain conformation induce dimerization. Nat. Chem. Biol. 9, 428–436 (2013)

    Article  CAS  Google Scholar 

  7. Brennan, D. F. et al. A Raf-induced allosteric transition of KSR stimulates phosphorylation of MEK. Nature 472, 366–369 (2011)

    Article  CAS  ADS  Google Scholar 

  8. Haling, J. R. et al. Structure of the BRAF-MEK complex reveals a kinase activity independent role for BRAF in MAPK signaling. Cancer Cell 26, 402–413 (2014)

    Article  CAS  Google Scholar 

  9. Hu, J. et al. Mutation that blocks ATP binding creates a pseudokinase stabilizing the scaffolding function of kinase suppressor of Ras, CRAF and BRAF. Proc Natl Acad Sci USA 108, 6067–6072 (2011)

    Article  CAS  ADS  Google Scholar 

  10. Koveal, D. et al. A CC-SAM, for coiled coil-sterile α motif, domain targets the scaffold KSR-1 to specific sites in the plasma membrane. Sci. Signal. 5, ra94 (2012)

    Article  Google Scholar 

  11. Van Aelst, L., Barr, M., Marcus, S., Polverino, A. & Wigler, M. Complex formation between RAS and RAF and other protein kinases. Proc. Natl Acad. Sci. USA 90, 6213–6217 (1993)

    Article  CAS  ADS  Google Scholar 

  12. Vojtek, A. B., Hollenberg, S. M. & Cooper, J. A. Mammalian Ras interacts directly with the serine/threonine kinase Raf. Cell 74, 205–214 (1993)

    Article  CAS  Google Scholar 

  13. Zhang, X. F. et al. Normal and oncogenic p21ras proteins bind to the amino-terminal regulatory domain of c-Raf-1. Nature 364, 308–313 (1993)

    Article  CAS  ADS  Google Scholar 

  14. Jin, T. et al. RAF inhibitors promote RAS-RAF interaction by allosterically disrupting RAF autoinhibition. Nat. Commun. 8, 1211 (2017)

    Article  ADS  Google Scholar 

  15. Ritt, D. A., Monson, D. M., Specht, S. I. & Morrison, D. K. Impact of feedback phosphorylation and Raf heterodimerization on normal and mutant B-Raf signaling. Mol. Cell. Biol. 30, 806–819 (2010)

    Article  CAS  Google Scholar 

  16. Johannessen, C. M. et al. COT drives resistance to RAF inhibition through MAP kinase pathway reactivation. Nature 468, 968–972 (2010)

    Article  CAS  ADS  Google Scholar 

  17. Moffat, J. et al. A lentiviral RNAi library for human and mouse genes applied to an arrayed viral high-content screen. Cell 124, 1283–1298 (2006)

    Article  CAS  Google Scholar 

  18. Boussif, O. et al. A versatile vector for gene and oligonucleotide transfer into cells in culture and in vivo: polyethylenimine. Proc. Natl Acad. Sci. USA 92, 7297–7301 (1995)

    Article  CAS  ADS  Google Scholar 

  19. Thevakumaran, N. et al. Crystal structure of a BRAF kinase domain monomer explains basis for allosteric regulation. Nat. Struct. Mol. Biol. 22, 37–43 (2015)

    Article  CAS  Google Scholar 

  20. James, P., Halladay, J. & Craig, E. A. Genomic libraries and a host strain designed for highly efficient two-hybrid selection in yeast. Genetics 144, 1425–1436 (1996)

    CAS  PubMed  PubMed Central  Google Scholar 

  21. James, P. Yeast two-hybrid vectors and strains. Methods Mol. Biol. 177, 41–84 (2001)

    CAS  PubMed  Google Scholar 

  22. Spee, J. H., de Vos, W. M. & Kuipers, O. P. Efficient random mutagenesis method with adjustable mutation frequency by use of PCR and dITP. Nucleic Acids Res. 21, 777–778 (1993)

    Article  CAS  Google Scholar 

  23. Winn, M. D. et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D 67, 235–242 (2011)

    Article  CAS  Google Scholar 

  24. Sammito, M. et al. ARCIMBOLDO_LITE: single-workstation implementation and use. Acta Crystallogr. D 71, 1921–1930 (2015)

    Article  CAS  Google Scholar 

  25. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D 60, 2126–2132 (2004)

    Article  Google Scholar 

  26. Adams, P. D . et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D 66, 213–221 (2010)

    Article  CAS  Google Scholar 

  27. Kabsch, W. Xds. Acta Crystallogr. D 66, 125–132 (2010)

    Article  CAS  Google Scholar 

  28. Pape, T. & Schneider, T. R. HKL2MAP: a graphical user interface for macromolecular phasing with SHELX programs. J. Appl. Crystallogr. 37, 843–844 (2004)

    Article  CAS  Google Scholar 

  29. Sheldrick, G. M. Experimental phasing with SHELXC/D/E: combining chain tracing with density modification. Acta Crystallogr. D 66, 479–485 (2010)

    Article  CAS  Google Scholar 

  30. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007)

    Article  CAS  Google Scholar 

  31. Langer, G., Cohen, S. X., Lamzin, V. S. & Perrakis, A. Automated macromolecular model building for X-ray crystallography using ARP/wARP version 7. Nat. Protocols 3, 1171–1179 (2008)

    Article  CAS  Google Scholar 

  32. Chen, V. B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D 66, 12–21 (2010)

    Article  CAS  Google Scholar 

  33. Delaglio, F. et al. NMRPipe: a multidimensional spectral processing system based on UNIX pipes. J. Biomol. NMR 6, 277–293 (1995)

    Article  CAS  Google Scholar 

  34. Vranken, W. F. et al. The CCPN data model for NMR spectroscopy: development of a software pipeline. Proteins 59, 687–696 (2005)

    Article  CAS  Google Scholar 

  35. Fielding, L. NMR methods for the determination of protein-ligand dissociation constants. Curr. Top. Med. Chem. 3, 39–53 (2003)

    Article  MathSciNet  CAS  Google Scholar 

  36. Schrodinger, LLC. The PyMOL Molecular Graphics System, Version 1.8 (2015)

  37. Pettersen, E. F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004)

    Article  CAS  Google Scholar 

  38. Waterhouse, A. M., Procter, J. B., Martin, D. M., Clamp, M. & Barton, G. J. Jalview Version 2—a multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–1191 (2009)

    Article  CAS  Google Scholar 

  39. Bond, C. S. & Schüttelkopf, A. W. ALINE: a WYSIWYG protein-sequence alignment editor for publication-quality alignments. Acta Crystallogr. D 65, 510–512 (2009)

    Article  CAS  Google Scholar 

  40. Schumann, F. H. et al. Combined chemical shift changes and amino acid specific chemical shift mapping of protein-protein interactions. J. Biomol. NMR 39, 275–289 (2007)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank L. Volpon, M. Osborne, G. Seabrook, and the Institute for Research in Immunology and Cancer High-Throughput Screening and Proteomics facilities for experimental assistance. H.L. was supported by a Canadian Institutes for Health Research (CIHR) Banting fellowship. N.T. was supported by a CIHR scholarship. Research was supported by Impact Grants from the Canadian Cancer Society (702319 to M.T. and 704116 to F.S.) and by CIHR operating grants (MOP119443 and PJT152873 to M.T. and FDN143277 to F.S.). M.T. and F.S. hold Canada Research Chairs in Intracellular Signaling and in Structural Biology of Signaling. Synchrotron work at the Northeastern Collaborative Access Team beamlines was supported by National Institutes of Health grants (P41 GM103403 and S10 RR029205).

Author information

Authors and Affiliations

Authors

Contributions

H.L., M.S., P.M., S.A.M., F.S., and M.T. designed the experiments. H.L., P.M., F.S., and M.T. wrote the manuscript. H.L. performed BRET assays with T.J. M.S. performed co-IP and transactivation assays. H.L. and S.A.M. conducted Y2H experiments. P.M., N.T., and I.K. performed structural and biophysical experiments.

Corresponding authors

Correspondence to Frank Sicheri or Marc Therrien.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks D. Barford, S. Hubbard and the other anonymous reviewer(s) for their contribution to the peer review of this work.

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 Figure 1 MEK levels modulate the formation of BRAF–KSR1 dimers.

a, MEK1 and MEK2 expression selectively modulates BRAF–KSR1 interaction but not the KRASG12V–BRAFRBD counter screen interaction by BRET. MEK1 and MEK2 expression was driven by the lentiviral vector pLX301. Error bars correspond to mean values ± s.d. of biological triplicates. b, MEK1 and MEK2 modulate BRAF–KSR1 interaction by BRET. c, MEK1 stimulates BRAF–KSR1 interaction in a dose-dependent manner. mCherry-tagged MEK1 was titrated in cells expressing the BRAF–KSR1 BRET biosensors. d, MEK1 promotes full-length BRAF–KSR1 interaction by co-IP. e, Reducing endogenous MEK1 and/or MEK2 levels by RNA interference (shMEK1 and/or shMEK2) decreases BRAF–KSR1 interaction by BRET. f, Efficiency of MEK1 and MEK2 shRNA knockdowns in d. g, MEK1 stimulates BRAF–KSR1 dimerization independently of its kinase activity. mCherry-tagged MEK1 or MEK1K97R similarly stimulate BRAF–KSR1 interaction by BRET. Cells expressing the BRAF–KSR1 BRET biosensor and increasing amounts of WT MEK1 were treated with the MEK inhibitor AZD8330 (10 μM). h, MEK1 stimulates BRAF–KSR1 interaction independently of the phosphorylation status of Serine 218 and 222. mCherry-tagged MEK1, MEK1S218A–S222A or MEK1S218D–S222D similarly stimulate BRAF–KSR1 interaction by BRET. i, and j, mCherry-tagged MEK1 selectively modulates KSR-containing dimers. The impact of MEK was measured on a series of ten biosensors comprising BRAF (h) or CRAF (i) as donor probes expressed with ARAF, BRAF, CRAF, KSR1, or KSR2 as acceptor probes. k, mCherry-tagged kinase-dead MEK1K97R selectively modulates KSR-containing dimers. The impact of MEK1K97R was measured on a series of five biosensors comprising BRAF as the donor probe expressed with ARAF, BRAF, CRAF, KSR1, or KSR2 as acceptor probes. l, mCherry-tagged interaction-defective MEK1F311S does not modulate dimerization of five biosensors comprising BRAF as the donor probe expressed with ARAF, BRAF, CRAF, KSR1, or KSR2 as acceptor probes. In c and gl, BRET log2-transformed fold-changes were reported as a function of log10-transformed mCherry relative fluorescence units. Experiments were repeated at least three times. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 2 Identification of mutations disrupting KSR–MEK and RAF–MEK interactions.

a, Crystal structure of KSR2 bound to MEK1. The structure shows a heterotetrameric arrangement with two KSR2 protomers assembling as a side-to-side dimer. Each KSR2 protomer in turn interacts in a face-to-face manner with one molecule of MEK1. Zoom-in views (dashed boxes) highlight the involvement of the αG helices and the activation segments (AS) of both kinase domains in the face-to-face contact (Protein Data Bank accession number 2Y4I). b, Screening scheme used to identify MEK1 and KSR1 binding-defective mutants by error-prone-PCR random mutagenesis using the Y2H system. c, Yeast spotting assay showing that the F311S mutation in MEK1 helix αG abrogates interaction with KSR1, BRAF, and CRAF. d, Yeast spotting assay showing that KSR1 and BRAF helix αG mutations (W831R and I666R, respectively) disrupt their interaction with MEK1. Experiments in c and d were repeated at least three times. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 3 Mutations in helix αG of MEK1 and KSR1 (F311S and W831R, respectively) do not perturb protein fold but abrogate stoichiometric assembly of the BRAF–KSR1–MEK1 complex.

a, SEC–MALS analysis of MEK1 WT and F311S mutant. The horizontal black and blue lines correspond to SEC–MALS calculated masses for the indicated proteins. A summary table of measured and calculated molecular masses for each protein is shown below. The experiment was performed in triplicate. b, Fluorescence polarization (FP) binding analysis of MEK1 WT and F311S mutant to a 25 nM fluorophore-conjugated ATP tracer (BODIPY–ATP). Error bars correspond to mean values ± s.d. of technical triplicates. Indicated Kd values ± s.d. are representative of two independent experiments. c, MEK1 and MEK1 F311S expressed in HEK293T cells show equal ability to bind the ATP analogue desthiobiotin-ATP as assessed using a Pierce Kinase Enrichment Kit (Thermo-Fisher Scientific). d, e, WT KSR1 and the W831R mutant assemble to the same extent with the known interaction partners HSP90, CDC37, and 14-3-3, indicating that the protein fold of KSR1 is not overtly perturbed by the W831R mutation. Interaction partners were confirmed by mass spectrometry and by western immunoblotting. f, WT KSR1 and the W831R variant expressed in HEK293T cells show equal binding to the ATP analogue desthiobiotin-ATP as assessed using a Pierce Kinase Enrichment Kit (Thermo-Fisher Scientific). g, BRAF–KSR1–MEK1 complexes were purified to near-homogeneity by immunoaffinity purification on anti-Flag–agarose resin after treatment of the transfected cells with GDC-0879 (10 μM). BRAF and KSR1 form near-stoichiometric complexes in the presence of MEK1 overexpression (compare lanes 3 and 4). The KSR1 W831R mutant (lane 5) and the MEK1 F311S mutant (lane 7) but not the BRAF I666R mutant (lane 6) perturbed this near-stoichiometric interaction. Note that the BRAF–KSR1–MEK1 complex also associates at near-stoichiometric levels with endogenous 14-3-3 isoforms. In addition, in the absence of MEK1 overexpression, appreciable but sub-stoichiometric levels of endogenous MEK are detected by immunoblotting. Protein identity was determined by mass spectrometry. Asterisks indicate the position of non-specific proteins associating with the anti-Flag–agarose resin. h, The basal association of BRAF–KSR1 dimers (top panel, first lane) depends on endogenous MEK. Reducing MEK1 and MEK2 levels by short hairpin RNA (shRNA) interference (combination of shMEK1 and shMEK2) decreases the amount of Pyo-KSR1 co-purifying with Flag–BRAF. NT, non-target control shRNA. Experiments were repeated at least three times. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 4 MEK1 binding to KSR1 is required for the transactivation of BRAF by side-to-side dimerization.

a, BRAF activation by KSR1 depends on an intact side-to-side dimerization surface in each protomer. Co-transfection of WT BRAF and KSR1 induces a strong pMEK signal while KSR1R665H or BRAFR509H mutants do not. b, Human KSR1 and KSR2 can both transactivate BRAF. C-terminal GFP10 fusions of KSR1 and KSR2 were co-transfected with BRAF and the resulting pMEK signal was monitored by western immunoblotting. c, Disruption of KSR1–MEK1 interaction with the W831R mutation perturbs BRAF transactivation by KSR1. The BRAF I666R mutation also prevents MEK phosphorylation. Similarly, MEK1F311S is not phosphorylated when co-transfected with WT BRAF and KSR1. d, Transactivation of Drosophila RAF (dRAF) by Drosophila KSR (dKSR) requires MEK binding (W896R mutation is homologous to W831R in human KSR1) and dRAF–dKSR dimerization (R732H mutation is homologous to R665H in human KSR1). e, The transactivation potential of FRB–dKSRKD towards FKBP–dRAFKD was assessed by monitoring the levels of phosphorylated MEK in the presence or absence of rapamycin in S2 cells (1 μM). Blocking MEK binding to dKSR (W896R mutation) and dRAF–dKSR dimerization (R732H mutation) abrogated RAF transactivation and MEK phosphorylation as judged by western immunoblotting. f, MEK phosphorylation induced by co-transfection of BRAF and KSR1 depends on the catalytic activity of BRAF and not on the integrity of the active site region of KSR1. g, KSR1 A637F (AF) and A637F/YLQE602-604>DDEE602-604 (AFDDEE) gain-of-function mutants are dependent on binding to endogenous MEK1 or MEK2 to dimerize with endogenous BRAF and to activate the cellular pool of RAF as judged by the pMEK1 or pMEK2 signal. The W831R substitution in KSR1 was used to abrogate KSR1AF and KSR1AFDDEE mutant binding to MEK. Experiments were repeated at least three times. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 5 Identification of MEK1 mutants rescuing the interaction with KSR1W831R.

a, Screening scheme used to identify MEK1 mutants rescuing the interaction with KSR1W831R. Screening was conducted in two rounds. In round 1, MEK1 was randomly mutagenized by error-prone-PCR, which identified 16 mutants corresponding exclusively to M219V and N221Y substitutions. In round 2, saturation mutagenesis was used to test all possible residue combinations in the region encoding amino acids 219–221. This identified 70 mutants that either complemented the Y2H interaction on SD-TLH or on both SD-TLH and SD-TLA. Sequence diversity of the recovered mutants is summarized by sequence logos shown on the right (see Supplementary Table 2 for details of mutant sequences). After further phenotypic screening, the fittest mutant on SD-TLA corresponded to MEK1M219W–A220L. b, Position of MEK1 mutations that rescue the interaction with KSR1W831R. Mutations systematically mapped to the MEK1 activation segment between Ser218 and Ser222, which are target phosphorylation sites for RAF proteins. Multiple sequence alignments of MAP2K1-7 activation segments illustrate that residues 219–221 vary within the MAP2K family, suggesting that substitutions of this sequence do not drastically affect enzyme function. The M219V substitution recovered in our screen is also found at the homologous position in MAP2K3 and MAP2K6. c, MEK1 activation segment mutations MEK1M219V, MEK1N221Y, MEK1M219V/N221Y, and MEK1M219W/A220L stimulate MEK1 binding to endogenous KSR1, BRAF, and CRAF. Co-IP of Flag-tagged MEK1 variants was performed in HEK293T cells. d, MEK1M219W–A220L but not WT MEK1 stimulates KSR1W831R–BRAF dimerization in BRET assays. mCherry-tagged MEK1 WT or MEK1M219W–A220L were titrated in cells expressing the BRAF–KSR1W831R BRET biosensors. BRET log2(fold-changes) were reported as a function of log10-transformed mCherry relative fluorescence units. e, Strategy to distinguish activator and substrate MEK in BRAF transactivation assays. Eight repeats of a G4S flexible linker (symbolized by an octagon) were added onto WT MEK1 to separate it by size from co-transfected MEK variants lacking the linker. Experimental conditions corresponding to lanes 7 and 8 of Fig. 2d, respectively, are depicted at the top and bottom of the diagram. Experiment in e was repeated at least three times. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 6 BRAF and KSR1 NTRs selectively interact.

a, MEK expression promotes BRAF–KSR1 interaction but not CRAF–KSR1 interaction. b, BRET titration experiments were used to monitor the interaction between BRAFNTR, CRAFNTR, and KSR1NTR. The indicated donor and acceptor pairs were tested for interaction in titration experiments. c, Y2H assays confirm interaction between BRAFNTR and KSR1NTR, and KSR1NTR self-association. d, Domain architecture of full-length mammalian RAF and KSR isoforms and schematic of truncations used to map the molecular determinants of BRAFNTR–KSR1NTR interaction. Within the NTRs are highlighted the RBDs (orange), cysteine-rich domain (CRD; yellow), serine/threonine-rich (S/T-rich; brown) region, proline-rich (P-rich; grey) region, BRAF-specific (BRS; blue) domain, and the CC-SAM (green) domain. Co-IP results derived from multiple experiments are indicated (yes/no) with representative examples shown in e and f. e, Co-IP of Pyo-KSR1NTR with the indicated fragments of Flag-tagged BRAF. f, Co-IP of the indicated fragments of Pyo-KSR1 with Flag-tagged BRAFNTR. Experiments were repeated at least three times. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 7 The contact surfaces between BRS and CC-SAM domains are required for the physical and functional interactions between BRAF and KSR1.

a, Phylogenetic profiling of BRAF and KSR1 domains demonstrates conservation of the BRS and CC-SAM domains throughout metazoan evolution. A concomitant loss of both domains is observed in the nematode lineage (indicated by ‘domain loss’). The conservation of each of BRAF and KSR1 domains with its human orthologue was colour-coded by BLAST E-value. RPS-BLAST was used to assess conservation of the CC-SAM domain. Conservation of the BRS was established by calculating the sum of the COILS score of amino acids located upstream of the RBD in each BRAF orthologue. b, Sequence alignment of the BRAF-specific (BRS) sequence from 14 metazoan species shows a conserved pattern of hydrophobic residues consistent with a coiled-coil region. Secondary structure elements identified in the BRS crystallographic structure are delineated above the sequence alignment. The heptad repeat (AbcDefg) typical in coiled-coil domains is shown below the alignment: positions A and D usually comprise hydrophobic amino acids while positions b, c, e, f, and g usually comprise polar or charged residues. BRS mutations used in this study are indicated by arrowheads. c, Sequence alignment of the CC-SAM sequence from 15 metazoan species. Secondary structure elements identified in the apo and BRS-bound CC-SAM structures are delineated above the sequence alignment. CC-SAM mutations used in this study are indicated by arrowheads below the alignment. Species in the phylogenetic tree and the sequence alignments are Homo sapiens, Mus musculus, Gallus gallus, Xenopus tropicalis, Danio rerio, Strongylocentrotus purpuratus, Saccoglossus kowalevskii, Lingula anatina, Lottia gigantea, Drosophila melanogaster, Anopheles gambiae, Haemonchus contortus, Dictyocaulus viviparus, Necator americanus, Caenorhabditis briggsae, Caenorhabditis brenneri, Caenorhabditis remanei, Caenorhabditis elegans, Loa loa, Brugia malayi, Bursaphelenchus xylophilus, Trichinella papuae, Trichuris trichiura, Trichuris suis, Nematostella vectensis, Amphimedon queenslandica, and Trichoplax adherens. d, Y2H mating matrix of BRAF and KSR1 NTR domains and chimaeric fusions (the dashed box indicates the interaction between isolated BRS and CC-SAM domains). e, The BRS and CC-SAM domains specifically interact as determined by BRET titration experiments. In contrast, the BRS–BRS BRET pair did not generate a saturatable BRET signal. f, Swapping of the BRS and CC-SAM domains supports the BRAF–KSR1 interaction and transactivation of BRAF by KSR1. Fusing the BRS domain to CRAF NTR stimulates its interaction with KSR1NTR by co-IP (g) and by Y2H (h). Experiments in dh were repeated at least three times. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 8 NMR characterization of the physical interaction between the BRS and CC-SAM domains.

a, Superposition of the [1H–15N]HSQC spectra of the [15N]CC-SAM domain of human KSR1 (hCC-SAM) alone (black) and [15N]hCC-SAM in the presence of the BRS domain of human BRAF (hBRS) (green) at a 1:2 molar ratio. Binding of BRS to CC-SAM induced both CSPs and peak broadening. b, Superposition of the [1H–15N]HSQC spectra of [15N]hBRS (black) and [15N]hBRS in the presence of hCC-SAM (blue) at the indicated molar concentrations. In the [15N]hBRS HSQC spectra (black), numerous peaks are in intermediate (peak broadening) timescale conformational exchange, giving rise to only 50 visible peaks out of 84 expected. In the [15N]hBRS:hCC-SAM HSQC spectra (blue), the binding of CC-SAM to BRS increases conformational exchange to a faster time regime, giving rise to 65 visible peaks. c, CSP of the mouse KSR1 [15N]CC-SAM (mCC-SAM) domain as a function of titrated hBRS. Superposition of the [1H–15N]HSQC spectra of [15N]mCC-SAM alone (black) and [15N]mCC-SAM in the presence of hBRS (green) at the indicated molar concentrations (top left). The Kd value was calculated as the mean ± s.d. of five independent peak profiles (right) for which one representative CSP profile is shown (bottom left). d, CSP of the [15N]hBRS domain as a function of titrated hCC-SAM. Superposition of the [1H–15N]HSQC spectra of [15N]hBRS alone (black) and [15N]hBRS in the presence of hBRS (blue) at the indicated molar concentrations (top left). The Kd value was calculated as the mean ± s.d. of six independent peak profiles (right) for which one representative CSP profile is shown (bottom left). e, Superposition of the [1H–15N]HSQC spectra of the [15N]mCC-SAM alone (black) and in the presence of the hBRS (green) at a 1:2 molar ratio. Binding of hBRS to mCC-SAM induced both CSPs and peak broadening. f, Peak intensity change and CSPs versus residue number for the [1H–15N]HSQC spectra of [15N]mCC-SAM alone versus [15N]mCC-SAM with hBRS at a 1:2 molar ratio. Cut-off values indicated by the red line were calculated using the corrected s.d. method40. Vertical green and grey bars indicate residues identified as being part of the interacting surface and unassigned residues, respectively. g, Significant CSPs and peak intensity changes greater than the cut-off value determined in f are highlighted in green on the surface of the mCC-SAM domain (Protein Data Bank accession number 2LPE). Most CSPs and peak intensity changes map to the coiled-coil region of mCC-SAM. Non-shifting peaks are indicated in pale grey while unassigned peaks in the HSQC are coloured in dark grey.

Extended Data Figure 9 Crystal structures of the BRS domain and the BRS:CC-SAM complex.

a, X-ray data collection and refinement statistics table for the crystal structures of the hBRS domain and the hCC–SAM–hBRS complex. b, Ribbon representation of the crystal structure of the BRS domain of human BRAF. Residues comprising the hydrophobic core of the coiled-coil interactions are shown as sticks. c, Structure of the BRAF BRS and KSR1 CC-SAM domain complex. Contents of the asymmetric unit revealed two CC-SAM:BRS complexes interacting in trans. Disordered (GS)4 linker between CC-SAM and BRS domains is represented by dashed lines. The connectivity between the BRS and the CC-SAM domains was deduced as trans because in this configuration, but not the cis configuration, the C terminus of the CC-SAM domain is sufficiently close to the N terminus of the BRS domain (15 Å) to be physically spanned by the disordered linker comprising 15 amino-acid residues. For comparison, the connection in cis would need to minimally span 80 Å. d, Representative final 2FoFc electron density map contoured at 1.0σ for the BRS:CC-SAM complex at a region of drastic conformational rearrangement in CC-SAM relative to the isolated CC-SAM structure (indicated by the arrow). e, SEC–MALS analysis of the indicated crystallization fusion proteins (top). The horizontal black, red, and green lines correspond to SEC–MALS calculated masses for the indicated fusion proteins. Summary table (bottom) of measured and theoretical calculated molecular masses for each protein complex analysed. Experiments were performed in triplicate.

Extended Data Figure 10 Mutational and functional characterization of molecular interactions involving the NTRs and kinase domains of BRAF and KSR1.

a, Peak intensity change and CSP versus residue number for the [1H–15N]HSQC spectra of the [15N]mCC-SAM L45D mutant (black) and the [15N]mCC-SAM L45D mutant in the presence of hBRS (green) (left), the [1H,15N]HSQC spectra of the [15N]mCC-SAM C60D (black) versus [15N]mCC-SAM C60D in the presence of hBRS (green) (middle), and the [1H–15N]HSQC spectra of [15N]mCC-SAM (black) and [15N]mCC-SAM in the presence of the hBRS M53D mutant (green) (right) at the indicated molar ratio. Cut-off values indicated by the red line were calculated using the corrected s.d. method40. Residue coordinates of the CC-SAM correspond to mouse KSR1; the corresponding human CC-SAM mutations are L47D and C62D. b, The BRS and CC-SAM domains specifically interact as determined by BRET. The CC-SAM L47D and C62D mutations reduce the BRS:CC-SAM BRET binding signal. Similarly, the BRS M53D mutation disrupts the BRS:CC-SAM BRET binding signal. c, Disruption of the K88–E72 salt interaction disrupts BRS:CC-SAM BRET interaction. Restoring the salt bridge with reverse polarity as in the K88E–E72K mutant pair rescues complex formation. d, MEK1 expression induces the formation of full-length BRAF–KSR1 heterodimers, but does not induce the formation of heterodimers between the isolated NTRs of BRAF and KSR1 as detected by co-IP. e, BRET analysis demonstrating that MEK expression promotes KSR1 kinase domain homodimerization. Helix αG mutations in either KSR1 (W831R) or MEK1 (F311S) abolish MEK-induced homodimerization of the KSR1 kinase domain. BRET log2-transformed fold-changes were reported as a function of log10-transformed mCherry relative fluorescence units. f, Co-IP analysis of Pyo-KSR1 with Flag–KSR1 reveals that MEK1 promotes full-length KSR1 homodimerization. Experiments in bf were repeated at least three times. For gel source data, see Supplementary Fig. 1.

Supplementary information

Life Sciences Reporting Summary (PDF 75 kb)

Supplementary Figure 1

This file contains gel source data for immunoblots presented in the main and extended data figures including molecular weight markers (kDa). (PDF 864 kb)

Supplementary Table 1

The raw data of BRET-based whole-kinome screen for modulators of BRAF-KSR1 kinase domains interaction. (XLSX 95 kb)

Supplementary Table 2

MEK1 mutations that enable KSR1W831R-MEK1 interaction. (XLSX 17 kb)

Conformational modification of the CC-SAM domain induced by binding of the BRS domain

Conformational modification of the CC-SAM domain induced by binding of the BRS domain. (MP4 4456 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lavoie, H., Sahmi, M., Maisonneuve, P. et al. MEK drives BRAF activation through allosteric control of KSR proteins. Nature 554, 549–553 (2018). https://doi.org/10.1038/nature25478

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature25478

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

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