Main

Given its evolutionary conservation in metazoans, simpler biological models such as Drosophila and Caenorhabditis elegans have been extensively employed to identify new components of the RAS–ERK pathway and to characterize complex regulatory mechanisms1. For example, genetic screens in flies and nematodes identified the pseudokinase KSR, a close RAF homolog, as a key component acting between RAS and RAF2,3,4. The characterization of KSR revealed that its kinase domain heterodimerizes with the kinase domain of RAF, through a surface that is conserved across all RAF family members, to allosterically activate RAF kinase activity5. Subsequent studies demonstrated that kinase domain dimerization is an essential step for the activation of all catalytically competent RAF family members, including ARAF, BRAF and CRAF in mammals6,7. Furthermore, recent work with mammalian and fly proteins demonstrated that MEK binding to KSR selectively drives RAF–KSR heterodimerization and RAF activation, thus identifying KSR proteins as MEK-dependent allosteric activators of RAF8.

Similar to KSR, the protein Connector eNhancer of KSR (CNK) was identified by genetic means in Drosophila as a key regulatory component of the RAS–ERK pathway9. CNK is a scaffolding protein comprising an amino-terminal conserved sterile alpha motif (SAM), conserved region in CNK (CRIC) and PSD-95/Dlg/ZO-1 (PDZ) domains, followed by a central pleckstrin homology (PH) domain10. In flies, CNK, like KSR, acts at a step between RAS and RAF to facilitate RAF activation, but its precise mechanism of action remains unresolved. Notably, the N-terminal part of CNK (denoted CNKNT) encompassing the SAM, CRIC and PDZ domains and a 260-residue linker preceding the PH domain can recruit the KSR–MEK complex in a manner dependent on its physical association with the SAM-domain-containing protein hyphen (HYP; also known as Aveugle)11,12. The interaction of CNK and HYP is mediated in part by a canonical head-to-tail interaction of their SAM domains13. Whether the interaction of CNK–HYP with KSR–MEK is direct, and the specific role of domains beyond the SAM domain of CNK in RAS–ERK pathway regulation, remain open questions. Here, we functionally and structurally characterized the KSR–MEK–CNK–HYP complex to reveal the mechanism by which KSR–MEK binds CNK–HYP and how this contributes to RAS–ERK signaling.

Results

Structure of CNKNT bound to HYP

In previous experiments, the minimal fragment of CNK interacting with the KSR–MEK complex in the presence of HYP extended from amino acids 2 to 549 (CNK2–549; Fig. 1a)11. These experiments did not address whether KSR and/or MEK could independently associate with CNK2–549. To address this point, we separately expressed a GST-CNK2–549 fusion construct along with the different pairwise combinations of Pyo-tagged HYP, V5-tagged KSR and Myc-tagged MEK constructs in S2 cells and assessed their ability to associate following GST pull-downs. Interestingly, although basal interactions with CNK2–549 were observed for HYP, KSR and MEK when tested alone or in pairs (Fig. 1b, lanes 1–6), co-expression of HYP, KSR and MEK greatly enhanced their association with CNK2–549 (lane 7). These observations suggest cooperative interaction between CNK, HYP, KSR and MEK.

Fig. 1: Crystal structure of a CNK–HYP complex.
figure 1

a, Domain architecture of Drosophila CNK and HYP. Amino acid positions are indicated on top of each depiction. Boundaries of constructs used in this study are represented as solid lines underneath. b, S2 cells were transfected with the indicated plasmid constructs. GST pull-downs were conducted on total cell lysates. Antibodies used to detect protein levels in the GST pull-downs or total cell lysates are indicated to the right of each blot. Co-expressed Pyo-HYP, V5-KSR and Myc-MEK proteins associate with GST-CNK2–549 and GST-CNK2–384, but not with GST-CNK2–209, which lacks the PDZ domain. Results shown here are representative of at least three independent experiments. Signal quantifications relative to control lanes 1–3 (set at 1.00) are shown under the blots. *P ≤ 0.05 using a one-way ANOVA test. Exact P values for quantified signals are as follows. KSR levels: lane 4, P = 0.9775; lane 6, P = 0.9999; lane 7, P < 0.0001; lane 8, P < 0.0001; lane 9, P = 0.4961. MEK levels: lane 5, P > 0.9999; lane 6, P = 0.0018; lane 7, P < 0.0001; lane 8, P < 0.0001; lane 9, P = 0.891. HYP levels: lane 4, P = 0.9611; lane 5, P = 0.9997; lane 7, P = 0.0012; lane 8, P = 0.0007; lane 9, P > 0.9999. c, Ribbon and surface representations of the CNK2–330–HYP complex obtained by X-ray crystallography.

Source data

To identify the minimal portion of CNK able to associate with the KSR–MEK complex, we probed the requirement of domains in addition to the SAM domain of CNK. To this end, we generated two additional carboxy-terminal deletions fused N-terminally to GST; namely, CNK2–384 lacking most of the linker region between the PDZ and PH domains, and CNK2–209 that additionally lacks the PDZ domain (Fig. 1a). As assessed using GST pull-downs, co-expression of epitope-tagged HYP, KSR and MEK proteins also interacted strongly with the CNK2–384 truncation (lane 8). By contrast, the CNK2–209 truncation was severely impaired in its ability to associate with all three proteins (Fig. 1b, lane 9). Together, these results indicate that the formation of a tight complex between CNK, HYP, KSR and MEK is dependent on sequences in addition to the SAM domains; in particular, the PDZ domain.

Given that the CNKSAM–HYPSAM domain interaction is necessary but not sufficient for KSR–MEK recruitment, we sought structural information on CNK beyond the SAM domain using CNK2–330, which minimally comprises the SAM, CRIC and PDZ domains. We determined a 2.1 Å resolution structure of the complex of CNK2–330 with HYP2–106 by single anomalous dispersion phasing using selenomethionine substituted proteins (see Extended Data Fig. 1 and Table 1 for X-ray diffraction and refinement statistics). The CNK–HYP complex adopts a rigid four-lobe torus (or ring) with extensive interactions between all four domains (Fig. 1c). As expected, the HYPSAM domain interacts with the CNKSAM domain in a prototypical head-to-tail manner with a buried contact surface of 518 Å2 described previously (Extended Data Fig. 1b)13. The atypically long helix α5 in HYPSAM facilitates a non-canonical interaction with the CNKCRIC domain (Fig. 1c), which itself adopts a four helical bundle resembling the FAT domain of the FAK kinase family14 (Extended Data Fig. 1c). A buried contact surface of 961 Å2 between HYPSAM and CNKCRIC is composed predominantly of hydrophobic residues on helix α5 of HYPSAM and on helices α2 and α3 of CNKCRIC (Extended Data Fig. 2a). The PDZ domain adopts a canonical fold (Extended Data Fig. 1d) and, despite its C-terminal position within the SAM-CRIC-PDZ sequence, adopts a central position between the CNKSAM and CNKCRIC domains in the toroid structure, making no direct contacts with HYPSAM (Fig. 1c). Furthermore, none of the inter-domain interactions mediated by the PDZ domain are canonical in nature15. A buried contact surface between CNKPDZ and CNKSAM of 667 Å2 comprises a mixture of hydrophobic and hydrophilic interactions (Extended Data Fig. 2d). A buried contact surface between CNKPDZ and CNKCRIC of 296 Å2 is mediated largely by hydrophilic interactions (Extended Data Fig. 2g).

Table 1 Data collection and refinement statistics

Next, we sought to validate the functional relevance of the CNK–HYP structure by testing the impact of specific point mutations on the binding of CNK to HYP in cells and on the ability of CNK and HYP to induce RAS–ERK pathway signaling as evidenced by the phosphorylation of MEK. With respect to the CNKCRIC–HYPSAM interface (Extended Data Fig. 2a), point mutations targeting hydrophobic interactions between helix α3 of CNKCRIC (I123D, L152D and I123D_L152D) and helix α5 of HYP (L90D, I97D and L90D_I97D) significantly reduced the association between Flag-CNK2–549 and GST-HYP as assessed by GST pull-down (Extended Data Fig. 2b). Previous work showed that the phosphorylation of MEK in S2 cells is greatly induced by the CNK–HYP complex in the presence of co-expressed RASV12, RAF and KSR13. Although single-site HYP and CNK mutations had little to no adverse effects on the induction of pMEK levels under these conditions, the double-site mutations in either CNKCRIC or HYPhelix α5 strongly impaired phospho-MEK (pMEK) levels (Extended Data Fig. 2c), thus validating the functional relevance of the affected binding interfaces. Interestingly, the presence of overexpressed RASV12, RAF, KSR and MEK mitigated, to various degrees, the impact of the single point mutations on the CNK–HYP interaction (compare Extended Data Fig. 2b and 2c). However, this effect was not seen with the double mutants, which probably explains their greater impact on pathway signaling (that is, pMEK levels) (Extended Data Fig. 2c). This observation is consistent with the noted stabilizing influence of the KSR–MEK complex on the CNK–HYP interaction (Fig. 1b).

With respect to the CNKSAM–CNKPDZ interface (Extended Data Fig. 2d), mutagenesis of contact residues in each domain, singly or in combination, prevented CNK–HYP association in most instances (Extended Data Fig. 2e). The same mutations were tested for their impact on CNK–HYP-induced pMEK levels, but most had no effects (Extended Data Fig. 2f). Moreover, the effect of these mutations on the CNK–HYP interaction were also offset by the presence of RASV12, RAF, KSR and MEK proteins (Extended Data Fig. 2f). Nevertheless, two sets of double mutations (CNKPDZ_E250R_V252R and CNKSAM_F74R_CNKPDZ_E250R) greatly affected HYP binding in the presence of the other RAS pathway components and considerably reduced pMEK levels (Extended Data Fig. 2f). These results confirm the importance of the CNKSAM–CNKPDZ contact surface for complex formation and signaling activity.

Finally, with respect to the CNKCRIC–CNKPDZ interface (Extended Data Fig. 2g), we introduced mutations to disrupt a salt interaction between D81 and K283. An individual mutation of each residue to the opposite charge (D81R and K283D) reduced complex formation partially, while the combined mutation restored CNK–HYP interaction to wild-type levels (Extended Data Fig. 2h). This result is consistent with the notion that a salt interaction between positions 81 and 283 (with either polarity) contributes favorably to CNK–HYP structural integrity. We also assessed the impact of these mutations on CNK–HYP-induced MEK phosphorylation. Only the K283D single-site mutation weakly but reproducibly reduced pMEK levels (Extended Data Fig. 2i). Consistent with this result, the K283D mutation also hampered complex formation in the presence of the RAS pathway components (Extended Data Fig. 2i). Importantly, this defect, as well as the reduction in pMEK levels, were fully rescued by co-introduction of the D81R mutation (Extended Data Fig. 2i), which restores a favorable salt interaction but with opposite polarity. These results support the relevance of the CNKCRIC–CNKPDZ interface and provide further evidence for the validity of the toroid CNKSAM-CRIC-PDZ–HYPSAM crystal structure.

CNK–HYP binds directly to KSR–MEK

Our structural and functional characterization of the CNK–HYP complex delineated the minimal portion of CNK required for HYP binding and for KSR–MEK recruitment. We previously showed that an isolated kinase domain of KSR bound to MEK is also efficiently recruited to CNK–HYP in cells15. However, it is unclear whether this interaction is direct. To address this issue, we separately produced and purified the CNK2–330–HYP and KSR654–966–MEK complexes to homogeneity using bacterial and insect cell (Sf9) expression systems, respectively and then determined their ability to form a higher-order complex upon mixing. In gel filtration chromatography, the CNK–HYP and KSR–MEK complexes exhibited distinct elution profiles (Extended Data Fig. 3a,e). Markedly, when the two complexes were pre-incubated together, a significant proportion of the CNK–HYP complex shifted to higher molecular weight fractions that overlapped with the KSR–MEK complex (Extended Data Fig. 3a). These results indicate that CNK, HYP, KSR and MEK proteins can associate directly to form a higher-order complex.

Mapping interaction surfaces on CNK and KSR

We next sought to map the surfaces on CNK–HYP that mediate KSR–MEK binding. We reasoned that if the interaction were functionally relevant, then the contact residues would be evolutionarily conserved. Projection of conserved residues onto the surface of the CNK–HYP structure identified multiple residues as candidates for mutagenesis, with the majority on the same surface of the toroid structure (Fig. 2a and Supplementary Figs. 1 and 2). Of ten sites mutagenized on HYP and CNK, a single mutation (M197E) within the linker joining the CNKCRIC and CNKPDZ domains reproducibly reduced the binding of CNK–HYP to KSR, as determined by GST pull-downs (Fig. 2b,c). Follow-on mutagenesis of adjacent conserved CNK residues, including D195K and V198E (but not Q200A), also disrupted the interaction of CNK–HYP with KSR–MEK (Fig. 2b,d). Importantly, none of these mutations altered the interaction of CNK with HYP (Fig. 2e). As an orthogonal approach, we tested a subset of purified mutant proteins for interaction by gel filtration chromatography. Although the D195K, M197E and V198E mutations in CNK did not impair the elution profile of the CNK–HYP complex on their own, they hampered the ability of CNK–HYP to assemble with KSR–MEK (Extended Data Fig. 3b–d). Together, these results identify a conserved surface in the vicinity of the CRIC-PDZ linker of CNK that mediates a direct interaction with KSR–MEK.

Fig. 2: The CRIC-PDZ linker region of CNK is required for recruiting KSR–MEK.
figure 2

a, Ribbon and surface representations of the CNK2–330–HYP complex. Projection of evolutionarily conserved residues is shown as a grayscale (black being the most conserved) onto the surface of the CNK–HYP structure. The positions of conserved residues targeted for mutagenesis are indicated and color-coded according to their domain location. b, Ribbon representation of the CNKCRIC-PDZ linker region. Side chains of residues targeted for mutagenesis are depicted as sticks. c–g, S2 cells were transfected with the indicated plasmid constructs. Antibodies used to detect protein levels are indicated to the right of each blot. GST pull-downs in c–f were conducted on total cell lysates for each condition. The M197E mutation targeting the CNKCRIC-PDZ linker region reduces KSR–MEK recruitment to the CNK–HYP complex (c); the D195K, V198E and M197E mutations in the CNKCRIC-PDZ linker region greatly impede the ability of CNK–HYP to recruit KSR–MEK (d); mutagenesis of the CNKCRIC-PDZ linker region does not impinge on the formation of the binary CNK–HYP interaction (e), but impairs CNK–HYP’s ability to promote RAF activation as assessed by pMEK levels following co-expression of Ha-RASV12, Pyo-RAF, V5-KSR and Myc-MEK (RRKM) (f). g, S2 cells were incubated with dsRNA targeting the 3′UTR of endogenous cnk transcripts 24 h before transfection. SEVS11 expression was induced by 30 min heat shock at 37 °C and incubated for 2 h before collecting the cells. The ability of CNK to promote SEVS11-induced MAPK activation is impeded by the double D195K_V198E mutation targeting the CNKCRIC-PDZ linker region as assessed by pMAPK levels. Experiments in c–g were repeated at least three times. Signal quantifications relative to control lanes (set at 1.00) are shown under relevant blots. *P ≤ 0.05 using a one-way ANOVA test. Exact P values for quantified signals in c are as follows. KSR levels: lane 3, P = 0.9994; lane 4, P = 0.9992; lane 5, P = 0.9968; lane 6, P = 0.9994; lane 7, P = 0.0761; lane 8, P = 0.9891; lane 9, P = 0.9821; lane 10, P = 0.9868; lane 11, P = 0.9385; lane 12, P = 0.9995; lane 13, P = 0.9968. CNK levels: lane 3, P = 0.8486; lane 4, P = 0.6118; lane 5, P = 0.6821; lane 6, P = 0.858; lane 7, P = 0.1317; lane 8, P = 0.9996; lane 9, P = 0.1848; lane 10, P = 0.8816; lane 11, P = 0.4895; lane 12, P = 0.9872; lane 13, P = 0.9877. Exact P values for quantified signals in d are as follows. KSR levels: lane 3, P = 0.9946; lane 4, P < 0.0001; lane 5, P < 0.0001; lane 6, P < 0.0001; lane 7, P = 0.0009; lane 8, P < 0.0001. MEK levels: lane 3, P = 0.9478; lane 4, P < 0.0001; lane 5, P = 0.0001; lane 6, P < 0.0001; lane 7, P = 0.037; lane 8, P < 0.0001. CNK levels: lane 3, P = 0.9937; lane 4, P = 0.0003; lane 5, P = 0.0182; lane 6, P = 0.0003; lane 7, P = 0.8326; lane 8, P = 0.0004. Exact P values for quantified signals in e are as follows. CNK levels: lane 3, P = 0.7654; lane 4, P = 0.5781; lane 5, P = 0.7979; lane 6, P = 0.6404; lane 7, P = 0.8454; lane 8, P = 0.9997. Exact P values for quantified signals in f are as follows. KSR levels: lane 4, P > 0.9999; lane 5, P = 0.0004; lane 6, P = 0.0066; lane 7, P = 0.0004; lane 8, P = 0.4077. MEK levels: lane 4, P = 0.9997; lane 5, P < 0.0001; lane 6, P < 0.0001; lane 7, P < 0.0001; lane 8, P = 0.0036. CNK levels: lane 4, P = 0.7734; lane 5, P = 0.0006; lane 6, P = 0.0019; lane 7, P = 0.0005; lane 8, P = 0.2225. pMEK levels: lane 4, P = 0.2198; lane 5, P = 0.0081; lane 6, P = 0.4627; lane 7, P = 0.0138; lane 8, P = 0.9997. Exact P values for quantified signals in g are as follows. pMAPK levels: lane 3, P = 0.0017; lane 4, P = 0.3245; lane 5, P = 0.0011.

Source data

To ascertain the functional relevance of the predicted binding surface on CNK for KSR–MEK, we tested mutations targeting this area for impact on CNK–HYP-induced pMEK levels in the presence of RASV12, RAF, KSR and MEK. As shown in Fig. 2f, the three strongest mutations in CNK2–549 affecting complex formation with KSR–MEK (D195K, M197E, V198E) significantly decreased pMEK levels.

We next asked whether expression of CNK mutants with impaired interaction with KSR–MEK would adversely affect phospho-MAPK (pMAPK) levels induced by an activated RTK after depletion of endogenous CNK by RNAi. To this end, we introduced the D195K_V198E double mutation into a full-length CNK construct lacking its natural 3′UTR. Expression of a constitutively active sevenless RTK (SEVS11 (ref. 16)) under a heat-shock promoter in S2 cells induces pMAPK levels (Fig. 2g). Co-transfection of a double-stranded RNA (dsRNA) targeting the 3′UTR of cnk transcripts reduced pMAPK levels as expected (Fig. 2g, lane 3) and this effect was fully restored by co-expressing an RNAi-insensitive CNK wild-type construct (lane 4). By contrast, the D195K_V198E mutation in CNK failed to rescue pMAPK levels under these conditions (lane 5). Together, these results provide compelling evidence that a surface involving the CRIC-PDZ linker region of CNK plays an important role in RTK-driven RAS–MAPK signaling in Drosophila cells by mediating direct interaction with KSR–MEK.

After defining the probable binding surface on CNK for KSR–MEK, we next sought to identify the complementary interaction surface on the KSR–MEK complex. Projection of conserved residues onto the surface of the known KSR2–MEK1 crystal structure (PDB 7JUR (ref. 17)) identified candidate regions for mutagenesis on KSR (Fig. 3a and Supplementary Fig. 3). We previously found that mutations in KSR that impaired MEK binding also affected CNK binding11, suggesting that the binding surface on KSR could be close to MEK. Therefore, we first examined three conserved residues (M902, K903 and Y925) located at the base of the C-lobe of KSR, adjacent to the C-lobe of MEK. Notably, the three single-site mutations M902D, K903D and Y925D strongly reduced binding of KSR–MEK to CNK–HYP and also impaired pMEK induction to varying degrees (Fig. 3b). By contrast, a negative control mutation (E926R) had no effect.

Fig. 3: Identification of surface residues on KSR enabling interaction with CNK–HYP.
figure 3

a, Ribbon representations of a face-to-face KSR2–MEK1 kinase domain dimer (PDB 2Y4I (ref. 25)). Projection of evolutionarily conserved residues (shown in grayscale, with black being most conserved) onto the surface of the KSR structure. Residues targeted for mutagenesis are indicated by arrows. The red-colored mutation targets a residue within the αG helix of KSR (W896) that is a known contact point with MEK8. b,c, S2 cells were transfected with the indicated plasmid constructs. Antibodies used to detect protein levels are indicated to the right of each blot. GST pull-downs (b) or anti-V5 immunoprecipitations (c) were conducted on total cell lysates. b, Mutagenesis of conserved residues located at the base of the C-lobe of KSR prevents interaction with the CNK–HYP complex and impedes RAF activation. The mutation of a non-conserved residue in close vicinity (E926R) has no effect. RRM denotes co-transfection of Ha-RASV12, Pyo-RAF and Myc-MEK. c, Mutations of the conserved residues in b (with the exception of W896A) have no or a marginal effect on the binary KSR–MEK interaction. Experiments in b and c were repeated at least three times. Signal quantifications relative to control lanes (set at 1.00) are shown under relevant blots. *P ≤ 0.05 using a one-way ANOVA test. Exact P values for quantified signals in b are as follows. KSR levels: lane 5, P < 0.0001; lane 6, P < 0.0001; lane 7, P < 0.0001; lane 8, P = 0.702. MEK levels: lane 5, P < 0.0001; lane 6, P < 0.0001; lane 7, P < 0.0001; lane 8, P < 0.0001. CNK levels: lane 5, P = 0.0153; lane 6, P = 0.0718; lane 7, P = 0.0773; lane 8, P= 0.9763. pMEK levels: lane 5, P < 0.0001; lane 6, P < 0.0001; lane 7, P < 0.0001; lane 8, P = 0.697. Exact P values for quantified signals in c are as follows. MEK levels: lane 3, P = 0.0015; lane 4, P = 0.9683; lane 5, P = 0.0335; lane 6, P = 0.5215; lane 7, P = 0.0421. CNK levels: lane 3, P = 0.0338; lane 4, P = 0.0096; lane 5, P = 0.026; lane 6, P = 0.0255; lane 7, P = 0.973.

Source data

Given that the three conserved residues in KSR that are relevant for CNK–HYP binding are close to the MEK binding surface, we next tested whether their mutations affected binding of KSR to MEK (that is, to rule out an indirect mechanism of action). Compared with the αG helix mutation W896R in KSR, which is known to abrogate MEK binding8, the M902D and Y925D mutations had no adverse effect on KSR binding to MEK, whereas the K903D mutation had an intermediate effect (Fig. 3c). As an orthogonal approach, we tested the KSRM902D and KSRY925D mutations for their effect on KSR–MEK interaction with CNK–HYP by gel filtration chromatography. Both mutations on their own or combined perturbed the interaction of KSR–MEK with CNK–HYP (Extended Data Fig. 3e–h). Together, these results identify a surface on the C-lobe of KSR centered on M902 and Y925 as the probable direct binding site for CNK–HYP.

Cryogenic electron microscopy structure of KSR–MEK–CNK–HYP

To visualize the binding mode of KSR–MEK to CNK–HYP, we pursued an atomic structure of the minimal KSR–MEK–CNK–HYP complex in the presence of AMPPNP and the MEK inhibitor trametinib by cryogenic electron microscopy (cryo-EM). The resultant structure, determined to 3.3 Å resolution (Fig. 4a, Table 2 and Extended Data Fig. 4), revealed unambiguous Coulomb potential density for the entire CNK–HYP subcomplex, all of MEK including bound AMPPNP and trametinib with the exception of helix α1 (Extended Data Fig. 5a,b), and all of KSR kinase domain with the exception of the N-lobe (Extended Data Fig. 5c). Disorder of the N-lobe of KSR correlated with the absence of ligand in the kinase active site. Aside from the apparent disorder of the KSR N-lobe, the binding mode of KSR to MEK is similar to that observed for the isolated KSR–MEK structure17, with a small 3° pivot of the KSR C-lobe with respect to MEK that eliminates potential for steric clash with CNKCRIC (Extended Data Fig. 6). The activation segments of KSR and MEK adopt similar inactive-like conformations stabilized by an intermolecular antiparallel β-sheet as previously observed17,18. We note, however, that the activation segment of KSR in the KSR–MEK–CNK–HYP structure adopts a more unstructured conformation in comparison to the isolated KSR–MEK structure, which may be because of the disordered nature of the N-lobe of KSR in the former structure (Extended Data Fig. 6d).

Fig. 4: Cryo-EM structure of the KSR–MEK–CNK–HYP complex.
figure 4

a, Cryo-EM map (left) and ribbon representations (right) of the KSR–MEK–CNK–HYP complex. CNK contacts the C-lobes of KSR and MEK kinase domains. KSR engages the CRIC domain of CNK, while MEK engages both the SAM and CRIC domains of CNK (see Extended Data Figs. 6 and 7 for details). b–e, S2 cells were transfected with the indicated plasmid constructs. Antibodies used to detect protein levels are indicated to the right of each blot. Anti-V5 (b) and anti-Pyo (c) immunoprecipitations were conducted on total cell lysates. The CNK–HYP complex enhances KSR–MEK complex formation (b), which drives RAF–KSR heterodimerization and RAF transactivation (c). Mutations at the binding interface between CNK and KSR disrupt the ability of CNK–HYP to drive RAF–KSR heterodimerization and RAF transactivation. The ability of CNK–HYP to promote RAF activity depends on the canonical side-to-side interface of the RAF–KSR heterodimer (d) and on the BRS-CC-SAM interaction between the RAF and KSR N-terminal regulatory regions8 (e). f, Model depicting the role of the CNK–HYP complex in KSR–MEK-driven RAF activation. Experiments in b–e were repeated at least three times. Signal quantifications relative to control lanes (set at 1.00) are shown under relevant blots. *P ≤ 0.05 using a one-way ANOVA test. Exact P values for quantified signals in b are as follows. MEK levels: lane 3, P = 0.0432. Exact P values for quantified signals in c are as follows. KSR levels: lane 3, P = 0.7991; lane 4, P = 0.0125; lane 5, P = 0.869; lane 6, P = 0.36. pMEK levels: lane 4, P = 0.0042; lane 5, P = 0.5287; lane 6, P = 0.9031. Exact P values for quantified signals in d are as follows. pMEK levels: lane 3, P = 0.9977; lane 4, P = 0.9065; lane 5, P < 0.0001; lane 6, P = 0.9976; lane 7, P = 0.9784. Exact P values for quantified signals in e are as follows. pMEK levels: lane 3, P < 0.0001; lane 4, P < 0.0001; lane 5, P = 0.0608.

Source data

The binding surface on CNK–HYP for KSR–MEK is composed of all three domains of CNK (SAM, CRIC and PDZ), whereas HYP makes no direct contact with KSR–MEK (Fig. 4a). The largest contact is made between CNK and KSR (765 Å2 buried surface area) and involves the αG–αH linker and helix αH of KSR and helices α1 and α4 of CNKCRIC and the CNKCRIC–CNKPDZ linker (Extended Data Fig. 7a,b). The second largest contact is made between CNKCRIC and MEK (719 Å2 buried surface area) and involves helices α1 and α2 of CNKCRIC and the αG–αH and the αH–αI linkers of MEK (Extended Data Fig. 7c). The third largest contact is made between CNKSAM and MEK (330 Å2 buried surface area) and involves the α3–α4 linker, helix α4 and helix α5 of CNKSAM and the APE motif-αF linker, αH–αI linker and helix αI of MEK (Extended Data Fig. 7d). Lastly, the C-terminal extension of the PDZ domain makes minor contact with the C-lobe of the KSR kinase domain (216 Å2 buried surface area; Extended Data Fig. 5d).

Overall, the CNK–HYP complex binds across the bottom surfaces of both KSR and MEK kinase C-lobes. Consistent with our contact mapping studies (Figs. 2 and 3), the interaction perturbing D195K, M197E and V198E mutations in CNK and the M902D, K903D and Y925D mutations in KSR directly map to the interface between CNK–HYP and KSR–MEK (Extended Data Fig. 7a). To further validate the KSR–MEK–CNK–HYP structure, we sought to mutagenize residues at additional points of contact between CNK–HYP and KSR–MEK. As above, the mutants were tested for their ability to disrupt binding of CNK–HYP to KSR–MEK (while preserving the binary interactions of CNK with HYP and of MEK with KSR) and for the ability to perturb pathway signaling (Extended Data Fig. 7). Although most single mutations (KSR R913E, MEK E323R, CNKSAM R46F and Y76R, CNKCRIC N96E and R99E) had little to no effect on quaternary complex formation or pathway signaling, three mutations (CNKCRIC N82E, N82W and Q84E) disrupted the interaction between CNK–HYP and KSR–MEK, while the CNKCRIC N82W mutation additionally abolished pathway signaling (Extended Data Fig. 7e,f). Given that binding of CNK–HYP to KSR–MEK appeared resistant to single mutations in earlier studies (Extended Data Fig. 2), we tested the effect of the new mutations in combination. Strikingly, all tested combinations strongly disrupted the interaction between CNK–HYP and KSR–MEK and reduced MEK phosphorylation (Extended Data Fig. 7g,h) without affecting the binary interactions of KSR with MEK or CNK with HYP (Extended Data Fig. 7i,j). Together, these results validate the functional relevance of the cryo-EM structure of CNK–HYP bound to KSR–MEK.

Comparison with 14-3-3 bound RAF complexes

The structure of KSR determined here corresponds to an isolated kinase domain, which raises the question of how other domains and binding partners of KSR might relate to and/or influence binding to CNK–HYP. In quiescent cells, a 14-3-3 dimer stabilizes BRAF in a monomeric auto-inhibited state through simultaneous binding of the 14-3-3 dimer to a pair of phospho-sites that N-terminally and C-terminally flank the kinase domain of BRAF19,20. Additional stabilizing interactions are afforded by the CRD domain of BRAF and by the binding of BRAF to MEK (Extended Data Fig. 8a, left)19,20. As the two 14-3-3 phospho-regulatory sites, the kinase domain and the CRD domain of BRAF are conserved in KSR proteins21, we reasoned that KSR could in principle adopt a similar auto-inhibited conformation. Superposition of the auto-inhibited structure of BRAF with the cryo-EM structure of KSR–MEK–CNK–HYP reveals a small steric clash between CNKPDZ and 14-3-3 bound to the N-terminal phospho-site in BRAF (Extended Data Fig. 8a, right). A 14-3-3 dimer also stabilizes the active state of BRAF by binding to the C-terminal phospho-regulatory sites of two separate BRAF molecules, thereby orienting the RAF kinase domains in a productive side-to-side dimer configuration (Extended Data Fig. 8b, left)19,20,22. As the C-terminal regulatory sites and the side-to-side kinase domain dimerization surfaces of BRAF are conserved in KSR proteins, KSR probably adopts a similar active state structure when bound to RAF. Superposition of the KSR–MEK–CNK–HYP structure with the 14-3-3 promoted BRAF active-state structure revealed no steric clashes. (Extended Data Fig. 8b, right). Although CNK–HYP binding to KSR appears compatible with KSR modeled in both 14-3-3 promoted active-dimer and inactive-monomer states, in the absence of subtle conformational changes to accommodate minor steric clashes, we reason that CNK–HYP binding may favor the former.

CNK–HYP activates RAF by stabilizing KSR–MEK

Our cryo-EM structure reveals an extensive interaction of CNK–HYP with both KSR and MEK kinase domains, suggesting that CNK–HYP stabilizes the interaction between KSR and MEK. We previously showed that MEK binding to KSR1 leads to BRAF allosteric activation by promoting the dimerization of KSR1 and BRAF kinase domains8. This led us to reason that the mechanism by which CNK–HYP activates RAS–ERK pathway signaling is mediated through the same allosteric mechanism.

To investigate this model, we tested the ability of the CNK2–549–HYP complex to enhance the association between KSR and MEK in pull-down experiments. Compared to control (Fig. 4b, lane 2), co-expression of CNK2–549–HYP in S2 cells strongly increased the ability of MEK to co-immunoprecipitate with KSR. Thus, it appears that the CNK–HYP complex promotes and/or stabilizes the formation of the KSR–MEK complex.

Next, we tested whether CNK–HYP binding to KSR–MEK affects the ability of KSR–MEK to interact with RAF. RAF weakly associates with KSR when co-expressed in S2 cells (Fig. 4c, lane 2) and, as expected, this interaction is enhanced by co-expression of MEK (lane 3). Strikingly, co-expression of the CNK–HYP complex in these conditions strongly enhanced RAF association with KSR, which correlated with an enhancement of pMEK levels (lane 4). Demonstrating the relevance of the direct interaction between CNK–HYP and KSR–MEK complexes in RAF–KSR dimerization, the CNKD195K and KSRM902D variants, which abrogates CNK–HYP binding to KSR–MEK (Figs. 2 and Fig. 3 and Extended Data Fig. 3), also disrupted the ability of CNK–HYP to promote RAF–KSR interaction and to enhance pMEK levels (Fig. 4c, lanes 5 and 6).

To confirm that RAF–KSR dimerization induced by CNK–HYP is canonical in nature, we examined the effect of RAFR481H and KSRR732H mutations on the side-to-side dimerization surfaces of the kinase domains, which disable allosteric activation5. As shown in Fig. 4d, CNK–HYP did not stimulate MEK phosphorylation with the RAFR481H or KSRR732H dimer interface mutants (lanes 6 and 7) compared to wild-type proteins (lane 5). Thus, the stimulatory effect of CNK–HYP on MEK phosphorylation depends on the allosteric mechanism by which KSR normally transactivates RAF.

The dimerization of BRAF with KSR1 also involves a direct interaction between the N-terminal regulatory region of both proteins, mediated by the BRS domain in BRAF and the CC-SAM domain in KSR8. To determine whether binding of the BRS and CC-SAM domains is also required for pathway activation by CNK–HYP, we introduced charge reversal mutations in two evolutionarily conserved residues (K62E in BRS and E64K in CC-SAM) known to form a salt bridge in the human BRAF–KSR1 dimer8. As shown in Fig. 4e, both mutations impaired CNK–HYP-induced pMEK levels. However, co-expression of the two mutant proteins fully restored pMEK levels, probably caused by the restoration of a functional salt bridge with opposite polarity between these two residues (lane 5). Together, these findings confirm that the CNK–HYP complex stimulates RAS–ERK signaling by promoting the KSR–MEK interaction, which in turn induces RAF–KSR dimerization and transactivation.

Discussion

RAS-mediated RAF activation is a multi-step process initiated by the interaction of GTP-loaded RAS with the RAS-binding domain of RAF at the plasma membrane21. Relocalization of RAF to the plasma membrane is followed by dephosphorylation of the 14-3-3 N-terminal phospho-site of RAF that precedes the kinase domain. This releases inhibitory binding of one protomer in a 14-3-3 protein dimer while leaving the C-terminal phospho-site interaction intact. The kinase domain of RAF then undergoes dimerization with itself or another member of the RAF family (including the KSR pseudokinases), stabilized by a 14-3-3 dimer binding to the C-terminal phospho-sites of the two kinases. Together, these events allosterically induce the catalytic function of RAF21. Our study reveals a new twist to the RAF activation mechanism by showing how a scaffolding complex composed of two proteins, CNK and HYP, contributes to RAF activation. RAF activation in flies is highly dependent on its dimerization with the pseudokinase KSR8, and recent work has demonstrated how this event relies on the interaction between KSR and MEK (the substrate of RAF)8. Here, we show that CNK in the presence of HYP binds directly to KSR and MEK. The absence of a direct binding interaction between HYP and KSR–MEK suggests that HYP acts by rigidifying CNK in a conformation competent for KSR–MEK binding (Fig. 4f). The binding of CNK–HYP simultaneously to KSR and MEK in turn reinforces the KSR–MEK interaction, which stimulates the formation of RAF–KSR heterodimers and, ultimately, the allosteric activation of RAF (Fig. 4f). Given that RAF can exist as both symmetric homodimers and asymmetric heterodimers with KSR, the ability of CNK–HYP to drive KSR heterodimerization with RAF has the potential to influence the equilibrium between dimer configurations to fine-tune the activation of the RAS–ERK pathway. Overall, our work shows that the KSR–MEK interaction can serve as an effective point of control for RAS–ERK pathway regulation, which was previously not thought to be the case23.

Interestingly, our structure shows that the CNK–HYP complex binds to the same region of KSR and MEK corresponding to the conserved helices αG and αH of the protein kinase fold (Supplementary Fig. 4). This region is close to the region that mediates substrate recruitment (in particular, helix αG) by KSR, MEK18, BRAF24 and also by other protein kinases such as the RNA-dependent protein kinase PKR25 and the cyclic AMP-dependent protein kinase PKA26. This observation further highlights the importance of the αG–αH region as a common element of protein kinase function and control.

Mammalian genomes encode four proteins related to Drosophila CNK; namely, CNK1, CNK2, CNK3 and IPCEF1 (ref. 10). Although some of these have been shown to interact with RAF or KSR proteins and modulate RAS–ERK signaling27,28,29, their molecular function within the pathway remains largely unexplored. Interestingly, human CNK2 and the HYP homolog SAMD12 have recently been shown to control cancer cell migration by mediating ARF6 activation induced by AXL signaling30. This result indicates that minimally, the CNK–HYP complex itself may provide an opportunity for targeted therapeutic intervention. Finally, given that other scaffolding proteins have been found to influence RAS–ERK pathway function (for example, β-arrestin31, IQGAP1 (ref. 32), paxillin33), it will be interesting to determine whether they function in a manner that similarly impinges on RAF dimerization.

Methods

DNA constructs

Copper-inducible pMet constructs were used for most functional and binding studies conducted in S2 cells. pMet constructs used to express Flag-CNK, GST-HYP, V5-KSR, Myc-MEK, Pyo-RAF and Ha-RASV12 have been previously described11,13. The pBS-pCNK-3×Flag-CNKΔ3′UTR used for rescue experiments has been previously described16. pMet-GST-CNK2–549 was generated by cloning a Drosophila cnk DNA fragment produced by PCR (encoding residues 2–549) between the Age1 and Not1 sites of the pMet-GST-His-TEV vector11. The truncated CNK variants (2–384 and 2–209) were generated by introducing a stop codon at residues 385 and 210, respectively, in pMet-GST-CNK2–549. pMet-Pyo-HYP was generated by cloning a PCR product encoding Drosophila HYP (positions 2–106) that included three Pyo epitopes (MEYMPME) at its N terminus into the Kpn1 and Not1 sites of the pMet vector. Novel point mutations in CNK, HYP, KSR and RAF constructs were introduced using the QuickChange II Site-Directed mutagenesis kit (Agilent). All constructs, including those described below, were fully verified by Sanger sequencing.

For protein production in bacteria, PCR products encoding wild-type or mutant versions of CNK (2–330) or HYP (2–106) proteins were inserted in tandem in a pGEX-4T backbone. CNK fragments were cloned in-frame at the C terminus of GST-TEV and were followed by 6×His-TEV-HYP sequences that included a synthetic trc (trp-lac) promoter at the 5′ end. The resulting construct, pGEX-4T GST-TEV-CNK2–330::trc 6×His-TEV-HYP, allowed the co-expression of CNK and HYP proteins in the same bacterial cells, which enhanced the production of soluble CNK proteins.

We used the pbiGBac baculovirus co-expression system34 for producing the Drosophila complex of KSR kinase domain (KSRKD) and MEK in SF9 cells. Two separate library vectors (pLIB) were first generated, namely, pLIB-GST-6×His-TEV-KSR654–966, which corresponds to the kinase domain of KSR fused to GST at its N terminus, and pLIB-MEK, which encodes full-length MEK isoform B (393 residues). Next, the two gene expression cassettes consisting of the polyhedrin promoter, gene of interest and SV40-terminator sequences were amplified by PCR from the pLIB constructs. Linearized pBIG1a vector and the two PCR products were then assembled in a Gibson assembly reaction as previously described34 to generate the pBIG1a-GST-6×His-TEV-KSR654–966-MEK construct.

Cell culture, transfections and RNAi

S2 cells (DGRC) were grown in EX-CELL 420 serum-free medium (Sigma-Aldrich) at 27 °C, while heat-inducible SevS11-expressing S2 cells9 were grown in Schneider’s insect medium (Sigma-Aldrich) supplemented with 10% FBS at 27 °C. For transfections, 7 × 106 S2 cells were seeded in a 60 mm tissue culture dish and transfected the next day with appropriate plasmids using Effectene (Qiagen). dsRNAs were produced and used essentially as previously described35. Cells transfected with copper-inducible vectors were induced 24 h post transfection by adding 0.7 mM CuSO4 to the medium, and cells were typically collected 36 h later for analysis. For rescue experiments, 2 × 106 SevS11-expressing S2 cells were seeded overnight in six-well plates ± dsRNAs (15 μg ml−1) targeting cnk 3′UTR sequences. Cells were transfected 24 h later with the indicated constructs and cultured for 48 h. Cells were then heat-shocked for 30 min at 37 °C and collected 2 h later.

Oligonucleotide sequences (Integrated DNA Technologies) for the synthesis of cnk 3′UTR dsRNAs were:

Top strand

5′-GAATTAATACGACTCACTATAGGGAGAGGATTAGCCCCCGTTTACTTATG-3′

Bottom strand

5′-GAATTAATACGACTCACTATAGGGAGAATTATGTACAGTTGACTTTATTC-3′

Cell lysates, GST pull-downs, immunoprecipitations and immunoblots

For preparation of cell lysates, cells were gently scraped off dishes with a cell lifter, transferred to 15 ml tubes and spun at 1,000 rpm, 4 °C. Media were then removed by aspiration and cell pellets were resuspended in 500 μl of ice-cold Triton lysis buffer (50 mM Tris at pH 7.5, 150 mM NaCl, 10% glycerol, 0.2% Triton X-100, 1 mM EDTA, 1× phosphatase inhibitor cocktail (Sigma-Aldrich), 1 mM sodium vanadate, 20 μM leupeptin, aprotinin (0.15 U ml−1), 1 mM phenylmethylsulfonyl fluoride (PMSF)). Lysing cells were incubated for 15 min in Eppendorf tubes on ice with gentle rocking and then centrifuged at 14,000×g, 4 °C for 10 min to remove cell debris.

GST pull-down assays were conducted as previously described13 except that assays were run overnight at 4 °C and beads were washed four times with 1 ml of ice-cold lysis buffer. Immunoprecipitation assays were conducted by adding antibodies and protein A/G agarose beads (Calbiochem) to cell lysates (~1–2 mg of total proteins), followed by gentle rocking overnight at 4 °C. Beads for both assays were washed three times with ice-cold lysis buffer and resuspended in 100 μl of SDS-loading buffer. Samples were denatured for 5 min in boiling water and resolved on 7%, 10% or 12 % SDS–PAGE. Proteins were transferred to nitrocellulose membranes (Pall Corporation) and blocked with 2% BSA (Sigma-Aldrich) when probed with most primary antibodies, or with 5% SKIM MILK (BioShop) when probed with anti-V5 or anti-pMAPK antibodies. Band intensities were quantified on at least three biological replicates using ImageJ (v. 1.54e)36. Statistical analysis was conducted using GraphPad Prism (v. 9.5.1).

Sources and dilutions for primary antibodies were as follows: rabbit polyclonal anti-GST (1:1,000; Cell Signaling, cat. no. 2622), mouse monoclonal anti-Flag, clone M2 (1:5,000; Millipore Sigma-Aldrich, cat. no. F3165), mouse monoclonal anti-V5, clone SV5-Pk1 (1:5,000; Invitrogen, cat. no. 46-0705), rabbit polyclonal anti-MEK1/2 (1:1,000 Cell Signaling Technology, cat. no. 9122), rabbit polyclonal anti-pMEK1/2 S217/221 (1:1,000; Cell Signaling Technology, cat. no. 9121), rabbit monoclonal p44/42 MAPK, clone 137F5 (1:1,000; Cell Signaling, cat. no. 4695), mouse monoclonal anti-pMAPK and clone MAPK-YT (1:2,000; Millipore Sigma-Aldrich, cat. no. M8159) antibodies. Mouse monoclonal anti-CNK (1:10; clone 26A6A2), anti-Ha (1:200; clone 12CA5), anti-Myc (1:5; clone 9E10) and anti-Pyo (1:5) antibodies were obtained from the supernatants of hybridoma originally generated in the laboratory of G. M. Rubin37.

Protein expression and purification

The construct pGEX-4T GST-TEV-CNK2-330:6×His-TEV-HYP was transformed into BL21-CodonPlus DE3-RIL bacteria (Agilent Technologies) for protein production. Protein expression was performed in TB or M9 minimal media supplemented with l-selenomethionine. Cells were induced at an optical density of 1.5 followed by overnight growth at 18 °C by the addition of IPTG at 1 mM. Collected bacterial pellets were resuspended in lysis buffer composed of 100 mM Bis-Tris pH 6.0, 500 mM NaCl, 25 mM imidazole, 10% glycerol, 5 mM MgCl2, 1 mM TCEP and 1 mM PMSF, and then lysed by homogenization. Lysates were clarified by centrifugation for 40 min at 18,000×g. Proteins were bound to nickel-affinity resin (GE Healthcare), eluted with lysis buffer containing 400 mM of imidazole. Proteins eluted from nickel-affinity resin were then bound to glutathione affinity resin (GE Healthcare) and eluted by cleavage of the GST tag with TEV protease. TEV protease was subtracted from the purified CNK–HYP complex using nickel-affinity resin. The CNK–HYP complex was concentrated and then buffer exchanged by size-exclusion chromatography using a Superdex75 24 ml column (GE Healthcare) equilibrated in running buffer (100 mM Bis-Tris pH 6.0, 500 mM NaCl, 10% glycerol, 5 mM MgCl2, 1 mM TCEP). All purification steps were carried out at room temperature (20–25 oC) to avoid protein precipitation.

The GST-6×His-TEV-KSR654–966 and full-length MEK proteins were co-expressed in Sf9 cells (Thermo Fisher, cat. no. 11496015). Cell pellets were lysed by sonication and homogenization in 50 mM HEPES pH 7.0, 250 mM NaCl, 1 mM TCEP, benzonase and protease inhibitor cocktails (Roche). After centrifugation, proteins were purified using glutathione resin. Proteins were eluted by cleavage of the GST tag using TEV protease. Proteins were further purified by gel filtration using a 120 ml Superdex 200 column equilibrated with 20 mM HEPES pH 7.0, 500 mM NaCl and 1 mM TCEP. All protein fractions corresponding to greater than 95% purity (as assessed by SDS–PAGE) were pooled, concentrated to 40 mg ml−1, and flash-frozen in liquid nitrogen. Protein concentration was determined by ultraviolet–visible absorption spectroscopy at 280 nm wavelength using a NanoDrop spectrophotometer (Thermo Fisher Scientific) and theoretical extinction coefficients. We probed the phosphorylation state of the activation segment of KSR and MEK following gel filtration purification by trypsin digest mass spectrometry. All the detected peptides of the activation segment of KSR correspond to the non-phosphorylated form, suggesting that the activation segment of KSR is not phosphorylated. We detected both phosphorylated and non-phosphorylated forms of peptides corresponding to the activation segment of MEK, indicating that Ser241 (equivalent to S222 in humans) was partially phosphorylated in our insect cell protein preparations.

Analysis of complex formation by size-exclusion chromatography

The formation of the KSR–MEK–CNK–HYP complex was analyzed by mixing an equimolar ratio of wild-type or mutant forms of the CNK–HYP complex with wild-type or mutant forms of the KSR–MEK complex at a final concentration of 34 µM in a final volume of 100 µl in buffer containing 20 mM HEPES pH 7.0, 500 mM NaCl and 1 mM TCEP. Samples were injected into an S200 HiLoad 10/300 column at 0.75 ml min−1 with a fraction size of 1 ml. Fractions were analyzed by SDS–PAGE gel stained by Coomassie blue.

Protein crystallization and crystal structure determination

CNK–HYP protein complexes were crystallized at 20 mg ml−1 at 20 °C in sitting drops with a 1:1 mix of protein and solution containing 0.1 M Bis-Tris propane pH 7.0 and 1.0 M ammonium tartrate dibasic pH 7.0. A native dataset was collected on a flash-frozen crystal cryo-protected in mother liquor containing 20% glycerol at 100 K on the NE-CAT beamline station 24-ID-E, at the Advanced Photon Source at a wavelength of 0.9792 Å. l-selenomethionine-labeled CNK–HYP protein complexes were mixed 50%:50% or 75%:25% with native protein complex and crystallized at 20 mg ml−1 at 20 °C in sitting drops with a 1:1 mix of protein and solution containing 0.1 M Bis-Tris propane pH 6.0–6.25 and 1.5–1.3 M ammonium tartrate dibasic pH 7.0. X-ray diffraction was measured on flash-frozen crystals cryo-protected in mother liquor containing 20% glycerol at 100 K on the NE-CAT beamline station 24-ID-C, at the Advanced Photon Source at a wavelength of 0.9791 Å. Data reduction was performed using the XDS package (v. 20170601)38. The software BLEND (v. 0.6.23) was then used to select, scale and merge 16 different datasets39. Phasing was performed by selenomethionine single-wavelength anomalous dispersion at 5.0 Å using HKL2MAP (v. 0.4b-beta)40 and the SHELX pipeline (SHELXC v. 2016/1; SHELXD v. 2013/2; SHELXE v. 2018/2)41. The atomic model obtained by selenomethionine single-wavelength anomalous dispersion was then used as a search model for molecular replacement with a higher-resolution native dataset using PHASER (v. 2.8.1)42. Model building and refinement were performed using COOT (v. 0.9)43 and PHENIX (v. 1.20.1-4487)44, respectively. Ramachandran statistics are 97.87% favored, 2.04% allowed and 0.1% outliers. All models were validated using MolProbity (v. 1.20.1-4487)45. The data collection and refinement statistics are reported in Table 1.

Cryo-EM specimen preparation and structure determination

The KSR–MEK–CNK–HYP complex was prepared by mixing 0.48 g l−1 (6 µM) KSR–MEK complex with 0.34 g l−1 (6.7 µM) CNK–HYP complex to obtain a final concentration of the quaternary complex at 6 µM at 1.1 molar excess of CNK–HYP, supplemented with 1 mM MgCl2, 1 mM AMP–PNP and 50 µM trametinib. The sample was equilibrated at 4 °C for 30 min before centrifugation at 18,000×g for 20 min at 4 °C. The supernatant was spotted on glow-discharged Quantifoil R2/2 Cu grids 200 mesh, and plunge-frozen using a Vitrobot Mark IV (FEI) at 4 °C under 100% humidity. Grids were stored in liquid nitrogen. Data were recorded on a 200 kV Talos Arctica electron microscope (Thermo Fisher Scientific) equipped with Gatan K2 summit direct electron detector. Movies were automatically collected in counting mode using SerialEM (v. 4.0.9)46 with a pixel size of 0.93 Å. The defocus range was 0.5–2.6 µm and each movie contained 44 frames with a dose per frame of 1.15 e2 for a total dose of 50.53 e2.

All cryo-EM data processing was conducted in cryoSPARC (v. v3.2)47 (Extended Data Fig. 4). Beam-induced full-frame motion was corrected with a built-in cryoSPARC routine, and the contrast transfer functions were estimated using Gctf (v. 1.06)48. A total of 1,179,823 blob-picked particles were extracted from 1,652 manually selected movies and corrected for beam-induced local motion. After several rounds of 2D classification, 2D templates from 154,791 selected particles were used for template picking. Then 1,421,783 template-picked particles were extracted from 1,652 manually selected movies and corrected for beam-induced local motion. After several rounds of 2D classification, 2D templates from 83,668 selected particles were used for topaz picking. A total of 1,006,192 topaz-picked particles were extracted and corrected for beam-induced local motion. After several rounds of 2D classification, 208,623 particles were selected and used for 3D ab initio reconstruction and homogenous refinement, yielding an initial map at ~3.8 Å resolution that improved to ~3.3 Å after per-particle CTF refinement and several rounds of 3D classification using hetero-refinement and local refinement using a soft mask encompassing the entire complex.

The crystal structure of the CNK–HYP complex (Fig. 1) and the structure of the KSR–MEK complex from PDB 7JUR17 were placed into the electron microscopy map. This initial model was manually fitted with UCSF Chimera (v. 1.15)49. Model building was performed using COOT (v. 0.9)43 and refined with PHENIX (v. 1.20.1-4487)44 and ISOLDE (v. 1.4)50. Structural analysis was performed with UCSF Chimera. Figures were made with UCSF Chimera (v. 1.15), UCSF ChimeraX (v. 1.6.1)51 and PyMol (v. 2.5.3) (www.pymol.org). All cryo-EM data collection and refinement statistics are reported in Table 2.

Table 2 Cryo-EM data collection, refinement and validation statistics

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.