Abstract
The mechanistic target of rapamycin complex 1 (mTORC1) controls cell growth and metabolism in response to nutrients, energy levels and growth factors. It contains the atypical kinase mTOR and the RAPTOR subunit that binds to the TOS motif of substrates and regulators. mTORC1 is activated by the small GTPase RHEB and inhibited by PRAS40. Here we present the 3.0 Å cryo-EM structure of mTORC1 and the 3.4 Å structure of activated RHEB-mTORC1. RHEB binds to mTOR distally from the kinase active site, yet causes a global conformational change that allosterically realigns active-site residues, accelerating catalysis. Cancer-associated hyperactivating mutations map to structural elements that maintain the inactive state, and we provide biochemical evidence that they mimic RHEB relieving auto-inhibition. We also present crystal structures of RAPTOR-TOS motif complexes that define the determinants of TOS recognition, of an mTOR FKBP12-rapamycin-binding (FRB) domain–substrate complex that establishes a second substrate-recruitment mechanism, and of a truncated mTOR-PRAS40 complex that reveals PRAS40 inhibits both substrate-recruitment sites.
mTORC1 controls multiple aspects of cell growth and homeostasis, including protein synthesis, lipogenesis, glucose metabolism, autophagy, lysosome biogenesis, proliferation and survival, in response to environmental cues ranging from levels of amino acids, glucose, energy and oxygen to growth factors1–3. mTOR and other pathway proteins are frequently mutated in cancer, and mTOR inhibitors are approved for the treatment of cancer4,5. mTORC1 is a ~1 MDa dimeric complex consisting of the phosphoinositide 3 kinase-related protein kinase (PIKK)6 mTOR, and the subunits RAPTOR and mLST87–10. mTORC1 activation requires nutrients, which are sensed as amino acid levels and induce mTORC1 recruitment to lysozomal membranes through RAPTOR2. There, mTORC1 meets its activator, the small GTPase RHEB, which conveys a second set of signals from environmental cues including energy, oxygen levels, and growth factors3,11,12. In metazoan, growth factors also relieve the inhibition of mTORC1 by PRAS4013–16.
The best-studied mTORC1 substrates are the eIF4E-binding protein 1 (4EBP1), whose phosphorylation destabilizes the 4EBP1-eIF4E complex and activates cap-dependent translation17,18, and the ribosomal S6 kinase 1 (S6K1), which promotes multiple aspects of protein synthesis and anabolic pathways1,17,18. mTORC1 exhibits multiple levels of substrate specificity. Like canonical kinases, mTOR has a sequence preference around the phosphorylation site, but it is limited to the P+1 position being a proline, bulky hydrophobic or aromatic residue19,20. It also employs a substrate-recruitment mechanism whereby RAPTOR binds to a ~5 amino acid Tor signaling sequence (TOS) motif present in the 4EBP1 and S6K1 substrates21,22, as well as the PRAS40 inhibitor13–16. In addition, the crystal structure of an N-terminally truncated mTOR kinase complex (mTORΔN-mLST8) suggested the FKBP12-rapamycin binding domain (FRB) recruits the S6K1 substrate into the recessed active site23. The overall architecture of mTORC1 has been described from 5.9 Å and 4.4 Å cryo-EM reconstructions10,24, but the biochemical and structural mechanisms of regulation by RHEB and PRAS40, and of substrate recruitment remain poorly understood. To address these questions, we determined the cryo-EM structures of mTORC1 and RHEB-mTORC1, at 3.0 Å and 3.4 Å, respectively, and the crystal structures of PRAS40-mTORΔN-mLST8, RAPTOR-TOS, and FRB-S6K1 peptide complexes. The crystallographic findings are shown in the context of the mTORC1-RHEB cryo-EM structure in the composite image of Figure 1a.
Structure of the S6K1 recruitment peptide bound to the FRB domain
To investigate FRB-mediated substrate recruitment23 by crystallography, we fused a 26-residue S6K1 sequence to the FRB domain due to the low affinity of this interaction (KM= 430 μM, Extended Data Figs 1a, b). In the 1.75 Å crystal structure, a 13-residue S6K1 portion adopts an amphipathic helix and packs with the rapamycin-binding site of the FRB (Fig. 1b; Extended Data Fig. 1e). Central to the interface is the S6K1 Leu396 side chain that extends from the amphipathic helix and inserts into an FRB pocket (Phe2039, Trp2101, Tyr2105 and Phe2108) where rapamycin inserts a key aliphatic group25 (Fig. 1c; Extended Data Fig. 1c). Additional FRB contacts are made by Val395 and Val399 that flank Leu396.
We confirmed the importance of these interactions by alanine-scanning mutagenesis of a 38-residue S6K1 substrate polypeptide (thereafter S6K1367–404). The L396A mutation reduced phosphorylation by the truncated mTORΔN-mLST8 complex to 3 % of the wild type level, and the other mutations had effects commensurate with either their FRB contacts or helix-stabilizing roles (Figs 1c and d; Extended Data Fig. 1b). We also assayed full-length, kinase-inactive S6K1 (thereafter S6K1ki) harboring the L396A mutation and found that its levels of Thr389 phosphorylation by mTORC1 or the truncated mTORΔN-mLST8 complex were 53 % and 37 % of the respective w.t. S6K1ki reactions, irrespective of the TOS motif (Fig. 1e). The L396A mutation also reduced Thr389 phosphorylation of S6K1ki transiently overexpressed in HEK293 cells to 34 % of the w.t. level (Fig. 1f).
The mTORC1 substrates1,4 GRB1019,26, TFEB27, MAF128 and LIPIN29 also contain hydrophobic amino acids within ~15 residues of reported phosphorylation sites (Fig. 1g). We thus mutated one or more hydrophobic residues in each substrate in the context of 20-residue synthetic peptides. The mutations reduced phosphorylation of all four substrates by a factor of 4 to 13, consistent with these substrates utilizing the FRB docking site to enter the catalytic cleft (Fig. 1g).
We also investigated 4EBP1, whose Thr37 and Thr46 phosphorylation sites are followed by an amphipathic helix that binds to eIF4E30. Appreciable Thr37 phosphorylation required extending the peptide to the first turn of the amphipathic helix, which when mutated reduced phosphorylation back to a barely detectable level (Fig. 1g). With the Thr46 site, the 20-residue peptide that reaches partway to the amphipathic helix was phosphorylated ~5-fold less than a peptide encompassing the entire helix. In this longer peptide, mutation of the first and second sets of hydrophobic residues reduced phosphorylation by a factor of ~8 and ~3, respectively (Fig. 1g). Consistent with the amphipathic helix being recruited by the FRB, addition of eIF4E, which sequesters the helix, reduced full-length 4EBP1 phosphorylation by both mTORC1 and mTORΔN-mLST8 by a factor of ~4 (Fig. 1h).
Crystal structures of Raptor-TOS motif complexes
We determined the crystal structures of Arabidopsis thaliana Raptor (atRaptor) bound to TOS-motif peptides from human 4EBP1, S6K1 and PRAS40, at 3.0, 3.1 and 3.35 Å resolution, respectively, and apo-atRaptor at 3.0 Å (Fig. 2a and Extended Data Table). The atRaptor residues that contact the TOS peptides are identical in human RAPTOR, and the atRaptor-4EBP1 TOS interface (Fig. 2b) does not differ discernibly in the 3.0 Å human mTORC1-4EBP1 cryo-EM structure described bellow (Extended Data Fig. 2).
RAPTOR has a sausage-like shape, with the N-terminal caspase-homology domain31 at one end, an α–α solenoid of ~8 armadillo repeats in the middle, and a C-terminal WD40 domain at the other end (Fig. 2a). The caspase homology domain can be superimposed on caspase6 with a root-mean square deviation of ~3 Å for 175 Cα atoms, but RAPTOR lacks the caspase Cys-His catalytic dyad31(Supplementary Information discussion). The TOS peptides bind to a groove between the caspase fold and the solenoid, ~65 Å away from the kinase active site (Fig. 1a). One side of this groove is formed by a 4-helix insertion in the caspase fold, and the other side by the first 3 armadillo repeats (Fig. 2a).
All three TOS peptides have an equivalent 8-residue segment ordered in the crystals (Extended Data Figs 3a to d). The key phenylalanine side chain of the TOS consensus21,22 FXΦDΦ (Φ hydrophobic, X any residue) binds to a pocket together with the preceding TOS residue (thereafter F−1 position). In 4EBP1, this unit consists of the Phe114 aromatic ring making π–π stacking interactions with the Gln113 side chain, whereas in S6K1 and PRAS40 the Phe5 and Phe129 aromatic rings make functionally analogous van der Waals contacts with Val4 and Leu128, respectively (Figs 2b to d). In addition, the phenylalanine backbone amide group hydrogen bonds to Tyr475 (human RAPTOR numbering). Alanine mutation of the 4EBP1 Gln113 reduces its affinity for human RAPTOR by a factor of 20, confirming the importance of its stacking with the phenylalanine and recognition of the residue pair as a unit (Kd values in Extended Data Figs 3e and f).
The other conserved TOS residues make overall conserved RAPTOR contacts (Figs 2b to d). The hydrophobic F+2 side chain (Met116, Ile7 and Met131 of 4EBP1, S6K1 and PRAS40, respectively) binds into a tight pocket at the bottom of the groove, and accordingly its alanine mutation reduces binding by nearly two orders of magnitude (Extended Data Fig. 3e). The F+3 aspartic acid side chain forms a hydrogen bond network involving Arg305 and the F+1 backbone carbonyl group, and its mutation increases the Kd 5-fold (Extended Data Figs 3d and e). By contrast, the hydrophobic F+4 side chain (4EBP1 Ile118 and S6K1 Leu9; PRAS40 Glu133 is disordered) binds to a shallow, solvent exposed surface pocket, and its mutation reduces binding only modestly (Extended Data Fig. 3e).
Consistent with the conservation of contacts in the three structures, the PRAS40 inhibitor has a TOS-RAPTOR Kd similar to those of S6K1 and 4EBP1 (Extended Data Fig. 3f). This suggested that PRAS40 has additional mTORC1-interacting elements.
PRAS40 blocks the FRB substrate-recruitment site
To identify additional mTORC1-binding elements of PRAS40, we first tested whether the PRAS40 segment reported to be necessary for inhibition14 (PRAS40114–256) can also inhibit mTORΔN-mLST8 phosphorylating S6K1367–404. PRAS40114–256 inhibited this TOS-independent reaction with an apparent inhibitor constant (Ki) of ~52 μM (calculated from IC50; Fig. 3a), which is significantly lower than the ~430 μM KM of S6K1367–404 (Extended Data Fig. 1b). Deletion of TOS (PRAS40173–256) had no effect as expected, but deletion of 33 additional residues (PRAS40206–256) reduced inhibition by a factor of ~4 (Fig. 3a). Very similar results were obtained with the 4EBP142–64 substrate (Extended Data Fig. 4a).
We next determined the 3.4 Å co-crystal structure of PRAS40173–256 bound to mTORΔN-mLST8 (Extended Data Fig. 4f). The structure revealed two PRAS40 anchor points separated by an unstructured segment: an amphipathic α helix (residues 212–232) bound to the FRB domain and a β strand (residues 188–196) bound to the mLST8 WD40 domain (Fig. 1a and Extended Data Fig. 4b).
The amphipathic helix binds to the same FRB site as the S6K1 substrate, but at 5 turns it is substantially longer and makes more extensive contacts than S6K1 (Figs 3b and c; Extended Data Fig. 4c). PRAS40 uses the Met222 side chain to bind to the same rapamycin-binding pocket as the S6K1 Leu396 (Fig. 3c), and five additional hydrophobic side chains (Leu215, Ile 218, Ala219, Leu225 and Val 226) to contact an extended FRB surface.
The PRAS40 β strand has a phenylalanine side chain (Phe193) inserting into a pocket between two mLST8 β propeller blades (Tyr195, Trp197, Pro167, Pro212), while its peptide backbone makes three β sheet hydrogen bonds to the edge of one β propeller (Fig. 3d and Extended Data Fig. 4d). These PRAS40 interactions are consistent with reduced inhibition by PRAS40206–256, which lacks the β strand, compared to PRAS40173–256 (Fig. 3a).
We further confirmed the importance of the amphipathic helix by mutating Met222 and four additional FRB-interacting residues (L215A, I218A, A219G, L225A) in full-length PRAS40. As shown in Figure 3e, the mutations reduced inhibition of mTORC1 phosphorylating full-length 4EBP1 by a factor of ~50.
Cryo-EM structures of active RHEB-mTORC1 and apo-mTORC1
To address how RHEB activates mTORC1, we collected cryo-EM data on mTORC1 that was cross-linked in the presence of excess RHEB-GTPγS and 4EBP1 (Extended Data Fig. 2c). The 3D auto-refinement of 580,768 particles in C2 symmetry led to a consensus reconstruction extending to 3.2 Å resolution, as determined by the gold-standard fourier shell correlation (FSC) procedure32 (Extended Data Figs 5a and b). In subsequent 3D classification in C1, most classes appeared to belong to a continuum of conformational states. One class with ~20 % of the particles had an overall conformation distinct from the ensemble of the other classes, and this was the only class that contained RHEB, one on each mTOR of dimeric mTORC1 (Extended Data Figs 5c to e).
Because of the conformational flexibility in between and within the two mTOR-RAPTOR-mLST8 complexes (Extended Data Fig. 5c), we converted the particles to monomers with partial signal subtraction33, and calculated focused reconstructions with three partially overlapping masks (2.98, 2.95 and 2.96 Å; Extended Data Figs 5a and b). Using these three reconstructions with the composite map option of REFMAC534, we refined mTORC1 at 3.0 Å resolution (Extended Data Figs 2c, 5f, 6 and 7a). Using the same procedure, we refined RHEB-mTORC1 at 3.4 Å (Extended Data Figs 2c, 5a and b).
As with the mTORΔN-mLST8 crystal structure23, the PIKK-specific FAT domain adopts a C-shaped solenoid structure that clamps onto the kinase domain (KD), with the start of the solenoid interacting with the KD N lobe, and its end with both the N and C lobes (Fig. 4a). In keeping with the secondary structures of previous mTOR cryo-EM reconstructions24,35, the N-terminal segment missing from mTORΔN starts with an α–α solenoid of 18 HEAT repeats (N-heat), followed by a smaller middle solenoid of 7.5 HEAT repeats (M-heat), and a ~110 residue helical-repeat segment that is structurally contiguous with the subsequent FAT domain and will henceforth be included in the “FAT” descriptor (Extended Data Figs 6, 7b and c). The FAT solenoid acts as an organizing center of mTOR. In addition to its ends clamping onto the KD N and C lobes, the FAT mid-portion packs with the N-heat solenoid end, and the beginning of the FAT packs with the start of the M-heat solenoid, anchoring the two domains (Fig. 4a; Extended Data Figs 6 and 7d; Supplementary Information discussion).
The mTORC1 dimer forms through the C-terminal portion of N-heat packing with the M-heat of the second protomer in a reciprocal fashion (Fig. 4b). RAPTOR binds to this interface as well, resulting in a tripartite interface of 5,420 Å2 buried surface area, of which 38 % is from RAPTOR–N-heat, 23 % RAPTOR–M-heat, and 39 % N-heat–M-heat (Extended Data Figs 7e and f).
RHEB binds to the amino-terminal portions of the N-heat, M-heat and FAT domains, forming a 4-way interface (Figs 4a and c). The majority of the contacts are made by the RHEB switch I and switch II regions, which have GTP-dependent conformations as with other small GTPases36. Switch I (residues 33–41) binds to M-heat and FAT, whereas the longer switch II (residues 63–79) interacts with all three mTOR regions (Fig. 4c and Extended Data Fig. 8a).
RHEB induces a conformational change that activates mTORC1 allosterically
In apo-mTORC1, the N-heat RHEB binding site is far away from those on M-heat and FAT, displaced by ~18 Å relative to its position in the RHEB-bound state (Fig. 5a). On RHEB binding, the N-heat solenoid swings in towards M-heat through a ~19° rotation, reconstituting the RHEB-binding site and also inducing new interactions between the N-terminal portions of N-heat and FAT (Figs 4c and 5a; Supplementary Video 1). This causes a conformational change within the FAT domain whose middle portion gets twisted and dragged by the moving N-heat solenoid end that is anchored on it. The intra-FAT conformational change is entirely distinct from the conformational flexibility apo-mTORC1 exhibits (Extended Data Fig. 7f; Supplementary Information discussion). The two conformations are incompatible in a mixed dimer, as this would require a > 20 Å offset in the N-heat portion of the dimerization interface. This explains the lack of single-RHEB 3D classes, and suggests that two RHEB molecules bind to mTORC1 cooperatively. In support, we find that the RHEB-GTPγS response curve of mTORC1 phosphorylating 4EBP1 best fits a Hill slope model with a Hill coefficient of ~2.0 (Fig. 5b; the ~100 μM EC50 in solution likely not reflective of the membrane-surface reaction in vivo2,36).
The intra-FAT conformational change occurs at hinge regions that allow for relative rotations of flanking segments. One hinge around residue 1443 is associated with a major rotation of 30° between the FAT sub-domains TRD1 and TRD223, while two other hinges exhibit smaller rotations (Fig. 5c). These conformational changes in the FAT are coupled to its C-terminal portion moving away from the kinase, the N lobe of the kinase moving in to the space vacated by the FAT, the FAT–N lobe interface repacking into a looser arrangement, and the catalytic cleft between the N and C lobes closing by 8° (Figs 5c and d; Supplementary Video 2).
The closing of the catalytic cleft changes the relative orientation of the ATP-contacting and catalytic residues from the N and C lobes. This brings the ATP phosphate groups that are bound by the N lobe into closer proximity to critical C lobe residues that include the Mg2+ ligands23 (Asn2343 and Asp2357) and the two catalytic residues23 (Asp2338 and His2340) (Figs 5e and f; Extended Data Figs 8b and c).
This indicates that RHEB activates mTORC1 by allosterically realigning active site residues, bringing them into correct register for catalysis. To confirm this, we compared the steady state kinetic constants of S6K1367–404 phosphorylation by mTORC1 in the presence of 250 μM RHEB-GTPγS or RHEB-GDP. Activation is accounted entirely by a ~30-fold increase in kcat, from 0.09 s−1 to 2.9 s−1, whereas KM values remain essentially unchanged (Fig. 5g). RHEB similarly increased the apparent kcat of full-length 4EBP1 phosphorylation, including under single-turnover conditions, the latter indicating that the kcat effect involves the catalytic step and not a hypothetically rate-limiting product-release step (Extended Data Figs 8d and e). RHEB-GTPγS also accelerated idle ATP hydrolysis, a low-level activity common in protein kinases as well as PI3K37 (Extended Data Fig. 8f).
Cancer-associated hyperactive mTOR mutants
Cancer-associated hyperactivating mutations5,38–40 predominantly involve structure-stabilizing residues. They cluster at the major intra-FAT hinge, the FAT-N lobe packing transition, and the N lobe anchor in a pocket between the C lobe and FAT, suggesting that they act by lowering the barriers to the N lobe adopting the active conformation, mimicking RHEB’s effects (Fig. 6a and Extended Data Figs 9a to c).
This hypothesis predicts that the mutations should lower the EC50 of activation by RHEB, as part of the RHEB-mTORC1 binding energy must be used to affect the conformational change, and that they should not synergize with saturating RHEB. To test these predictions, we transiently expressed and purified four representative hyperactive mTORC1 mutants5,38: A1459P in the middle of a helix at the major intra-FAT hinge, T1977R buried at the FAT-N lobe transition, and S2215Y and E2419K at the N lobe-C lobe and juxtaposed C-lobe–FAT interfaces, respectively (Fig. 6a; Extended Data Figs 9a to c).
After confirming that the mutations increase the kcat of mTORC1 phosphorylating S6K1367–404 without affecting the KM (Extended Data Fig. 9d), we assessed their RHEB-GTPγS dose response curves. All four mutations shifted the response curve to lower RHEB concentrations compared to w.t. mTORC1, produced as the mutants (Fig. 6b). A1459P, S2215Y and E2419K reduced the EC50 comparably, by factors of 6.6, 7.0, and 7.4, respectively, whereas T1977R reduced it by a factor of 4.0 (Fig. 6b). Importantly, the mutations did not synergize with RHEB, as at the highest, nearly saturating RHEB concentration, the mutants exhibited S6K1367–404 phosphorylation levels within ~15% of the w.t. control (v/[E] of 3.7 s−1 for w.t., and 4.1–4.4 s−1 for the mutants; Fig. 6b).
The mutations also reduced the cooperativity of RHEB binding, suggesting they allow the formation of single-RHEB containing mTORC1 (Fig. 6b). We presume this is due to the destabilization of the inhibitory FAT clamp, either directly or indirectly at its anchors on the N and C lobes, allowing the FAT-bound N-heat to reach the tripartite dimerization interface of a mixed dimer.
Conclusion
We show that the TOS motif docking site is ~65 Å from the kinase active site, suggesting that it acts to increase the effective substrate concentration, and we establish a second substrate-docking site corresponding to the rapamycin-binding site at the entrance of the catalytic cleft. PRAS40 binds to both substrate-docking sites and an additional site on mLST8 to achieve inhibition. We also show that the low kinase activity of apo-mTORC1 is due to a misalignment of the kinase N and C lobes and their associated ATP-binding and catalytic residues. The FAT clamp, present in all PIKKs, is a key auto-inhibitory element that keeps the N lobe misaligned. RHEB, binding ~60 Å away from the active site, induces a movement of the N-heat domain, which pulls and twists the FAT clamp, freeing the N lobe to adopt the active conformation. The end result of this process is likely mimicked by cancer-associated mutations that activate mTOR.
METHODS
Protein expression and purification
For the FRB-S6K1389–414 fusion protein, a synthetic gene encoding the FRB domain of human mTOR (residues 2018–2114) followed by a three amino acid linker (SGG) and residues 389–414 of human S6K1αII was cloned into a modified pGEX4T3 vector. The fusion protein was overexpressed in the Escherichia coli strain BL21(DE3), and was purified by glutathione affinity chromatography, removal of the GST tag with TEV protease, and fractionation by ion exchange and size exclusion chromatography. The peak fractions were concentrated to 40 mg ml−1 in 20 mM Tris-HCl, pH 8.0, 0.5 M NaCl, 10% glycerol and 0.5 mM tris-(2-carboxyethyl) phosphine (TCEP).
Human S6K1367–404 wild type and mutant peptides were produced in E. coli as GST-tag proteins using a modified pGEX4T3 vector. Following glutathione affinity chromatography and on-bead cleavage with GST-TEV protease, the recombinant peptides were purified by reversed phase high performance liquid chromatography (HPLC) using a Zorbax 300SB-C3 column (Agilent). All other peptides used for kinase assays were purchased from Bio-Synthesis Inc. or Peptide 2.0 and further purified on HPLC using Zorbax 300SB-C3 or C18 columns when necessary. Molecular weights for all purified peptides were verified using MALDI-TOF mass spectrometry. Peptides used in crystallization experiments were purchased from PEPTIDE 2.0. 4EBP1 and S6K1ki and their mutants were prepared as previously described23. For the TOS-mutated S6K1ki, TOS−, residues 5 to 9 (FDIDL) were mutated to alanine, and for 4EBP1TOS−, residues 114 to 118 (FEMDI) were mutated to alanine.
Human eIF4E was produced by infecting High Five insect cells with a pFastBac1 baculovirus expressing the GST-tagged protein and purified similarly as described above, except that 0.1 mM 7-Methyl-GDP (Sigma) was added to the elution buffer during the glutathione affinity chromatography.
PRAS40 fragments were overexpressed and purified as with the FRB-S6K1389–414 fusion protein. For selenomethionine (SeMet) substituted PRAS40 fragments, E. coli transformants were cultured in LB media at 37°C until the OD600 reached 0.8, pelleted, washed, and resuspended in M9 media containing 50 mg l−1 Leu/Ile/Val, 100 mg L−1 Phe/Lys/Thr, and 90 mg L−1 SeMet. The culture was incubated for 30 minutes to shut down methionine biosynthesis. Protein expression was induced with 1 mM IPTG for 3 hours at 37°C.
Full-length human PRAS40 (PRAS40wt) was produced by infecting High Five insect cells with pFastBac1 baculovirus expressing the GST-tagged fusion protein. It was purified by glutathione affinity chromatography, removal of the GST tag with TEV protease, and by anion exchange and size exclusion chromatography. The peak fraction was concentrated to ~1 mM in 20 mM Tris-HCl, pH 8.0, 400 mM NaCl, 10% Glycerol and 10 mM DTT. PRAS40α-mut containing five mutations on it’s FRB binding α helix (L215A, I218A, A219G, M222A and L225A, generated by QuikChange Lightning Multi Site-Directed Mutagenesis kit, Agilent Technologies) was produced similarly.
Arabidopsis thaliana Raptor (Genebank Q93YQ1) containing an internal deletion (residues 883–942) was produced by infecting High Five insect cells with a pFastBac1 baculovirus expressing the GST-tagged protein. It was purified by glutathione affinity chromatography, removal of the GST tag with TEV protease, and by ion exchange and size exclusion chromatography. The peak fraction was concentrated to 7–9 mg ml−1 in 20 mM Tris-HCl, pH 8.0, 400 mM NaCl, 5% Glycerol, 0.5 mM TCEP.
The mTORC1 complex and mTORΔN–mLST8 complex were prepared from stably-transfected cell as previously described23. The cancer-derived mTORC1 mutants and the matched w.t. control were produced by transient transfection. FLAG-tagged human mTOR wild type and four cancer-derived hyperactivating mutants, A1459P, T1977R, S2215Y, and E2419K (generated with QuikChange Lightning Site-Directed Mutagenesis kit, Agilent Technologies), were cloned into the pcDNA3.1(+) vector. The corresponding mTOR plasmids (40 μg per 15 cm plate) were cotransfected with the pcDNA3.1(+)-human mLST8 plasmid (5 μg per 15 cm plate) using Lipofectamine 2000 into a HEK293-F cell line stably overexpressing the FLAG-tagged human RAPTOR. After incubation for 2 days, ~60 plates of cells per complex were harvested by gentle scraping and lysed in 50 mM Tris-HCl, pH 8.0, 500mM NaCl, 1 mM EDTA, 1 mM EGTA, 10% (v/v) glycerol, 2 mM dithiothreitol (DTT) and protease inhibitors using French Press. After lysate clarification, the Flag-tagged proteins were isolated using anti-Flag M2 agarose beads (Sigma), extensively washed with lysis buffer supplemented with 200 mM Li2SO4, and eluted with 0.2 mg ml−1 flag peptide in lysis buffer. They were then further purified by anion exchange chromatography (MonoQ). Peak fractions containing the wild type or mutant mTORC1 complexes were then concentrated to ~1 mg ml−1 (~2 μM) and stored in small aliquots in −80°C.
For the preparation of nucleotide-bound RHEB, RHEB was overexpressed and purified similarly as FRB-S6K1389–414 fusion protein. To charge RHEB with GDP, the protein was first dialyzed overnight against a buffer containing 20 mM Tris-HCl, pH 8.0, 300 mM NaCl, 5% Glycerol, 5 mM EDTA, 0.5 mM TCEP, and 100 mg L−1 acid-washed activated charcoal (Sigma). Next, the protein was incubated with 30-fold molar excess of GDP (Sigma-Aldrich) on ice for 30 minutes followed with addition of 15 mM MgCl2. To charge RHEB with GTPγS, the protein was incubated with 30-fold molar excess of GTPγS (Sigma), 20 mM EDTA, and 10 units of alkaline phosphatase (New England BioLabs) per mg of RHEB at 30°C for 1 hour followed with addition of 40 mM MgCl2. Both states of charged RHEB were then purified by size exclusion chromatography (Superdex 75) in 20 mM Bicine, pH 8.0, 200 mM NaCl, 5% glycerol, 5 mM MgCl2 and 0.1 mM TCEP. The peak fractions were concentrated to 25–40 mg ml−1.
Crystallization and X-ray data collection
FRB-S6K1389–414 crystals were grown by the hanging-drop vapor diffusion method at 16°C from 100 mM Bis-tris propane, pH 7.0, 30% Tacsimate (Hampton Research). Crystals were transferred to 100 mM Bis-tris propane, pH 7.0, 6.3 M NaFormate and flash-frozen in liquid nitrogen.
AtRaptor native crystals were grown at 4°C by the hanging drop vapor diffusion method from a crystallization buffer of 100 mM HEPES, pH 7.5, 50 mM Ca(OAc)2, 4–8% PEG 8000, 0.5 mM TCEP. For heavy atom derivatives, crystals were soaked in 0.4 to 0.8 mM thimerosal for 2 hours before cryo-protection (not shown). The atRaptor-TOS complexes were grown by streak seeding native crystals into atRaptor mixed with a 2-fold molar excess of synthetic TOS peptides corresponding to residues 99–118 of human 4EBP1 (RNSPEDKRAGGEESQFEMDI; only underlined residues are ordered in the crystals), residues 1–14 of human S6K1 (MAGVFDIDLDQPED), and residues 124–139 of human PRAS40 (DNGGLFVMDEDATLQD). Crystals were cryo-protected in crystallization buffer supplemented with 20% PEG 400 and flash-frozen in liquid nitrogen.
For the initial PRAS40–mTORΔN–mLST8 and SeMet-PRAS40 containing complex, 30 μM mTORΔN–mLST8 was incubated with 0.2 mM of truncated PRAS40, followed by crystallization by the hanging-drop vapor diffusion method. Co-crystals of the complex grew in the same condition as the apo mTORΔN–mLST8 crystals23, and subsequent PRAS40 crystals were prepared by soaking apo mTORΔN–mLST8 crystals with 0.5mM PRAS40 fragments. Crystals were cryo-protected as described23.
All diffraction data were collected at −170°C at the ID24C or ID24E beamlines of the Advanced Photon Source. Crystals containing SeMet were collected at the Se edge using the inverse beam strategy. Data were processed with the HKL suite41.
Crystal structure determination and refinement
For the FRB-S6K1 structure, initial phases were obtained by molecular replacement with PHASER42 using the FRB structure (PDB 1FAP) as the search model. The FRB-S6K1 and other crystal structures described in this study were built using O43 and refined with REFMAC542 and PHENIX44. The FRB-S6K1 model has a Molprobity clashscore of 0.36. The Ramachandran plot has 99.8, 0.2, and 0 % of the residues in the favored, allowed and outlier regions, respectively.
For atRaptor, initial phases were calculated from two thimerosal derivatives using both isomorphous and dispersive differences. The atRaptor used in crystallization has an internal deletion of residues 883–942, a region that is highly susceptible to proteolysis. In the crystals the residues 866–882 and 943–956 flanking the internal deletion, and residues 736–830 are disordered. The corresponding regions are also disordered in the 3.0 Å-refined human RAPTOR structure from our cryo-EM reconstruction. The atRaptor model has a Molprobity clashscore of 4.77. The Ramachandran plot has 89.66, 7.75 and 2.59 % of the residues in the favored, allowed and outlier regions, respectively.
The PRAS40–mTORΔN–mLST8 crystal structures were solved by molecular replacement using the mTORΔN–mLST8 structure23 as a search model. The model has a Molprobity clashscore of 1.92, and the Ramachandran plot has 93.74, 5.3 and 0.95 % of the residues in the favored, allowed and outlier regions, respectively.
Cryo-EM sample preparation and data collection
Cryo-EM samples were prepared using two different crosslinking procedures. In the initial, high-salt crosslinking procedure, 0.42 μM mTORC1 was incubated with 250 μM RHEB-GTPγS, 10 μM 4EBP1, 0.2 mM GTPγS, 1 mM AMP-PNP for 10 minutes, and crosslinked with 0.18 % (v/v) glutaraldehyde for 45 minutes on ice in 20 mM Bicine, pH 8.0, 300 mM NaCl, 10% glycerol, 5 mM MgCl2, and 0.5 mM TCEP. The reaction was quenched with 100 mM Tris-HCl, pH 8.0 and the mixture was purified by size exclusion chromatography (Superose 6) in 20 mM Tris-HCl, pH 8.0, 260 mM NaCl, 5 mM MgCl2 and 0.1 mM TCEP. Peak fractions were concentrated by ultrafiltration to 1 mg ml−1 and were supplemented with 100 μM RHEB-GTPγS, 10 μM 4EBP1, 0.2 mM GTPγS and 1 mM AMPPNP. The sample (3 μl) was applied to glow discharged UltrAuFoil 300 mesh R1.2/1.3 grids (Quantifoil). Grids were blotted for 2.2 s at 22°C and ~95% humidity and plunge-frozen in liquid ethane using a FEI Vitrobot Mark IV.
After determining that high ionic strength severely reduced RHEB activation (not shown), we reduced the salt concentration and modified the crosslinking procedure by first reducing the concentrations of mTORC1, RHEB-GTPγS and NaCl to 0.21 μM, 120 μM and 100 mM, respectively, then crosslinking with 0.24 mM BS3 for 45 minutes on ice, followed by the addition of 130 μM RHEB and 160 mM NaCl and further crosslinking with 0.18 % of glutaraldehyde for 45 minutes. Quenching and purification were performed as with the high-salt procedure in 260 mM NaCl. The main data set used for apo-mTORC1 reconstruction was collected at the NYSBC Simons Electron Microscopy Center using LEGINON45 on a Titan Krios microscope operated at 300 kV and equipped with a Gatan K2 Summit camera using defocus values ranging from −1.2 to −3μm. The camera was operated in counting mode with a 1.331 Å pixel size at the specimen level and a dose rate of 8.3 electrons per pixel per second. Each 12 second exposure was dose-fractionated into 60 frames and contained a total dose of ~56 e− per Å2. This dataset, which was acquired over 7 sessions, consisted of 4,502 micrographs from high-salt crosslinked samples and 3740 from lower-salt crosslinked samples. The second data set, which employed only the lower-salt crosslinking condition, was collected at the Sloan-Kettering Institute Titan Krios microscope/Gatan K2 Summit camera operated at 300 kV with a 1.089 Å pixel size and 8.0 electrons per pixel per second. Each 8 second exposure was dose-fractionated into 40 frames and contained a total dose of ~52 e− per Å2. This data set consisted of 8,354 micrographs.
Cryo-EM image processing
Motion correction was performed with MOTIONCORR and MOTIONCORR-246 for the first and second data sets, respectively. Contrast transfer function parameters were estimated with CTFFIND447, and subsequent 2D/3D classifications and 3D refinements were carried out with RELION-1.448 and RELION-2.049. All reported resolutions are from gold-standard refinement procedures with the FSC=0.143 criterion32 after post-processing by applying a soft mask, correction for the modulation transfer function (MTF) of the detector, temperature-factor sharpening, and correction of FSC curves to account for the effects of the soft mask as implemented in RELION48,49. Initial references for template-based particle picking were from 2D class averages of manually picked particles. Multiple rounds of 2D and, for some data subsets, 3D classifications were then used to remove false positives and particles that clustered in classes with poorly determined orientations. For the main data set, a total of 580,768 particles were retained (Extended Data Fig. 5a). After an initial 3D auto-refinement with C2 symmetry, the particles were improved by particle-based motion correction and radiation-damage weighting50. The resulting ‘polished’ particles were used for the 3D auto-refinement of a consensus mTORC1 dimer map in C2, yielding a 3.23Å reconstruction (Extended Data Fig. 5a). After 3D classification in point group C1, 114,879 (19.7 %) of the polished particles clustered to a single RHEB-containing class. The fraction of RHEB-containing particles was ~2.5 fold higher in the low-salt crosslinked particles compared to the high-salt ones. The remaining 6 apo-mTORC1 3D classes appeared to sample a continuum of conformational flexibility, both between and within the two mTOR-RAPTOR-mLST8 complexes (Extended Data Fig. 5a). Because of this conformational flexibility, we converted the particles to a “monomeric” form (illustrated in Extended Data Fig. 5a). For this, we duplicated the particle list and advanced the RELION rot angle by 180° to extract the signal of the 2nd copy of the complex superimposed on the first. We then subtracted the signal from the masked map of the second mTOR-RAPTOR-mLST8 complex in the consensus reconstruction (before post-processing) from each particle in the combined set as described33. 3D auto-refinement then yielded a monomeric mTORC1 consensus reconstruction to 3.11 Å. An alternate, “pseudo-monomer” set of particles was calculated by switching the two N-heat domains, as this domain seemed to move more relative to its own mTOR than that of the second protomer. 3D auto-refinement of the “pseudo-monomer” yielded a reconstruction to 3.02 Å. Upon further inspection of the modes of flexibility in 3D classification of the “monomeric” complexes, three partially overlapping soft masks were constructed for masked 3D auto-refinements with local searches of orientation angles. One focused on N-heat (17–903), FAT-KD (1261–2549), the KD-proximal M-heat portion (933–1005), and mLST8 of the actual monomer (mask1 in Extended Data Fig. 5a), and produced a reconstruction to 2.98 Å. The second used the pseudo-monomer focusing on N-heat of protomer 2, RAPTOR, and the RAPTOR-proximal M-heat portion of protomer 1 (1006–1222), and produced a reconstruction to 2.95 Å (mask2 in Extended Data Fig. 5a). The third also used the pseudo-monomer, focusing on N-heat, a smaller portion of the proximal M-heat (1027–1222), and nearby segments of RAPTOR (195–214, 237–255, 274–503), to 2.96 Å (mask3 in Extended Data Fig. 5a). This third reconstruction had the best density for N-heat. The same procedure of particle duplication and signal subtraction was used to convert the 114,879 RHEB-mTORC1 particles to monomer and pseudo-monomer forms (Extended Data Fig. 5a). These two sets of particles were combined with the RHEB-containing monomer and pseudo-monomer particles of the second data set that were processed similarly and which were down-scaled in fourier space with RELION to match the magnification of the first data set. The optimal scale factor was determined by comparing the KD C lobe portion of the refined structures from the two data sets to the crystal structure of mTORΔN–mLST8, minimizing the sum of differences squared in the inter-Cα distances. The combined total of 396,474 RHEB-containing particle sets were then 3D auto-refined as with the consensus structure, except for the inclusion of RHEB bound to M-heat mask 1, and in masks 2 and 3 the inclusion of RHEB and the minimal RHEB-interacting M-heat and FAT elements of protomer 2. The dimeric mTORC1-RHEB reconstruction was to 3.8 Å, the monomers to 3.58 Å, and the three focused 3D refinements to 3.43, 3.41 and 3.38Å (Extended Data Fig. 5a). The inclusion of the down-scaled particles from the second data set improved the resolution limits only marginally, but the maps had better continuity, especially in the relatively less ordered regions.
Cryo-EM structure refinement
Model refinement was done with REFMAC5 modified for cryo-EM34, with a composite map of the three focused 3D reconstructions assigned to the following coordinates: N-heat, and for RHEB-mTORC1, RHEB-GTPγS and minimal RHEB-interacting M-heat and FAT portions (both duplicated) to 3rd focused map; M-heat (961–1222), RAPTOR, TOS, and RAPTOR-bound β strand from M-heat FAT linker to 2nd focused map; M-heat (933–960), FAT-KD (1261–2549), AMPPNP, mLST8, and for RHEB-mTORC1, RHEB-GTPγS and minimal RHEB-interacting regions of N-heat (duplicated) of protomer 2 to 3rd focused map. The three focused maps were aligned on the corresponding regions of the consensus C2 map by first obtaining the rotation-translation matrix with CHIMERA51 and then applying the transformation with CCP442. The resulting composite maps were used for building a model of the monomeric complex using O43, and for refinement with REFMAC542 and PHENIX44. The apo-mTORC1 monomer was refined to 3.0 Å with weak secondary structure restraints (SSR) generated by ProSMART52, and the RHEB-mTORC1 monomer to 3.4 Å with tighter SSRs. RHEB was built based on the published structure36. Validation refinement was done as described34. To refine the dimeric complexes, two copies of the three focused maps of each complex were aligned on the corresponding C2 maps as above. The six resulting maps for each complex were then combined with the composite sfcalc option of REFMAC5 to construct the high-resolution structure factors as described for the monomers. The dimeric apo-mTORC1 and RHEB-mTORC1 models were refined against these structure factors using tight non-crystallographic symmetry (n.c.s.) restraints for the positions and B-factors of the atoms.
In vitro kinase assays
In vitro kinase assays were performed as described23, except reaction duration was 20 minutes. Briefly, reactants were assembled in a buffer of 25 mM HEPES, pH 7.4, 100 mM NaCl, 10 mM MgCl2, 2 mM DTT, 5 % (v/v) glycerol and allowed to incubate on ice for 5 minutes. Reactions were started by the addition of cold ATP (0.5 mM final concentration except for Extended Data Fig. 8f, which were added at the indicated concentrations) supplemented with 2 to 4 μCi [γ-32P] ATP (6000 Ci/mmol, Perkin-Elmer) per reaction. Reactions with RHEB-GTPγS or RHEB-GDP were supplemented with 200 μM of the corresponding nucleoside. For reactions with short peptide substrates (Fig. 1g), 0.01% Triton X-100 (Sigma) was added to reduce non-specific interaction of peptides with the test tube. Reactions with peptide substrates were resolved on 16% or 19% Tricine-urea-SDS-PAGE gels53. The Ki values of PRAS40 fragments inhibiting S6K1367–404 phosphorylation shown in Figure 3a were calculated using the IC50 values, the 10 μM S6K1367–404 substrate concentration, and the 430 μM KM of this substrate peptide (Extended Data Fig. 1b) according to the competitive inhibitor equation Ki = IC50/(1+([S]/KM)).
In vivo S6K1 phosphorylation assay
HA-S6K1ki wild type and mutants were cloned into the pcDNA3.1(+) vector. HEK-293F cells (Invitrogen) were maintained in DMEM medium with 10% fetal bovine serum (Sigma) at 37°C and 5% CO2. For transfection, cells were seeded into 6-well tissue culture plates, cultured to 70% confluence and exchanged into fresh media one hour prior to transfection. Cells were transfected with 2 μg each of HA-S6K1ki wild type or mutant plasmids using Lipofectamine 2000 (Invitrogen). 48 hours post transfection, cells were lysed in 50 mM Tris-HCl, 150 mM NaCl, 0.5 mM TCEP, 1% Triton-X100, 2 mM EDTA, 50 mM NaF, 10 mM β-glycerolphosphate and 10 mM Na-pyrophosphate, pH 7.5, and 1 tablet each of cOmplete protease inhibitor and PhosSTOP cocktail (Roche). Whole cell extract (W.C.E.) were adjusted to 2 mg ml−1 with lysis buffer and NuPAGE LDS sample loading buffer (Invitrogen), and boiled for 5 minutes. 20 μg W.C.E. were loaded on gel for immunoblotting with anti-HA antibody (Santa Cruz, SC805) or anti-phospho-S6K1 (T389) antibody (Cell signaling, #9205). The immunoblots were quantified by normalizing the anti-phospho-S6K1 signal to the anti-HA signal of each reaction.
ATPase assay
The ATP hydrolysis assays were set up similarly as the in vitro kinase assays, except without mTOR substrates. To vary the final ATP concentrations, cold ATP was serially diluted and supplemented with [γ-32P] ATP (4 μCi per reaction). The reaction was initiated by mixing the ATP with the enzyme (10 μl total volume), incubated for 20 minutes at 30°C, and stopped by adding 10 μl of 2 M formic acid. 2 μl of each reaction was then spotted on a PEI Cellulose TLC plate (Millipore), developed in 1 M formic acid and 0.5 M LiCl, dried, and quantified by phosphorimaging.
Fluorescence Polarization
FITC-labeled TOS peptides were purchased from Peptide 2.0 Inc. Peptides were quantified by A495 by >20 fold dilution in 10 mM Tris-HCl, pH 8.0, using an extinction coefficient of 75,000 cm−1M−1. A series of 60 μL binding conditions using serially diluted protein with 20 nM FITC-labeled TOS peptides were set up in buffer consisting of 10 mM Tris-HCl, 100 mM NaCl, 2.5% Glycerol and 1 mM TCEP, pH 8.0. Each binding condition was set up in triplicates and equilibrated at room temperature for 15 min. The fluorescence anisotropy measurements were taken with a Cary Eclipse Fluorescence Spectrophotometer with automated polarization accessory (Agilent Technologies), using 485 nm excitation (5 nm slit) and 512 nm emission (10 nm slit) wavelengths, and G factor of 1.5111. The apparent dissociation constants (Kd) values were obtained by fitting the data to a one-site binding model, by minimizing the sum of square of the differences.
Data availability
The cryo-EM maps, including the three focused reconstruction maps and the structure factors of their composite map used in model refinement, and the refined atomic models have been deposited with the Electron Microscopy Data Bank and the Protein Data Bank with accession numbers EMDB-7087 and PDB 6BCX for apo-mTORC1 and EMDB-7086 and PDB 6BCU for RHEB-mTORC1. The coordinates and structure factors of the FRB-S6K1 complex (5WBH), atRaptor (5WBI), atRaptor-4EBP1TOS (5WBJ), atRaptor-S6K1TOS (5WBK), atRaptor-PRAS40TOS (5WBL), mTORΔN-mLST8-PRAS40173–256 (5WBU), mTORΔN-mLST8-PRAS40114–207 (5WBU) have been deposited with the Protein Data Bank.
Extended Data
Extended Data Table.
atRaptor | atRaptor-4EBP199–118 | atRaptor-S6K11–14 | atRaptor- PRAS40124–139 | |
---|---|---|---|---|
Data collection | ||||
Space group | P212121 | P212121 | P212121 | P212121 |
Cell dimensions | ||||
a, b, c (Å) | 89.1, 112.6, 134.1 | 89.1, 113.1, 153.5 | 89.1, 113.1, 152.9 | 89.1, 113.1, 151.8 |
α, β, γ (°) | 90, 90, 90 | 90, 90, 90 | 90, 90, 90 | 90, 90, 90 |
Resolution (Å) | 50-3.00 (3.11-3.00) | 80-3.00 (3.11-3.00) | 80-3.10 (3.21-3.10) | 80-3.35 (3.47-3.35) |
Rsym | 0.086 (0.533) | 0.082 (0.800) | 0.084 (0.819) | 0.088 (0.758) |
Rpim | 0.057 (0.362) | 0.035 (0.307) | 0.035 (0.331) | 0.042 (0.377) |
I/σI | 19.8 (2.6) | 19 (1.53) | 19.2 (1.75) | 22.3 (2.3) |
CC-1/2 | (0.713) | (0.801) | (0.846) | (0.770) |
Completeness (%) | 97.4 (97.3) | 98.7 (99.6) | 99.2 (99.4) | 98.5 (97.6) |
Redundancy | 3.1 (3.0) | 6.2 (7.2) | 6.5 (6.9) | 5.1 (4.8) |
Refinement | ||||
Resolution (Å) | 20-3.0 | 20-3.0 | 20-3.11 | 20-3.35 |
No. reflections | 24264 | 27720 | 25050 | 19535 |
Rwork/Rfree (%) | 24.4/27.8 | 22.8/27.4 | 21.6/25.6 | 21.0/26.5 |
No. atoms | ||||
Protein | 8310 | 8378 | 8332 | 8331 |
Ligand/ion | 0 | 0 | 0 | 0 |
Water | 0 | 0 | 0 | 0 |
B factors | ||||
Protein | 75.3 | 102.4 | 104.4 | 115.3 |
Ligand/ion | – | – | – | – |
Water | – | – | – | – |
R.m.s. deviations | ||||
Bond lengths (Å) | 0.008 | 0.008 | 0.008 | 0.008 |
Bond angles (°) | 1.356 | 1.371 | 1.348 | 1.355 |
Values in parentheses are for highest-resolution shell. All datasets were collected from a single crystal each.
Supplementary Material
Acknowledgments
We thank the staff of the Advanced Photon Source, the NYSBC Simons Electron Microscopy Center, the HHMI Cryo-EM facility, the MSKCC Cryo-EM facility and Subangstrom LLC for help with data collection. Supported by HHMI and National Institutes of Health grant CA008748.
Footnotes
Author Contributions: H.Y., B.L. and N.P.P. designed the experiments, determined the mTOR structures and wrote the manuscript. N.P.P. and X.J. determined the Raptor crystal structures and biochemical constants. H.Y., H.J.Y., M.M., A.Y. and A.D. carried out the mTOR enzyme assays and biochemical experiments.
Author Information: The authors declare no competing financial interests.
SI Guide
Supplementary Information: Supplementary Discussion; Supplementary Data
This file contains the Supplementary Discussion sections about steady state kinetic constants, about other mTORC1 substrates in Figure 1g, about Raptor structure, about mTORC1 structure, and about apo-mTORC1 conformational flexibility. The file also contains Supplementary Figure 1 showing the source data (autoradiograms and immunoblots of gels) for the Main text and Extended Data figures.
References
- 1.Ben-Sahra I, Manning BD. mTORC1 signaling and the metabolic control of cell growth. Current opinion in cell biology. 2017;45:72–82. doi: 10.1016/j.ceb.2017.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Bar-Peled L, Sabatini DM. Regulation of mTORC1 by amino acids. Trends in cell biology. 2014;24:400–406. doi: 10.1016/j.tcb.2014.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Dibble CC, Cantley LC. Regulation of mTORC1 by PI3K signaling. Trends in cell biology. 2015;25:545–555. doi: 10.1016/j.tcb.2015.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Laplante M, Sabatini DM. mTOR signaling in growth control and disease. Cell. 2012;149:274–293. doi: 10.1016/j.cell.2012.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Grabiner BC, et al. A diverse array of cancer-associated MTOR mutations are hyperactivating and can predict rapamycin sensitivity. Cancer discovery. 2014;4:554–563. doi: 10.1158/2159-8290.CD-13-0929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Keith CT, Schreiber SL. PIK-related kinases: DNA repair, recombination, and cell cycle checkpoints. Science. 1995;270:50–51. doi: 10.1126/science.270.5233.50. [DOI] [PubMed] [Google Scholar]
- 7.Kim DH, et al. mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery. Cell. 2002;110:163–175. doi: 10.1016/s0092-8674(02)00808-5. [DOI] [PubMed] [Google Scholar]
- 8.Hara K, et al. Raptor, a binding partner of target of rapamycin (TOR), mediates TOR action. Cell. 2002;110:177–189. doi: 10.1016/s0092-8674(02)00833-4. [DOI] [PubMed] [Google Scholar]
- 9.Kim DH, et al. GbetaL, a positive regulator of the rapamycin-sensitive pathway required for the nutrient-sensitive interaction between raptor and mTOR. Mol Cell. 2003;11:895–904. doi: 10.1016/s1097-2765(03)00114-x. [DOI] [PubMed] [Google Scholar]
- 10.Aylett CH, et al. Architecture of human mTOR complex 1. Science. 2016;351:48–52. doi: 10.1126/science.aaa3870. [DOI] [PubMed] [Google Scholar]
- 11.Saucedo LJ, et al. Rheb promotes cell growth as a component of the insulin/TOR signalling network. Nat Cell Biol. 2003;5:566–571. doi: 10.1038/ncb996. [DOI] [PubMed] [Google Scholar]
- 12.Stocker H, et al. Rheb is an essential regulator of S6K in controlling cell growth in Drosophila. Nat Cell Biol. 2003;5:559–565. doi: 10.1038/ncb995. [DOI] [PubMed] [Google Scholar]
- 13.Sancak Y, et al. PRAS40 is an insulin-regulated inhibitor of the mTORC1 protein kinase. Mol Cell. 2007;25:903–915. doi: 10.1016/j.molcel.2007.03.003. [DOI] [PubMed] [Google Scholar]
- 14.Vander Haar E, Lee SI, Bandhakavi S, Griffin TJ, Kim DH. Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Nat Cell Biol. 2007;9:316–323. doi: 10.1038/ncb1547. [DOI] [PubMed] [Google Scholar]
- 15.Wang L, Harris TE, Roth RA, Lawrence JC., Jr PRAS40 regulates mTORC1 kinase activity by functioning as a direct inhibitor of substrate binding. J Biol Chem. 2007;282:20036–20044. doi: 10.1074/jbc.M702376200. [DOI] [PubMed] [Google Scholar]
- 16.Oshiro N, et al. The proline-rich Akt substrate of 40 kDa (PRAS40) is a physiological substrate of mammalian target of rapamycin complex 1. J Biol Chem. 2007;282:20329–20339. doi: 10.1074/jbc.M702636200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Burnett PE, Barrow RK, Cohen NA, Snyder SH, Sabatini DM. RAFT1 phosphorylation of the translational regulators p70 S6 kinase and 4E-BP1. Proc Natl Acad Sci U S A. 1998;95:1432–1437. doi: 10.1073/pnas.95.4.1432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Ma XM, Blenis J. Molecular mechanisms of mTOR-mediated translational control. Nat Rev Mol Cell Biol. 2009;10:307–318. doi: 10.1038/nrm2672. [DOI] [PubMed] [Google Scholar]
- 19.Hsu PP, et al. The mTOR-regulated phosphoproteome reveals a mechanism of mTORC1-mediated inhibition of growth factor signaling. Science. 2011;332:1317–1322. doi: 10.1126/science.1199498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Kang SA, et al. mTORC1 phosphorylation sites encode their sensitivity to starvation and rapamycin. Science. 2013;341:1236566. doi: 10.1126/science.1236566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Schalm SS, Fingar DC, Sabatini DM, Blenis J. TOS motif-mediated raptor binding regulates 4E-BP1 multisite phosphorylation and function. Curr Biol. 2003;13:797–806. doi: 10.1016/s0960-9822(03)00329-4. [DOI] [PubMed] [Google Scholar]
- 22.Nojima H, et al. The mammalian target of rapamycin (mTOR) partner, raptor, binds the mTOR substrates p70 S6 kinase and 4E-BP1 through their TOR signaling (TOS) motif. J Biol Chem. 2003;278 doi: 10.1074/jbc.C200665200. [DOI] [PubMed] [Google Scholar]
- 23.Yang H, et al. mTOR kinase structure, mechanism and regulation. Nature. 2013;497:217–223. doi: 10.1038/nature12122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Yang H, et al. 4.4 A Resolution Cryo-EM structure of human mTOR Complex 1. Protein & cell. 2016;7:878–887. doi: 10.1007/s13238-016-0346-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Choi J, Chen J, Schreiber SL, Clardy J. Structure of the FKBP12-rapamycin complex interacting with the binding domain of human FRAP. Science. 1996;273:239–242. doi: 10.1126/science.273.5272.239. [DOI] [PubMed] [Google Scholar]
- 26.Yu Y, et al. Phosphoproteomic analysis identifies Grb10 as an mTORC1 substrate that negatively regulates insulin signaling. Science. 2011;332:1322–1326. doi: 10.1126/science.1199484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Roczniak-Ferguson A, et al. The transcription factor TFEB links mTORC1 signaling to transcriptional control of lysosome homeostasis. Sci Signal. 2012;5:ra42. doi: 10.1126/scisignal.2002790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Michels AA. MAF1: a new target of mTORC1. Biochemical Society transactions. 2011;39:487–491. doi: 10.1042/BST0390487. [DOI] [PubMed] [Google Scholar]
- 29.Peterson TR, et al. mTOR complex 1 regulates lipin 1 localization to control the SREBP pathway. Cell. 2011;146:408–420. doi: 10.1016/j.cell.2011.06.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Marcotrigiano J, Gingras AC, Sonenberg N, Burley SK. Cap-dependent translation initiation in eukaryotes is regulated by a molecular mimic of eIF4G. Mol Cell. 1999;3:707–716. doi: 10.1016/s1097-2765(01)80003-4. [DOI] [PubMed] [Google Scholar]
- 31.Ginalski K, Zhang H, Grishin NV. Raptor protein contains a caspase-like domain. Trends Biochem Sci. 2004;29:522–524. doi: 10.1016/j.tibs.2004.08.006. [DOI] [PubMed] [Google Scholar]
- 32.Scheres SH, Chen S. Prevention of overfitting in cryo-EM structure determination. Nature methods. 2012;9:853–854. doi: 10.1038/nmeth.2115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Scheres SH. Processing of Structurally Heterogeneous Cryo-EM Data in RELION. Methods in enzymology. 2016;579 doi: 10.1016/bs.mie.2016.04.012. [DOI] [PubMed] [Google Scholar]
- 34.Brown A, et al. Tools for macromolecular model building and refinement into electron cryo-microscopy reconstructions. Acta Crystallogr D Biol Crystallogr. 2015;71:136–153. doi: 10.1107/S1399004714021683. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Baretic D, Berndt A, Ohashi Y, Johnson CM, Williams RL. Tor forms a dimer through an N-terminal helical solenoid with a complex topology. Nature communications. 2016;7:11016. doi: 10.1038/ncomms11016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Yu Y, et al. Structural basis for the unique biological function of small GTPase RHEB. J Biol Chem. 2005;280:17093–17100. doi: 10.1074/jbc.M501253200. [DOI] [PubMed] [Google Scholar]
- 37.Klink TA, Kleman-Leyer KM, Kopp A, Westermeyer TA, Lowery RG. Evaluating PI3 kinase isoforms using Transcreener ADP assays. Journal of biomolecular screening. 2008;13:476–485. doi: 10.1177/1087057108319864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Wagle N, et al. Activating mTOR mutations in a patient with an extraordinary response on a phase I trial of everolimus and pazopanib. Cancer discovery. 2014;4:546–553. doi: 10.1158/2159-8290.CD-13-0353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Ohne Y, et al. Isolation of hyperactive mutants of mammalian target of rapamycin. J Biol Chem. 2008;283:31861–31870. doi: 10.1074/jbc.M801546200. [DOI] [PubMed] [Google Scholar]
- 40.Urano J, et al. Point mutations in TOR confer Rheb-independent growth in fission yeast and nutrient-independent mammalian TOR signaling in mammalian cells. Proc Natl Acad Sci U S A. 2007;104:3514–3519. doi: 10.1073/pnas.0608510104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Otwinowski Z, Minor W. Processing of X-ray diffraction data collected in oscillation mode. Methods in enzymology. 1997;276:307–326. doi: 10.1016/S0076-6879(97)76066-X. [DOI] [PubMed] [Google Scholar]
- 42.Winn MD, et al. Overview of the CCP4 suite and current developments. Acta Crystallogr D Biol Crystallogr. 2011;67:235–242. doi: 10.1107/S0907444910045749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Jones TA, Zou JY, Cowan SW, Kjeldgaard M. Improved methods for building protein models in electron density maps and the location of errors in these models. Acta crystallographica Section A, Foundations of crystallography. 1991;47(Pt 2):110–119. doi: 10.1107/s0108767390010224. [DOI] [PubMed] [Google Scholar]
- 44.Adams PD, et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr D Biol Crystallogr. 2010;66:213–221. doi: 10.1107/S0907444909052925. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Suloway C, et al. Automated molecular microscopy: the new Leginon system. Journal of structural biology. 2005;151:41–60. doi: 10.1016/j.jsb.2005.03.010. [DOI] [PubMed] [Google Scholar]
- 46.Zheng SQ, et al. MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nature methods. 2017;14:331–332. doi: 10.1038/nmeth.4193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Rohou A, Grigorieff N. CTFFIND4: Fast and accurate defocus estimation from electron micrographs. Journal of structural biology. 2015;192:216–221. doi: 10.1016/j.jsb.2015.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Scheres SH. RELION: implementation of a Bayesian approach to cryo-EM structure determination. Journal of structural biology. 2012;180:519–530. doi: 10.1016/j.jsb.2012.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Kimanius D, Forsberg BO, Scheres SH, Lindahl E. Accelerated cryo-EM structure determination with parallelisation using GPUs in RELION-2. eLife. 2016;5 doi: 10.7554/eLife.18722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Scheres SH. Beam-induced motion correction for sub-megadalton cryo-EM particles. eLife. 2014;3:e03665. doi: 10.7554/eLife.03665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Pettersen EF, et al. UCSF Chimera–a visualization system for exploratory research and analysis. Journal of computational chemistry. 2004;25:1605–1612. doi: 10.1002/jcc.20084. [DOI] [PubMed] [Google Scholar]
- 52.Nicholls RA, Fischer M, McNicholas S, Murshudov GN. Conformation-independent structural comparison of macromolecules with ProSMART. Acta Crystallogr D Biol Crystallogr. 2014;70:2487–2499. doi: 10.1107/S1399004714016241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Schagger H. Tricine-SDS-PAGE. Nature protocols. 2006;1:16–22. doi: 10.1038/nprot.2006.4. [DOI] [PubMed] [Google Scholar]
Extended Data References
- 1.Yang H, et al. mTOR kinase structure, mechanism and regulation. Nature. 2013;497:217–223. doi: 10.1038/nature12122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Choi J, Chen J, Schreiber SL, Clardy J. Structure of the FKBP12-rapamycin complex interacting with the binding domain of human FRAP. Science. 1996;273:239–242. doi: 10.1126/science.273.5272.239. [DOI] [PubMed] [Google Scholar]
- 3.Baker NA, Sept D, Joseph S, Holst MJ, McCammon JA. Electrostatics of nanosystems: application to microtubules and the ribosome. Proc Natl Acad Sci U S A. 2001;98:10037–10041. doi: 10.1073/pnas.181342398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Brown A, et al. Tools for macromolecular model building and refinement into electron cryo-microscopy reconstructions. Acta Crystallogr D Biol Crystallogr. 2015;71:136–153. doi: 10.1107/S1399004714021683. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Kim DH, et al. mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery. Cell. 2002;110:163–175. doi: 10.1016/s0092-8674(02)00808-5. [DOI] [PubMed] [Google Scholar]
- 6.Grabiner BC, et al. A diverse array of cancer-associated MTOR mutations are hyperactivating and can predict rapamycin sensitivity. Cancer discovery. 2014;4:554–563. doi: 10.1158/2159-8290.CD-13-0929. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The cryo-EM maps, including the three focused reconstruction maps and the structure factors of their composite map used in model refinement, and the refined atomic models have been deposited with the Electron Microscopy Data Bank and the Protein Data Bank with accession numbers EMDB-7087 and PDB 6BCX for apo-mTORC1 and EMDB-7086 and PDB 6BCU for RHEB-mTORC1. The coordinates and structure factors of the FRB-S6K1 complex (5WBH), atRaptor (5WBI), atRaptor-4EBP1TOS (5WBJ), atRaptor-S6K1TOS (5WBK), atRaptor-PRAS40TOS (5WBL), mTORΔN-mLST8-PRAS40173–256 (5WBU), mTORΔN-mLST8-PRAS40114–207 (5WBU) have been deposited with the Protein Data Bank.