Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Feb 7.
Published in final edited form as: Mol Cell. 2019 Jan 3;73(3):446–457.e6. doi: 10.1016/j.molcel.2018.11.017

Multisite phosphorylation of S6K1 directs a kinase phospho-code that determines substrate selection

Abul Arif 1,2,, Jie Jia 1, Belinda Willard 3, Xiaoxia Li 4, Paul L Fox 1,*
PMCID: PMC6415305  NIHMSID: NIHMS1517579  PMID: 30612880

SUMMARY

Multisite phosphorylation of kinases can induce on-off or graded regulation of catalytic activity; however, its influence on substrate specificity remains unclear. Here we show multisite phosphorylation of ribosomal protein S6 kinase 1 (S6K1) alters target selection. Agonist-inducible phosphorylation of glutamyl-prolyl tRNA synthetase (EPRS) by S6K1 in monocytes and adipocytes requires not only canonical phosphorylation at Thr389 by mTORC1, but also phosphorylation at Ser424 and Ser429 in the C-terminus by cyclin-dependent kinase 5 (Cdk5). S6K1 phosphorylation at these additional sites induces a conformational switch, and is essential for high-affinity binding and phosphorylation of EPRS, but not canonical S6K1 targets, e.g., ribosomal protein S6. Unbiased proteomic analysis identified additional targets phosphorylated by multisite phosphorylated S6K1 in insulin-stimulated adipocytes, namely, coenzyme A synthase, lipocalin 2, and cortactin. Thus, embedded within S6K1 is a target-selective kinase phospho-code that integrates signals from mTORC1 and Cdk5 to direct an insulin-stimulated, post-translational metabolon determining adipocyte lipid metabolism.

Graphical Abstract

Arif et al report that phosphorylation of the important metabolism-controlling kinase, S6K1, at two sites near the protein terminus induces its phosphorylation of multiple targets related to lipid metabolism. These insulin-stimulated phosphorylation events in adipocytes (fat cells) might contribute to the known influence of S6K1 on obesity.

graphic file with name nihms-1517579-f0001.jpg

INTRODUCTION

Protein phosphorylation is a widespread, post-translational event determining metabolic and immunologic responses to a host of stimuli including normal and pathological stresses (Pawson and Scott, 2005). The pervasive significance of phosphorylation in humans is evidenced by the identification of over 500 kinases that drive phosphorylation, 150 phosphatases that reverse phosphorylation, and 250,000 non-redundant phosphorylation sites in over 20,000 proteins (Alonso et al., 2004; Hornbeck et al., 2012; Khoury et al., 2011; Manning et al., 2002). Protein phosphorylation effectively increases proteome complexity to an extent far exceeding the diversity imparted by the genome (Walsh et al., 2005). Despite intensive investigation of the catalytic mechanisms and consequences of protein phosphorylation, the determinants that define kinase-substrate relationships, and direct target specificity are poorly understood (Miller and Turk, 2018; Ubersax and Ferrell, 2007). Critical determinants, with notable exceptions, include (1) consensus sequences on the target protein in close proximity to the phospho-acceptor site and recognized by the kinase catalytic domain, and (2) distal docking sites on the kinase and substrate that facilitate their transient interaction (Goldsmith et al., 2007; Holland and Cooper, 1999; Mok et al., 2010). Phosphorylation of a kinase, generally regulates its activity, usually inducing activation, but occasionally causing inactivation, and can direct cellular localization (Johnson and Lewis, 2001; Ubersax and Ferrell, 2007). Kinases are often modified by both auto- and trans-phosphorylation, and occasionally at multiple nearby sites (Beenstock et al., 2016; Cohen, 2000; Endicott et al., 2012).

We have investigated stimulus-dependent phosphorylation of glutamyl-prolyl tRNA synthetase (EPRS) in myeloid cells and in adipocytes (Arif et al., 2018). In mouse and human monocytes and macrophages, interferon (IFN)-γ induces phosphorylation of EPRS at Ser999 in the linker domain that joins the catalytic domains in this unique bifunctional tRNA synthetase (Arif et al., 2009). Phosphorylated EPRS is released from its usual residence in the multi-tRNA synthetase complex (MSC) to join with other proteins to form the GAIT (IFN-γ-activated inhibitor of translation) complex that directs translational silencing of select inflammation-related transcripts (Arif et al., 2018). In adipocytes, insulin induces EPRS phosphorylation at the same Ser residue, and upon release from the MSC, P-EPRS binds and transports fatty acid transport-1 (FATP1) to the plasma membrane for long-chain fatty acid (LCFA) uptake and enhanced triglyceride synthesis (Arif et al., 2017). The physiological significance of Ser999 phosphorylation was evidenced by low fat mass and extended life-span in homozygous mice carrying phospho-deficient Ser999-to-Ala mutations (Arif et al., 2017).

Investigation of the upstream events responsible for EPRS Ser999 phosphorylation in macrophages revealed cytosolic p70 ribosomal protein S6 kinase polypeptide 1 (RPS6KB1, also known as S6K1) as the proximal kinase. The evidence included: (i) failure to phosphorylate EPRS following siRNA-mediated knockdown of S6K1 and in bone marrow-derived macrophages from S6K1−/− mice, (ii) in vitro phosphorylation of EPRS by recombinant S6K1 or by S6K1 immunoprecipitated from IFN-γ-activated monocytic cells, and (iii) stimulus-dependent interaction of EPRS with S6K1 (Arif et al., 2017). S6K1 is activated by signature phosphorylation at Thr389 by mammalian target of rapamycin complex 1 (mTORC1); the mTORC1-S6K1 kinase axis is central to anabolic pathways (Dibble and Cantley, 2015; Ma and Blenis, 2009; Ruvinsky et al., 2005; Saxton and Sabatini, 2017). Unexpectedly, IFN-γ-stimulated phosphorylation of EPRS at Ser999 required not only mTORC1 and S6K1, but also cyclin-dependent kinase 5 (Cdk5) and its obligate activator, Cdk5R1 (p35) (Arif et al., 2011; Arif et al., 2017). Metabolic or genetic interactions between Cdk5 and mTOR pathways have not been reported. Cdk5/p35 was once thought to be neuron-specific, but recent studies show its presence and activity in macrophages and adipocytes, and its dysregulation contributes to tumorigenesis and age-related pathologies (Dhavan and Tsai, 2001; Pozo and Bibb, 2016).

Here, we elucidate a previously unrecognized mechanism of kinase activation and target selection. We show that IFN-γ treatment of myeloid cells and insulin treatment of adipocytes induce Cdk5-dependent phosphorylation of two Ser residues in the carboxyl terminus domain (CTD) of S6K1, in addition to “classical” mTORC1-dependent phosphorylation of Thr389. Multisite phosphorylated S6K1 exhibits an altered conformation, and binds and phosphorylates EPRS. Global, proteomic analysis identified additional non-classical targets, namely, coenzyme A synthase (COASY), cortactin (CTTN) and lipocalin-2 (LCN2). Thus, S6K1 exhibits a target-selective phospho-code that controls a post-translational metabolon influencing adipocyte lipid metabolism.

RESULTS

Cdk5 is Required for S6K1-mediated Phosphorylation of EPRS but not Canonical Targets

We investigated the unexpected requirement of Cdk5, in addition to mTORCI and S6K1, for EPRS Ser999 phosphorylation (Arif et al., 2011; Arif et al., 2017). We considered the possibility that Cdk5 was specifically required for activation of mTORC1 or S6K1 in IFN-γ-activated monocytic cells. Cdk5 was inhibited by siRNA-mediated knockdown or by the inhibitor roscovitine in U937 cells. Following stimulation with IFN-γ, phosphorylation of mTORC1 and S6K1, and their downstream targets, was determined by immunoblot. Cdk5 inhibition did not suppress phosphorylation of mTOR at activation sites Ser2448 and Ser2481, nor did it inhibit phosphorylation of the mTORC1 target and S6K1 activation site, Thr389 (Figure 1A and Supplementary Figure S1A). Moreover, phosphorylation of several S6K1 targets, i.e., ribosomal protein S6 (RPS6), eukaryotic translation initiation factor 4B (eIF4B), and eukaryotic elongation factor 2 kinase (eEF2K), were unaffected by Cdk5 inhibition. In contrast, Cdk5 knockdown or inhibition completely blocked EPRS phosphorylation at Ser999. Virtually identical results were observed in IFN-γ-stimulated human peripheral blood monocytes (PBM) and in insulin-stimulated 3T3-L1 adipocytes (Supplementary Figure S1B,C). To establish the specific role of S6K1 in the observed differential target phosphorylation, S6K1 was immunoprecipitated from IFN-γ-activated monocytic cells subjected to Cdk5 inhibition, and incubated with recombinant Ser”9-containing EPRS linker and RPS6. S6K1 isolated from IFN-γ-activated cells phosphorylated both EPRS linker and RPS6 as detected by [γ-32P]ATP labeling; however, S6K1 isolated from Cdk5-inhibited cells phosphorylated RPS6, but not EPRS linker (Figure 1B). We investigated the role of Cdk5 in determining differential target specificity under multiple agonist-stimulated cell conditions. U937 cells, differentiated 3T3-L1 adipocytes, and HepG2 cells were treated with an array of mTORC1 agonists, and phosphorylation of EPRS and RPS6 was determined. In the same cell lysates, Cdk5 activity was measured by immunoprecipitation and in vitro phosphorylation of a Cdk5 substrate peptide. Only those agonists that induced Cdk5 activity, i.e., IFN-γ and lipopolysaccharide (LPS) in U937 cells, insulin and LPS in 3T3-L1 adipocytes, also induced EPRS Ser999 phosphorylation (Supplementary Figure S1C). In contrast, induction of S6 phosphorylation was seen in all cells with nearly all agonists. Together, these results show that activation of S6K1 by mTORC1 alone is not sufficient to phosphorylate EPRS in IFN-γ-stimulated myeloid cells or insulin-stimulated adipocytes, and suggest that an additional Cdk5-dependent mechanism of S6K1 activation is essential for differential phosphorylation of targets by S6K1.

Figure 1. Cdk5 suppression inhibits phosphorylation of EPRS Ser999, but not canonical S6K1 targets.

Figure 1.

(A) Immunoblot analysis of S6K1 target phospho-sites in Cdk5-suppressed and control (cont.) human U937 monocytes. Cells were transfected with control or siRNA targeting Cdk5 and allowed to recover for 24 hr. Cells were serum-depleted for 2 hr and treated with interferon- γ (IFN-γ; 500 units/mL/107 cells) for 8 or 24 hr. For pharmacologic inhibition, serum-depleted cells were treated with IFN-γ for 30 min and then with roscovitine (10 μM) for an additional 7.5 and 23.5 hr. Lysates were analyzed by immunoblot.

(B) Immunocomplex kinase analysis shows Cdk5 suppression inhibits phosphorylation of EPRS, but not RPS6. S6K1 was immunoprecipitated from Cdk5-suppressed, IFN-γ-treated U937 cells. The immunoprecipitates were incubated with purified recombinant, His-tagged Ser886-to-Ala (S886A) EPRS linker (i.e., bearing only Ser999 phospho-acceptor site) and RPS6 in presence of [γ-32P]ATP, and phosphorylation determined by PAGE and autoradiography.

Cdk5-mediated Phosphorylation of S6K1 at Ser424 and Ser429 in the C-terminus is Required for Differential Target Phosphorylation

We considered the possibility that Cdk5 alters S6K1 target selection by phosphorylation at sites distinct from Thr389. Consistent with this mechanism, Cdk5 knockdown reduced IFN-γ-stimulated S6K1 phosphorylation as shown by 32P-labeling of cells, followed by immunoprecipitation and autoradiography (Figure 2A). Cdk5, is a Pro-directed Ser/Thr kinase (Arif et al., 2011), and the major effect of knockdown was reduced Ser phosphorylation as detected by immunoblot with Ser-, Thr-, and Tyr-specific antibodies. S6K1 primarily exists in two isoforms, a nuclear p85S6K1 isoform (α1, bearing 23-amino acid N-terminus nuclear localization sequence), and a shorter, cytosolic p70S6K1 isoform (α2, termed here as S6K1) generated by utilization of an alternate ATG initiation site (Figure 2B) (Magnuson et al., 2012). S6K1 consists of an N-terminal domain (NTD) containing the TOR signaling (TOS) motif, kinase domain (KD), linker containing sites critical for catalytic activity including the signature mTORC1 phospho-site, Thr389, and C-terminal domain (CTD). The CTD contains a cluster of phospho-sites bearing the Ser/Thr-Pro motif, and is thought to contribute to S6K1 catalytic activity (Hou et al., 2007; Magnuson et al., 2012; Mukhopadhyay et al., 1992). However, S6 phosphorylation by S6K1 is only minimally influenced by removal of the CTD, or by phospho-defective or -mimetic mutations (Dennis et al., 1998; Ferrari et al., 1992; Weng et al., 1998). We therefore considered the possibility that Cdk5 phosphorylates CTD sites to re-direct substrate specificity of S6K1. Recombinant EPRS linker and RPS6 were subjected to in vitro phosphorylation by insect cell-derived, full-length S6K1 (including the upstream nuclear localization sequence, S6K1−23–502) and S6K1 lacking the CTD (S6K1−23–398). Deletion of the CTD from S6K1 specifically decreased phosphorylation of EPRS, but not RPS6 (Figure 2B). The ability of full-length, insect cell-derived S6K1 to phosphorylate EPRS linker is likely due to phosphorylation of sites in the CTD (Kozma et al., 1993), however, it was less effective towards EPRS substrate than endogenous S6K1 immunoprecipitated from IFN-γ-activated monocytic cells.

Figure 2. Cdk5-mediated Ser phosphorylation of S6K1 in the C-terminus directs EPRS phosphorylation.

Figure 2.

(A) Cdk5 knockdown reduces S6K1 Ser phosphorylation. Cells transfected with siRNA targeting Cdk5 (or non-targeting control) were labeled with 32P-orthophosphate in phosphate-free DMEM medium in absence or presence of IFN-γ for 4 hr. 32P-incorporation in S6K1 was determined by immunoprecipitation with anti-S6K1 antibody and autoradiography. Unlabeled S6K1 immunopreciptates also were analyzed by immunoblot using Ser, Thr, and Tyr phospho-specific antibodies.

(B) Role of C-terminal domain in EPRS phosphorylation. (Top) Domain map and critical phospho-sites in p85 and p70 S6K1. Phospho-site numbering is based on p70 S6K1. Arrows indicate alternate ATG initiation site for generation of S6K1 isoforms. CTD, C-terminal domain; KD, kinase domain; NLS, nuclear localization signal; NTD, N-terminal domain. (Bottom) S6K1 immunoprecipitate from IFN-γ-treated U937 cells (S6K1-IP), recombinant, active full-length (rS6K1FL, 1-502 aa plus NLS), and an S6K1 truncate lacking the CTD (rS6K11-398, 1-398 aa plus NLS) were used for in vitro phosphorylation of S886A EPRS linker and RPS6. Amount of each S6K1 form was added to give equivalent activity determined by in vitro phosphorylation of S6K1-specific peptide substrate KRRRLASLR. Aliquots were spotted on phosphocellulose P-81 filters, and 32P incorporation determined by scintillation counting (mean ± SEM; n = 3 experiments).

We investigated whether Cdk5-mediated phosphorylation of specific CTD sites contributed to altered substrate selection. Cdk5-dependent phosphorylation of sites that conform to the Ser/Thr-Pro recognition motif was determined in lysates from IFN-γ-treated cells using available site-specific phospho-antibodies against S6K1. Cdk5 knockdown significantly reduced Ser411 and Ser424 phosphorylation, whereas phosphorylation of canonical S6K1 activation sites, i.e., Thr229, Ser371, and Thr389 (as well as Ser418 and Thr421), were unaffected (Figure 3A and Supplementary Figure S2A). Motif-specific antibodies further validated Cdk5-dependent phosphorylation of Ser sites in S6K1 conforming to the Ser-Pro motif (Supplementary Figure S2B). Ser429 was considered a candidate Cdk5-dependent phosphorylation site based on sequence, but a phospho-antibody was not available. Thus, Cdk5-dependent phosphorylation of Ser411, Ser424, or Ser429, or a combination of these sites, might serve as a potential signal dictating altered S6K1 substrate specificity. To investigate specific phospho-site(s) required for EPRS phosphorylation, the three candidate Ser residues, alone and in combination, were mutated to Ala in Myc-tagged-pcDNA3 constructs containing S6K1 cDNA lacking the 3’UTR. U937 cells were co-transfected with mutant or wild-type S6K1 constructs, and with 3’UTR-specific siRNA to knock down endogenous S6K1, and then treated with IFN-γ. S6K1 constructs bearing Ala mutations at either Ser424 or Ser429 markedly reduced phosphorylation of EPRS at Ser999, but not RPS6 at Ser235/236 (Figure 3B). S6K1 bearing a Thr389-to-Ala mutation did not phosphorylate either target, consistent with its essential role in mediating S6K1 activation (Dennis et al., 1998). To further investigate the requirement for combinatorial phosphorylation of both Ser424 and Ser429 sites, a gain-of-function experiment was performed. Phospho-mimetic Ser-to-Asp S6K1 mutants were constructed at the three candidate sites, alone and in combination, in the context of a phospho-mimetic mutation at Thr389 to constitutively activate the canonical mTORC1 target site, even in the absence of agonist. U937 cells were co-transfected with Myc-tagged wild-type and mutated S6K1 cDNA, and with 3’UTR-specific siRNA to knock down endogenous S6K1. Phosphorylation of EPRS Ser999 was observed only when S6K1 contained phospho-mimetic mutations at both Ser424 and Ser429; RPS6 phosphorylation at Ser235/236 was independent of CTD phospho-mimetic mutations (Figure 3C). Essentially identical effects of S6K1 phospho-deficient and - mimetic mutations on differential target phosphorylation were observed in insulin-treated 3T3-L1 adipocytes. (Supplementary Figure S2C,D). To determine if Cdk5 directly phosphorylates Ser424 and Ser429, wild-type and mutated recombinant S6K1 were subjected to in vitro phosphorylation by Cdk5 immunoprecipitated from lysates from IFN-γ-treated U937 cells. In vitro phosphorylation of Cdk5-specific target peptide (PKTPKKAKKL) confirmed kinase activity of the immunocomplex (Figure 3D, right). Phosphorylation of S6K1 bearing single Ala mutants at Ser424 or Ser429 was decreased by about 50%, and by about 80-90% in double Ser424/Ser429 mutants (Figure 2D, left). In vitro phosphorylation of S6K1 mutants by recombinant Cdk5/p35 (i.e., Cdk5 activated by its binding partner, p35) gave similar results, establishing direct phosphorylation of S6K1 by Cdk5 at critical target specificity-determining sites. For brevity we designate the S6K1 proteoform phosphorylated at Thr389, Ser424, and Ser429 as S6K1*.

Figure 3. Ser424 and Ser429 phosphorylation in the S6K1 CTD, in addition to Thr389 phosphorylation, is required for EPRS Ser999 phosphorylation.

Figure 3.

(A) Cdk5 knockdown inhibits phosphorylation of specific Ser sites in S6K1 CTD. Schematic of S6K1 phospho-sites including critical mTORC1 site Thr389 and putative Cdk5 targets in CTD (top). Phosphorylation in cell lysates from IFN-γ-treated control and Cdk5-knockdown U937 cells was analyzed by immunoblot using phospho-specific antibodies (bottom).

(B) Ectopic expression of phospho-defective Ser424- and Ser429-to-Ala S6K1 mutants blocks EPRS Ser999 phosphorylation. Endogenous S6K1 expression was suppressed by transfection of 3’UTR-specific siRNA in U937 cells. Cells were co-transfected with C-terminus Myc-tagged wild-type (WT) S6K1 ORF (1-502 minus 23-aa NLS), phospho-mimetic (Thr389-to-Glu, T389E) and phospho-defective Thr389-to-Ala (T389A) mutant S6K1, and single, double, and triple Ser-to-Ala point mutants as shown. Cells were treated with IFN-γ, and phosphorylation of EPRS and RPS6 analyzed by immunoblot.

(C) Ectopic co-expression of Ser424-to-Asp (S424D) and Ser429-to-Asp (S429D) phospho-mimetic S6K1 mutants phosphorylate EPRS Ser999 in absence of IFN-γ. U937 cells were transfected as in (B) but with phospho-mimetic, instead of phospho-defective mutants.

(D) Cdk5 phosphorylates Ser424 and Ser429 in the S6K1 CTD. Cdk5 was isolated by Immunoprecipitation of IFN-γ-treated U937 cell lysates with anti-Cdk5 antibody. Activity was determined by in vitro phosphorylation of Cdk5-specific peptide substrate, PKTPKKAKKL, in the presence of [γ-32P]ATP (right). Recombinant wild-type and mutant S6K1 bearing N-terminus His- and C-terminus Myc-tag were generated by in vitro translation in wheat germ extract and purified by Ni-affinity chromatography utilizing the His-tag (Myc-S6K1). Immunoprecipitated Cdk5 was used for in vitro phosphorylation of Myc-S6K1 in the presence of [γ32P]ATP. (left). Phosphorylation was quantitated densitometrically and expressed as percentage of γ32-P incorporation compared to wild-type control. Purified Myc-S6K1 was detected by silver stain. Likewise, recombinant Cdk5 activated by p35 (Cdk5/p35) also was used to phosphorylate Myc-tagged S6K1 (bottom).

Cellular Functions of EPRS Phosphorylated by S6K1*

We investigated the cellular functions of EPRS phosphorylated by S6K1* by determining its ability to reconstitute known monocyte and adipocyte functions. We previously showed that IFN-γ-mediated phosphorylation of EPRS in monocytic cells induces binding to NS1-associated protein (NSAP1), phosphorylated ribosomal protein L13a (RPL13a), and glyceraldehyde 3-phosphate dehydrogenase (GAPDH), to form the GAIT complex that inhibits translation of multiple inflammation-related transcripts (Arif et al., 2018; Ray and Fox, 2007; Sampath et al., 2004). Moreover, the phosphorylated linker domain of EPRS is sufficient for formation of a functional complex with the other GAIT constituents (Arif et al., 2009; Jia et al., 2008). Myc-tagged S6K1 constructs bearing phospho-mimetic mutations were transfected into U937 cells, and collected by immunoprecipitation with anti-Myc tag antibody. The immunocomplexed kinases exhibited the expected substrate specificity as shown by EPRS Ser999 phosphorylation by S6K1 bearing T389E mutation plus S424D and S429D mutations, and RPS6 Ser235/236 phosphorylation by all S6K1 forms bearing a T389E mutation (Figure 4A). Purified EPRS linker, after in vitro phosphorylation by the S6K1 immunocomplexes, was incubated with other GAIT complex constituents, i.e., phospho-RPL13a, NSAP1, and GAPDH, in a wheat germ extract. GAIT element-bearing reporter mRNA, i.e., luciferase (luc) upstream of the ceruloplasmin (Cp) GAIT mRNA element, was subjected to in vitro translation in the extract. Translation was markedly inhibited by S6K1 forms bearing T389E, S424D, and S429D mutations; moderate inhibition was observed by S6K1 forms bearing the T389E mutation (Figure 4B).

Figure 4. Isolated S6K1 triple-phosphomimetic at Thr389, Ser424, and Ser429 (S6K1*) exhibits EPRS-selective specificity, and dictates EPRS-related cell functions.

Figure 4.

(A) Immunocomplexed triple phospho-mimetic S6K1* bearing T389E, S424D and S429D mutations exhibits selectivity for EPRS Ser999 phosphorylation. Myc-tagged S6K1 bearing phospho-mimetic mutations was transfected into U937 cells, and immunoprecipitated with anti-Myc antibody. Immunoprecipitates were incubated with purified recombinant His-tagged S886A EPRS linker, or RPS6, in presence of ATP, and phosphorylation determined by immunoblot with phospho-specific antibodies.

(B) EPRS linker phosphorylated by S6K1* reconstitutes GAIT complex activity. His-tagged S886A EPRS linker was in vitro phosphorylated by immunocomplexed Myc-S6K1, purified, and pre-incubated with other GAIT complex constituents, i.e., phospho-RPL13a, NSAP1, and GAPDH. Reconstituted complexes were added to a wheat germ extract for in vitro translation of luciferase (Luc)-ceruloplasmin (Cp) GAIT element reporter transcript in the presence of [35S]Met. T7 gene 10 RNA was co-translated as specificity control. Luc translation was quantified by densitometry, normalized by T7 gene 10, and expressed as percent of control (mean ± SEM, n = 3 experiments).

(C) Triple phospho-mimetic S6K1* activates EPRS linker to stimulate fatty acid uptake in adipocytes. siRNA targeting S6K1 3’UTR and Myc-tagged S6K1 ORFs were co-transfected into 3T3-L1 adipocytes. Non-transfected cells treated with insulin (100 nM) for 4 hr served as positive control. Lysates were analyzed by immunoblot. 5 × 104 cells were incubated with bodipy-dodecanoate (C12) for 4 h, and fatty acid uptake assayed by relative fluorescence (485 nm excitation and 515 nm emission) after 30-min. Uptake was expressed as fold-change relative to untreated, non-transfected control cells (mean ± SEM, n = 3 experiments in duplicate).

We previously reported that following insulin stimulation of adipocytes, EPRS phosphorylated by mTORC1/S6K1 binds FATP1 and transports it to the plasma membrane for enhanced fatty acid uptake (Arif et al., 2017). To investigate the specific phosphorylation requirement for this S6K1 activity, differentiated 3T3-L1 adipocytes were co-transfected with Myc-tagged wild-type and mutated S6K1 and with Flag-tagged EPRS linker. As in monocytic cells, Ser999 phosphorylation of EPRS linker was observed only when S6K1 contained phospho-mimetic mutations at both Ser424 and Ser429, but RPS6 phosphorylation only required the T389E mutation (Figure 4C). Transfection of S6K1 containing both S424D and S429D mutations at both Ser424 and Ser429, nearly doubled uptake of bodipy-C12, a fluorescently-labeled LCFA reporter. In a positive control experiment, insulin treatment of adipocytes induced phosphorylation of both EPRS linker and RPS6, and also doubled fatty acid uptake.

Altered Binding Properties and Conformation of S6K1*

A reciprocal co-immunoprecipitation experiment showed that IFN-γ induces S6K1 binding to EPRS in monocytic cells (Arif et al., 2017). Following co-transfection of 3T3-L1 adipocytes with Flag-tagged EPRS linker and Myc-tagged S6K1 bearing phospho-mimetic mutations, an interaction between S6K1 and EPRS linker was observed only when S6K1 contained phospho-mimetic mutations at both Ser424 and Ser429 (Figure 5A); a similar result was seen in transfected U937 monocytes (Supplementary Figure S3A). To investigate whether this interaction is direct, binding of recombinant S6K1 forms with EPRS linker was determined by surface plasmon resonance (SPR) spectrometry. Recombinant, His-tagged wild-type S6K1 and S6K1 bearing phospho-mimetic mutations were generated by in vitro translation in wheat germ extract. Wild-type S6K1 was also phosphorylated by incubation with a lysate from insulin-treated adipocytes, repurified, and used as phosphorylated wild-type (P-WT) S6K1 control. EPRS linker was immobilized on a CM5 SPR sensor chip, and S6K1 proteins flowed over the chip for 5 min to permit binding, and then stopped to allow dissociation (Figure 5B). The association and dissociation curves were used to calculate the equilibrium dissociation constants, KD. Importantly, the binding of S6K1 mutant T389E/S424D/S429D (i.e., S6K1*) to EPRS linker exhibited a dissociation constant (56 nM) nearly two orders of magnitude lower than wild-type S6K1 (5.0 μM). The binding affinity of EPRS to S6K1* and to P-WT S6K1 control (40.0 nM) were similar, verifying that the putative phospho-mimetic mutations are highly mimetic and that other phosphorylation sites in S6K1 do not contribute substantially to binding. Finally, the T389E mutant did not exhibit high-affinity binding to EPRS indicating essential requirement for phosphorylation of the CTD sites.

Figure 5. S6K1* exhibits altered conformation compared to S6K1, facilitating high-affinity binding to EPRS for site-selective Ser999 phosphorylation.

Figure 5.

(A) Triple phospho-mimetic S6K1* binds and phosphorylates EPRS linker in absence of insulin. 3’UTR-specific S6K1 siRNA, C-terminus Myc-tagged S6K1 ORFs, and Flag-tagged EPRS linker were co-transfected into 3T3-L1 adipocytes. Lysates were subjected to co-immunoprecipitation with anti-Myc antibody to probe Flag-EPRS linker/Myc-S6K1 interaction. EPRS linker Ser999 phosphorylation was determined by immunoblot.

(B) SPR analysis of binding of S6K1 forms to EPRS linker. Wild-type (WT) and mutant S6K1 proteins (1.25 nmol) were flowed over EPRS linker immobilized on a Biacore CM5 sensor chip. Binding was measured as relative response units. Smoothed association and dissociation curves determined by BIAevaluation software was used to calculate binding affinities expressed as dissociation constants, Kd.

(C) Susceptibility of S6K1 forms to limited proteolysis. To establish appropriate limited proteolysis condition, purified Myc-S6K1* (75 ng) bearing T389E, S424D, and S429D mutations was treated with prolyl endopeptidase for 20 min, and probed with anti-Myc antibody (right). Myc-S6K1 bearing phospho-defective and phospho-mimetic mutations were cleaved with prolyl endopeptidase (40 ng, 20 min), and probed with anti-Myc antibody. Myc-S6K1 mutants generated by in vitro translation in wheat germ extract, and purified by Ni-affinity, were detected by silver staining.

(D) Conformational alteration of S6K1* determined by antibody affinity. WT and S6K1 mutants (1.25 nmol) were flowed over anti-S6K1 antibody raised against human S6K1 C-terminus (amino acids 490-502) immobilized on a CM5 Biacore sensor chip. Binding was determined as in (B).

The specific high-affinity binding of EPRS to S6K1* suggests that the phosphorylated CTD sites are an essential part of the EPRS binding site, or that phosphorylation of the sites induces a conformational shift that exposes the binding site. To investigate the latter possibility, we subjected recombinant forms of C-terminus Myc-tagged S6K1 to limited proteolysis with prolyl endopeptidase after determining appropriate conditions for controlled cleavage (Figure 5C, right). Wild-type S6K1 was almost completely degraded by the treatment, however, forms containing both S424D and S429D mutations (in the context of T389E mutation) exhibited a unique, protected ~37 kDa fragment (Figure 5C, left). This result suggests that CTD phosphorylation alters protein conformation as evidenced by decreased susceptibility of the C-terminus half of S6K1* to degradation. In an independent assessment of S6K1 conformation we determined the binding affinity of an antibody specific for the C-terminal 13 amino acids of S6K1 (490-502). The binding results nearly paralleled those observed for binding of S6K1 forms to EPRS linker. The antibody binds with higher affinity to S6K1* or wild-type, endogenously phosphorylated S6K1 compared to wild-type, unmodified protein or the Thr389 phospho-mimetic form (Figure 5D). Together these experiments support a mechanism in which phosphorylation of S6K1 Ser424 and Ser429 induces a conformational shift that increases surface exposure of a region near the C-terminus that contains the EPRS linker docking site.

Additional Targets Uniquely Phosphorylated by S6K1*

A proteomic approach was taken to identify additional S6K1* targets based on the premise that these targets, like EPRS, will exhibit selective, high-affinity binding to S6K1*. HEK293T cells were transfected with Myc-tagged, wild-type or triple phospho-mimetic S6K1*, or pcDNA3 vector control (Figure 6A). Following pull-down with anti-Myc antibody and PAGE, the lanes were segregated into 8 regions subjected to elution, trypsinization, and LC-mass spectrometry. Peptides derived from 605, 999, and 846 proteins (combined total of 958 proteins) were detected in the control, S6K1 and S6K1* lanes, respectively. To select the protein sub-set that specifically binds S6K*, the data were filtered using a cut-off of at least 14 spectral counts in the S6K1* sample, and a S6K1*-to-S6K1 ratio of at least 4. The screen identified 10 candidate S6K1*-binding proteins including EPRS (Figure 6A, Supplementary Table 1). Binding of the candidates to S6K1 was investigated by immunoprecipitation with anti-S6K1 antibody in lysates from insulin-treated 3T3-L1 adipocytes; specific binding to S6K1* was supported by inhibition following knockdown with Cdk5 siRNA (and with S6K1 knockdown as control). Of the nine newly identified candidates, three were shown to bind S6K1 in a Cdk5-dependent way, namely, bifunctional coenzyme A (CoA) synthase (COASY), neutrophil gelatinase-associated lipocalin (LCN2), and Src substrate cortactin (CTTN) (Figure 6B). RPS6, a canonical S6K1 phosphorylation target, did not stably bind either S6K1 or S6K1*, as shown by comparison to vector control, consistent with the transient nature of typical interactions between kinases and their substrates (de Oliveira et al., 2016). Inability to validate several candidates, e.g., BPIFB1 and WDR77, is inconclusive as it might be due to low adipocyte expression, low antibody sensitivity, or both. Phosphorylation of the candidates was determined by 32P-labeling followed by immunoprecipitation. EPRS, COASY, LCN2, and CTTN all exhibited, insulin-stimulated phosphorylation that was blocked by siRNA targeting S6K1 and Cdk5, consistent with S6K1* targets (Figure 6C). RPS6 was inducibly phosphorylated, but Cdk5-independence indicates S6K1* activity is not required. siRNA-mediated knockdown of the S6K1* targets in insulin-treated 3T3-L1 adipocytes reduced their expression and phosphorylation verifying the specificity of the antibodies (Supplementary Figure S4). Immunoprecipitation followed by immunoblot with phospho-specific antibodies revealed insulin induced phosphorylation primarily on Ser on all four S6K1* targets (Figure 6D).

Figure 6. Proteomic identification of S6K1* targets.

Figure 6.

(A) Proteomic analysis identified candidate proteins that preferentially bind S6K1*. Myc-tagged wild-type S6K1, S6K1* bearing phospho-mimetic T389E, S424D, and S429D mutations, and pcDNA3 vector control (Cont.) were transfected in HEK293T cells, and lysates probed by immunoblot (left). Lysates were subjected to co-immunoprecipitation with anti-Myc antibody, and resolved on 4-20% Tris-MOPS-SDS-PAGE followed by imperial blue stain (center). Stained gels were divided into eight areas, excised, and subjected to in-gel tryptic digest. Digests were subjected to LC-MS/MS, and data searched using all high-energy, collision-induced dissociation (HCD) spectra against the human SwissProtKB database with Sequest and Mascot programs. Proteins that specifically bind S6K* with a cut-off of ≥14 spectral counts in the S6K1* lane, and a S6K1*/S6K1 ratio ≥5 (right). RPS6 is included to show lack of specific interaction with S6K1*. #, undefined (division by 0).

(B) Validation of Cdk5-dependent interaction of S6K1* with candidate targets. 3T3-L1 adipocytes transfected with siRNA targeting S6K1 or Cdk5 were treated with insulin for 4 hr. Lysates were probed by immunoblot before (right) and after (left) co-immunoprecipitation with anti-S6K1 antibody.

(C) Validation of Cdk5- and S6K1-dependent phosphorylation of candidate targets of S6K1*. 3T3-L1 adipocytes were labeled with 32P-orthophosphate in phosphate-free DMEM medium ± insulin for 4 hr. 32P-incorporation was determined by immunoprecipitation and autoradiography (left). Immunoprecipitated protein determined by immunoblot (right).

(D) Identification of Ser phosphorylation of S6K1* targets. 3T3-L1 adipocytes were immunoprecipitated with antibodies targeting COASY, CTTN, and LCN2, and probed with phospho-Ser/Thr/Tyr antibodies. EPRS Ser999 phosphorylation is shown as control.

In addition to determining stimulus- and kinase-dependent requirements, we directly assessed the requirement of multisite-phosphorylated S6K1* for target protein interaction and phosphorylation. 3T3-L1 adipocytes were transfected with Myc-tagged wild-type, triple phospho-mimetic (S6K1*), and T389E forms of S6K1, and binding to targets was determined by immunoprecipitation with anti-Myc antibody. EPRS, COASY, CTTN, and LCN2 all bound S6K1* but not wild-type or T389E forms of S6K1 (Figure 7A). Neither RPS6 nor isoleucyl tRNA synthetase (IARS) bound any S6K1 forms. Phosphorylation of all candidates by S6K1*, but not by wild-type or T389E S6K1, was shown by metabolic labeling with 32P followed by immunoprecipitation and autoradiography (Figure 7B). Together, these experiments verify COASY, CTTN, and LCN2, in addition to the founding family member EPRS, as authentic S6K1* targets in insulin-stimulated adipocytes, possibly defining a coordinately regulated, post-translational metabolon influencing lipid synthesis (Figure 7C).

Figure 7. S6K1* binds and phosphorylates EPRS, COASY, CTTN, and LCN2.

Figure 7.

(A) Ectopically expressed S6K1* binds targets in cells. cDNA encoding Myc-tagged wild-type, triple phospho-mimetic (S6K1*), and T389E S6K1 were transfected into 3T3-L1 adipocytes. Lysates were subjected to co-immunoprecipitation and analyzed by immunoblot. Detection with anti-IARS and -RPS6 antibodies served as negative controls.

(B) Ectopically expressed S6K1* phosphorylates targets in cells. S6K1 forms were transfected into 3T3-L1 adipocytes as above, and labeled with 32P-orthophosphate in phosphate-free DMEM medium for 4 hr. Lysates were immunoprecipitated with target-specific antibodies and phosphorylation determined by autoradiography.

(C) Schematic of insulin-induced noncanonical activation of S6K1, the kinase phospho-code, and downstream influence on lipid metabolism.

DISCUSSION

Our studies establish an unprecedented molecular mechanism in which disparate upstream signaling cascades are integrated by combinatorial, multisite phosphorylation of S6K1, which in turn dictates substrate selection. Inducible phosphorylation of EPRS, CTTN, COASY, LCN2 by differentially phosphorylated S6K1 is evidence for a target-selective phospho-code embedded in the kinase. Canonical mTORC1 agonists phosphorylate S6K1 at Thr389, and together with phosphorylation events in the kinase and linker domains (i.e., Thr229 and Ser371), activate S6K1 to induce phosphorylation of canonical targets that in turn drive global stimulation of translation (Figure 7C) (Ma and Blenis, 2009; Ruvinsky et al., 2005). In contrast, insulin stimulation of adipocytes and IFN-γ stimulation of myeloid cells, activate both the mTORC1 and Cdk5/p35 pathways, inducing phosphorylation of S6K1 at Thr389, and additional phosphorylation events at CTD sites Ser424 and Ser429. Multisite phosphorylated S6K1* phosphorylates canonical S6K1 targets, but also additional targets including COASY, CTTN, LCN2, and EPRS.

The mTORC1-S6K1 axis integrates intracellular and extracellular signals and controls nutrient-driven metabolic pathways with a well-established link to inflammation (Fontana et al., 2010; Hotamisligil and Erbay, 2008). Our results show an unusual contribution of Cdk5 in addition to mTORC1 in integrating inputs from IFN-γ and insulin for substrate-selective activation of S6K1*. Cdk5 is a ubiquitously expressed atypical member of the Cdk family of kinases with no known role in cell cycle progression. Cdk5 was considered to be a neuron-specific kinase due to abundant expression of its obligate activators, p35 and p39, in neurons (Dhavan and Tsai, 2001; Pozo and Bibb, 2016). However, context- and stimulus- dependent activation of Cdk5 has been detected in non-neuronal cells, including monocytes and adipocytes (Arif et al., 2011; Chen and Studzinski, 2001; Lalioti et al., 2009; Pareek et al., 2010). In adipocytes, insulin- and obesity-induced activation of Cdk5 is critical for glucose uptake and stimulation of a diabetogenic expression program (Banks et al., 2015; Lalioti et al., 2009; Li et al., 2011). Cdk5/p35 is also critical for insulin secretion by pancreatic islets in response to glucose challenge (Ubeda et al., 2004; Wei et al., 2005). Likewise, Cdk5 is linked to multiple inflammation-related responses such as differentiation of leukemic cells to monocytes, induction of T-cell-mediated inflammatory disorders, and anti-tumor immunity via regulation of the cell death ligand, PD-L1 (Chen et al., 2004; Dorand et al., 2016; Pareek et al., 2010). Our results revealing Cdk5- and mTORC1-dependent S6K1 activation link the kinase trio to the regulation of pro-inflammatory gene expression and metabolic activities in myeloid and fat cells, respectively.

Cdk5- and mTORC1-directed multisite phosphorylation drives activation and substrate selectivity of S6K1*. Multisite phosphorylation, defined as phosphorylation of two or more sites in a single polypeptide chain, can modulate or fine-tune protein activity, localization, stability, and interaction partners (Cohen, 2000; Holmberg et al., 2002). In the case of a single kinase inducing phosphorylation of multiple sites, then a simple input can be amplified into a complex response. As an example, temporally regulated multisite phosphorylation of ELK-1 by ERK directs rapid target activation and subsequent inactivation (Mylona et al., 2016). In contrast, if multiple sites are phosphorylated by distinct kinases, then integration of upstream signaling pathways is a key system feature (Ferrell, 1996; Nash et al., 2001; Park et al., 2006). For example, the tumor suppressor proteins p53 and retinoblastoma have multiple phospho-sites that combinatorially generate a range of responses by stimulus- and context-dependent interactions with multiple downstream effectors (Murray-Zmijewski et al., 2008; Rubin, 2013).

Multisite phosphorylation of multiple kinases, including S6K1, can control enzymatic activity via orthosteric or allosteric regulation (Cohen, 2000; Magnuson et al., 2012; Nussinov et al., 2012). However, multisite phosphorylation has not been reported to influence target selection, which is generally determined by the intrinsic property of the substrates (Kang et al., 2013; Miller and Turk, 2018). In response to diverse inputs, S6K1 can be phosphorylated at eight or more well-defined sites (Magnuson et al., 2012). In addition to the signature mTORC1-mediated phosphorylation of Thr389, phosphorylation of Thr229 and Ser371 is considered essential for canonical S6K1 activation (Ma and Blenis, 2009; Moser et al., 1997; Pullen et al., 1998). Stepwise phosphorylation of multiple CTD sites in S6K1 by several proline-directed Ser/Thr kinases including Cdk5 has been suggested to contribute to S6K1 activation, possibly by relieving an inhibitory interaction between N- and C-termini (Ferrari et al., 1992; Hou et al., 2007; Lai et al., 2015; Mukhopadhyay et al., 1992; Papst et al., 1998; Sarker and Lee, 2004; Shah et al., 2003). However, the minimal effect of CTD truncation or site-specific CTD mutations on activation suggests a complex relationship between proline-directed Ser/Thr kinases and S6K1 (Cheatham et al., 1995; Dennis et al., 1998; Edelmann et al., 1996; Magnuson et al., 2012; Mahalingam and Templeton, 1996; Weng et al., 1998). Furthermore, stimulus-dependent activation of Cdk5 in a limited set of non-neuronal cells suggests a specialized role for proline-directed Ser/Thr kinases in mTORC1-S6K1 signaling, and possibly explains the cell-type specificity observed for EPRS phosphorylation (Supplementary Figure S1B)(Arif et al., 2017).

Our findings reveal a new function for multisite phosphorylation of a kinase, endowing S6K1 with a “barcode” that, depending on phosphorylation state of key sites, directs differential target recognition emblematic of kinase plasticity. Phospho-codes have been described for several key regulatory proteins, including, retinoblastoma protein, histones, and p53 tumor suppressor (Murray-Zmijewski et al., 2008; Rubin, 2013; Sims and Reinberg, 2008). Phospho-codes in kinases have been reported, but they generally determine binding partners and signaling pathways, not substrate selection and target post-translational modification as observed for S6K1* (Nussinov et al., 2012). Receptor tyrosine kinases exhibit phospho-codes that determine binding to adaptors without directly influencing target selection or phosphorylation (Olsson et al., 2006). Likewise, insulin-induced multisite phosphorylation of the kinase Akt2 generates distinct temporal signaling profiles but the mechanism, downstream effector, and output cellular response remain undefined (Humphrey et al., 2015). Importantly, the elucidation of a Cdk5- and mTORC1-directed phospho-code embedded in S6K1 establishes a foundation and incentive for discovery of analogous mechanisms in other kinases to specify differential, context-dependent target selection. Moreover, the S6K1 phospho-code provides a conceptual foundation for exploration of kinase inhibitors that inhibit phosphorylation of a selected cluster of targets.

S6K1*-directed EPRS Ser999 phosphorylation is defined not by consensus sequence recognition, but by high-affinity binding to a conformational variant of the kinase. An unbiased, proteomic approach identified additional candidate proteins that stably interact with S6K1*. As for EPRS, three additional proteins, COASY, CTTN, and LCN2, were experimentally validated to be phosphorylated by S6K1* in adipocytes by an insulin- and Cdk5-dependent mechanism. All of these proteins are previously unrecognized targets of S6K1, with the exception of COASY that was observed to bind S6K1, but was not phosphorylated by the kinase (Nemazanyy et al., 2004). Each S6K1* target exhibits activities actually or arguably related to lipid metabolism; however, the specific effects of phosphorylation on their activities are unknown. Particularly significant is the targeting of bifunctional enzyme COASY, which catalyzes the final two reactions in the 5-step pathway converting pantothenate to coenzyme A (Zhyvoloup et al., 2002). For triacylglycerol synthesis, LCFA imported into adipocytes require initial activation by coenzyme A, catalyzed primarily by long-chain coenzyme A synthetase 1 (ACSL1) (Lobo et al., 2009). Phosphorylation of COASY might increase the catalytic rate of one or both of its rate-limiting enzymatic activities, thereby increasing available coenzyme A (Zhyvoloup et al., 2002). Alternatively, phosphorylation could also induce translocation of COASY from its principal residence at the outer mitochondrial membrane to the plasma membrane for improved accessibility to ACSL1 (Zhyvoloup et al., 2003).

Interestingly, ACSL1 is complexed to FATP1, the major insulin-stimulated LCFA importer, which in turn is transported to the plasma membrane by another S6K1* target, phospho-EPRS (Arif et al., 2017; Gargiulo et al., 1999; Richards et al., 2006). These transport process might be facilitated by CTTN, an actin-binding protein required for insulin-stimulated translocation of vesicles containing the Glut4 glucose transporter to the plasma membrane (Nazari et al., 2011; Tunduguru et al., 2017). This mechanism is supported by evidence that Cdk5 is stimulated by insulin and plays a key role in Glut4-mediated glucose uptake and FATP1-mediated long-chain fatty acid uptake in adipocytes (Arif et al., 2017; Lalioti et al., 2009). Possibly, S6K1*-mediated phosphorylation of CTTN enables actin-mediated transport of the LCFA import machinery, i.e., CoASY or EPRS/FATP1, to the plasma membrane. Lastly, LCN2 is a secreted adipokine elevated during obesity and promotes insulin resistance (Yan et al., 2007). Exogenous application of LCN2 can stimulate both LCFA uptake and β-oxidation in adipocytes (Law et al., 2010; Paton et al., 2013). Possibly, insulin-stimulated phosphorylation of LCN2 represses its β-oxidation activity, thereby favoring anabolic pathways promoting triglyceride synthesis. These results suggest that S6K1* likely represents a critical control point of an insulin-stimulated, post-translational metabolon determining lipid metabolism in adipocytes. Future studies will delineate the specific lipid-related functions of the phospho-forms of these S6K1* target proteins, identify additional targets, and determine the respective roles of S6K1 and S6K1* in health and disease.

STAR★METHODS

CONTACT FOR REAGENT AND RESOURCE SHARING

Further information and requests for resources and reagents should be directed to, and will be fulfilled by, the Lead Contact, Paul L. Fox (foxp@ccf.org).

Plasmids and proteins

Recombinant, His-tagged, wild-type and Ser886-to-Ala (S886A) mutant linker proteins spanning Pro683 to Asn1023 of human EPRS were expressed and purified as described (Arif et al., 2011; Arif et al., 2009). Recombinant, active, full-length S6K1 (rS6K1FL, 1502 aa plus NLS) was purchased from R&D Systems and Cell Signaling (Kozma et al., 1993). Full-length human S6K1 cDNA in pCMV6-Entry vector was purchased from Origene and recloned, deleting the 23-aa N-terminus nuclear localization sequence (NLS) and adding an in-frame upstream 6-His tag and downstream C-terminus Myc-tag in pcDNA3. Specific mutations were introduced using appropriate mutation-bearing primers and GENEART Site-Directed Mutagenesis System (Invitrogen). Recombinant, wild-type and mutant NLS-deleted S6K1 with N-terminus His- and C-terminus Myc-tag were generated by in vitro translation using wheat germ extract system (Promega), and purified by Ni-affinity chromatography (Thermo Scientific Pierce).

Cell lines and cell culture

Human U937 monocytic cells (CRL 1593.2) were obtained from American Type Culture Collection (ATCC), and cultured in RPMI 1640 medium and 10% fetal bovine serum (FBS) with penicillin and streptomycin at 37 °C in 5% CO2. Peripheral blood monocytes (PBM) from human blood were isolated by leukapheresis and elutriation under an Institutional Review Board-approved protocol adhering to the guidelines of American Association of Blood Banks. PBMs were cultured in RPMI 1640 medium and 10% FBS with penicillin and streptomycin at 37 °C in 5% CO2. U937 cells and PBMs (107 cells) were incubated with 500 U/ml IFN-γ (R&D Systems) for up to 24 hr as described (Arif et al., 2011; Arif et al., 2017). Human HepG2 hepatoma (HB-8065; ATCC) and mouse 3T3-L1 fibroblasts (CL-173; ATCC) were cultured in high-glucose containing Dulbecco’s modified Eagle’s medium (DMEM; Life Technologies), 10% FBS, 4 mM L-glutamine, and 1× antibiotic/antimycotic solution (Life Technologies) at 37 °C in 5 and 10% CO2, respectively. For differentiation into adipocytes, fibroblasts at 75% confluence were cultured in medium containing DMEM and 10% FBS supplemented with 1× solutions of insulin:dexamethasone:3-isobutyl-1-methylxanthine (Cayman) as described (Arif et al., 2017). After 72 h, the medium was replaced with 10% FBS and DMEM containing insulin and maintained for a week with three medium changes. Adipocytes were maintained in DMEM medium with 10% calf serum and 1× antibiotic/antimycotic for at least 3 days before use. Differentiated adipocytes were serum-deprived for 4 hr followed by treatment with 100 nM insulin (Sigma-Aldrich) for 4 hr. HEK293T cells (CRL-3216; ATCC) were cultured in the high-glucose DMEM medium at 10% FBS, 4 mM L-glutamine and 1× antibiotic/antimycotic at 37 °C in 5% CO2. Cell lysates were prepared using Phosphosafe extraction buffer (Novagen) supplemented with protease and phosphatase inhibitor cocktails.

Transfection procedures

Cells were transfected with endotoxin-free plasmid DNAs or siRNAs (target-specific and scrambled control) using 100 μl of nucleofector solution V (U937 cells) or solution L (differentiated 3T3-L1 adipocytes) from Amaxa nucleofection kit (Lonza) following manufacturer’s protocol. Transfected cells were immediately transferred to pre-warmed Opti-MEM media for 6 hr and then to RPMI 1640 (for U937 cells) and DMEM (for 3T3-L1 adipocytes) containing 10% FBS supplemented with 4 mM L-glutamine, penicillin, streptomycin, and geneticin (G418; 20 μg/ml) for 18 to 24 hr before treatment with insulin and inhibitors.

In vitro kinase and phosphorylation assays

Purified active, recombinant kinases or immunocomplexed kinases were pre-incubated with EPRS S886A linker or RPS6 for 5 min in kinase assay buffer (50 mM Tris-HCl, pH 7.6, 1 mM dithiotheitol, 10 mM MgCl2, 1 mM CaCl2, and phosphatase inhibitor cocktail) (Arif et al., 2011; Arif et al., 2009). Phosphorylation was initiated by addition of 5 μCi [γ-32P]ATP (Perkin-Elmer), and after 15 min was terminated using denaturing Laemmli gel-loading dye and heat denaturation. 32P incorporation was determined by autoradiography after resolution on Tris-glycine SDS-PAGE and fixation in 40% methanol plus 10% acetic acid. For kinase activity assays, 50 μM of synthetic oligopeptide substrates were phosphorylated with 1 μCi [γ-32P]ATP in kinase assay buffer. Equal volumes were spotted onto P81-phosphocellulose squares, washed in 0.5% H3PO4, and 32P incorporation determined by scintillation counting.

Immunocomplex kinase assay

Pre-cleared cell lysates were incubated with specific antibodies for 4 hr in kinase assay buffer, and immunocomplexes captured by incubating with protein A-Sepharose beads for an additional 4 hr. (Arif et al., 2011; Arif et al., 2009). Immunocomplexes were washed three times with kinase assay buffer supplemented with 0.1% Triton X-100. Finally, the immunocomplexes were resuspended in kinase assay buffer and used to phosphorylate EPRS S886 linker, RPS6, or peptide substrate. 32P incorporation into substrates was determined by SDS-PAGE coupled with autoradiography and scintillation counting (Arif et al., 2009; Arif et al., 2017).

Immunoblot analysis

Ccell lysates or immunoprecipitates were denatured in Laemmli sample buffer (Bio-Rad) and resolved on Tris-glycine SDS-PAGE (10, 12, or 15% polyacrylamide) prepared using 37.5:1 acrylamide:bis-acrylamide stock solution (National Diagnostics), or on precast Express Plus Tris-MOPS-SDS PAGE (8, 10, 12, 4-12%, 4-20%; GenScript). After transfer to polyvinyl difluoride membranes, proteins were probed with target-specific antibody, followed by incubation with horseradish peroxidase-conjugated secondary antibody (GE Healthcare), and detected with ECL prime western blotting detection reagent (GE Healthcare). Immunoblots shown are typical of experiments independently done at least three times.

Phosphorylation assay by 32P-metabolic labeling

24 hr post-transfection with siRNAs, U937 and 3T3-L1 cells were incubated with IFN-g and insulin, respectively, in presence of 250 μCi of 32P-orthophosphate (MP Biomedical) in phosphate-free RPMI 1640 medium for 4 hr. 32P-labeled cells were lysed in Phosphosafe extraction buffer (Novagen) supplemented with protease and phosphatase inhibitor cocktails, and subjected to immunoprecipitation. Target 32P-labeled proteins were immunoprecipitated by overnight incubation of pre-cleared lysates (1 mg) with specific antibodies in buffer containing 50 mM Tris-HCl (pH 7.6), 150 mM NaCl, 1% Triton X-100, 1% NP-40, 0.5% sodium deoxycholate, 0.1% SDS, and protease/phosphatase inhibitor cocktails. Immunoprecipitates were captured with Protein A-Sepharose beads (Sigma) for an additional 4 hr. Immunoprecipitated beads were washed three times in detergent-containing buffer (50 mM Tris-HCl, pH 7.6, 150 mM NaCl, and 0.1% Triton X-100) followed by three washes in detergent-free buffer. Finally, immunoprecipitated proteins were suspended in denaturing Laemmli gel-loading dye, resolved on SDS-PAGE, and phosphorylation determined by autoradiography.

In vitro reconstitution of GAIT complex and translation assays

Capped, poly(A)-tailed luciferase (Luc) reporter upstream of the ceruloplasmin (Cp) 3’UTR GAIT element (Luc-Cp GAIT) and T7 gene 10 reporter transcripts were prepared using mMessage mMachine SP6 and T7 kits (Ambion), respectively (Jia et al., 2008). For in vitro reconstitution of GAIT complex, His-tagged S886A EPRS linker (permitting site-specific Ser999 phosphorylation) was phosphorylated by immunocomplexed Myc-S6K1 (wild-type and mutant) in presence of ATP (10 μM), and re-purified using Ni-NTA purification kit (Thermo-Scientific). Purified NSAP1, phospho-L13a, and GAPDH were generated as described (Jia et al., 2008). The GAIT complex was reconstituted by incubating 5 pmol each of purified EPRS linker, NSAP1, GAPDH, and phospho-L13a. The reconstituted complex was incubated with Luc-Cp-GAIT (200 ng) and T7 gene 10 (200 ng) template transcripts, and added to wheat germ extract system (Promega) containing Met-free amino acid mixture and 35S-Met (Perkin-Elmer). Translation of Luc-Cp-GAIT and T7 gene 10 transcripts was visualized by resolution on 10% SDS-PAGE and autoradiography.

Fatty acid uptake assay

Fatty acid uptake was determined using the long-chain fatty acid analog, bodipy-dodecanoic acid and a QBT fatty acid uptake assay (Molecular Devices) (Arif et al., 2017; Liao et al., 2005). Untransfected and Myc-tagged S6K1 transfected 3T3-L1 adipocytes (5 × 104 cells per well in a 96-well plate) were incubated in serum-free Hanks balanced salt solution for 4 h, and then with the fatty acid analog for an additional 4 hr. As a positive control, untransfected adipocytes were treated with the fatty acid analog in the presence of 100 nM insulin. After 30 min, the relative fluorescence was determined at 485 nm excitation and 515 nm emission wavelengths in bottom-read mode (SpectraMax GeminiEM, Molecular Devices).

Limited proteolysis assay

Purified, recombinant, Myc-tagged wild-type and mutant S6K1 (75 ng) were subjected to proteolysis by prolyl endopeptidase (40 to 150 ng; Abcam) for 20 min at 30 °C in 50 mM Tris HCl (pH 7.5) buffer. The reaction was terminated using non-denaturing Laemmli gel-loading dye supplemented with 0.1 M EDTA. Proteolyzed S6K1 was analyzed by SDS PAGE followed by immunoblot with anti-Myc antibody.

Protein-protein interaction analysis by co-immunoprecipitation

Cell lysates were pre-cleared by incubation with protein A-Sepharose beads (Sigma) for 1 hr. Pre-cleared lysates (1 mg) were incubated for 12-16 hr with antibody conjugated to protein A-Sepharose beads in detergent-free buffer containing 50 mM Tris-HCl (pH 7.6), 150 mM NaCl, and EDTA-free protease/phosphatase inhibitor cocktail. Immunoprecipitated beads were washed three times in the same buffer followed by elution with 0.2 M glycine-HCl (pH 2.6), immediately neutralized with 50 mM Tris-HCl (pH 8.5), and suspended in denaturing Laemmli gel-loading dye. Co-immunoprecipitation was also performed by directly incubating the pre-cleared cell lysates (1 mg) with antibody overnight in detergent-free buffer, followed by a 4 hr incubation with protein A-Sepharose beads. The immunoprecipitates were washed as above, suspended in denaturing Laemmli gel-loading dye, and analyzed by Tris-glycine SDS-PAGE and immunoblot.

Protein-protein interaction analysis by SPR

Recombinant N-terminus, His-tagged human EPRS linker protein was purified using Ni-NTA purification kit (Thermo-Scientific). Subsequently, the His-tag was removed using enterokinase cleavage capture kit (Novagen), and re-purified by Ni-NTA column. N-terminus 6-His tagged wild-type and mutant S6K1 were generated by in vitro translation in wheat germ extract and purified by Ni-NTA column. Phosphorylated S6K1 was generated by in vitro phosphorylation of purified S6K1 by lysates from insulin-treated 3T3-L1 adipocytes in presence of ATP (10 μM) as above, followed by re-purification by Ni-NTA chromatography. Purified EPRS linker or anti-S6K1 antibody were immobilized on Biacore sensor chip CM5 using amine-coupling procedures per manufacturer’s instructions. Purified wild-type and mutant His-tagged S6K1 proteins were prepared in HBS-EP (0.01 M HEPES, pH 7.4, 0.15 M NaCl, 3 mM EDTA and 0.005% v/v surfactant P20) buffer (Biacore). Purified S6K1 proteins (1.25 nmol) were flowed over the sensor chip at 10 μl/min for 5 min. After a 2-min dissociation period, the bound proteins were removed with regeneration buffer containing 1.5 M NaCl and 10 mM glycine-HCl, pH 3.0 (30 μl/min, 30 s). Finally the sensor chip was equilibrated by washing with HBS-EP buffer (2.5 min). Kinetics of S6K1 binding (dissociation constant, Kd) to EPRS linker and S6K1 antibody was calculated using association and dissociation curves generated by Biaevaluation software (Biacore).

Proteomic analysis of S6K1*-interacting proteins

HEK293T cells (3 × 105 cells) in triplicate were transfected with 1 μg endotoxin-free DNA encoding Myc-tagged ORFs of S6K1, S6K1* with phospho-mimetic T389E, S424D, and S429D mutations, and pcDNA3 vector, using Fugene 6 transfection reagent (Promega) following manufacturer’s instruction. 36 hr post-transfection, the replicates were pooled and lysed in 50 mM Tris HCl, pH 7.6 by four freeze-thaw cycles, followed by passage several times through a 26-gauge needle, and centrifugation at 5,000 rpm for 5 min at 4 °C. Supernatant (1.5 mg protein) was subjected to co-immunoprecipitation by incubation with anti-Myc tag Sepharose-conjugated mouse mAb (Cell Signaling) for 16 hr. The precipitate was washed 4 times with 50 mM Tris HCl, pH 7.6, 150 mM NaCl containing protease and phosphatase inhibitor cocktail, and denatured in Laemmli sample buffer (Bio-Rad). The samples were resolved by 4-20% GenScript Express Plus Tris-MOPS-SDS PAGE, and protein detected by Imperial blue stain (ThermoFisher). Each lane was divided into eight regions, excised, washed/destained in 50% ethanol, 5% acetic acid, and dehydrated in acetonitrile. Excised regions were reduced with 1,4-dithiotreitol and alkylated with iodoacetamide prior to overnight in-gel digestion by trypsin (5 μl of 10 ng/μl in 50 mM NH4HCO3). Peptides were extracted using 30 μl 50% acetonitrile with 5% formic acid, solvent was evaporated, and peptides resuspended in 1% acetic acid for liquid chromatography-mass spectrometry (LC-MS) analysis using a Fusion Lumos mass spectrometer system (ThermoScientific). Extracts were injected into a reversed-phase capillary chromatography HPLC column (Dionex Acclaim Pepmap 100 C18 LC), and peptides eluted by acetonitrile/0.1% formic acid gradient at a flow rate of 0.3 μl/min with detection by a 2.5 kV microelectrospray ion source. Data-dependent multitask capability of the instrument was used to acquire full-scan mass spectra and generate high-energy, collision-induced dissociation (HCD) spectra to determine peptide molecular weights and amino acid sequences. The HCD spectra were searched against the human SwissProtKB database with Mascot and Sequest programs and peptide and protein validation was performed using the program Scaffold (Proteome Software, Portland, OR) to determine the specific S6K1* interactome.

QUANTIFICATION AND STATISTICAL ANALYSIS

Data obtained from in vitro phosphorylation of peptide substrate, autoradiography, and in vitro translation were analyzed using Student’s t test (GraphPad Prism 5) and results shown as mean ± SEM.

DATA AVAILABILITY

Comprehensive data from the proteomic analysis of S6K1*-interacting proteins is included as a single Excel spreadsheet.

Supplementary Material

2
3

Table S1. Proteomic analysis of S6K1*-interacting proteins, Related to Figure 6.

(A) Summary of results obtained from each lane divided into eight regions. (B) Summary of results by lane type, i.e., control, S6K1 and S6K1*. (C) Comparison of proteins in control, S6K1, and S6K1* lanes.

KEY RESOURCES TABLE

REAGENT or RESOURCE SOURCE IDENTIFIER
Antibodies
Anti-phospho-S6K1-Ser418 Abcam Cat# ab58534;
RRID:AB_883110
Anti-phospho-S6K1-Thr229 Abcam Cat# ab5231;
RRID:AB_304788
Anti-phospho-S6K1-Ser411 Novus Biologicals Cat# NBP1-60941
RRID:AB_11006602
Anti-phospho-S6K1-Ser424 Novus Biologicals Cat# NBP1-60942;
RRID:AB_11019606
Anti-phospho-S6K1-Thr421 Novus Biologicals Cat# NB100-92608;
RRID:AB_1217585
Anti-S6K1 Cell Signaling Technology Cat# 2708;
RRID:AB_390722
Anti-phospho-S6K1-Ser371 Cell Signaling Technology Cat# 9208;
RRID:AB_330990
Anti-phospho-S6K1-Thr421/Ser424 Cell Signaling Technology Cat# 9204;
RRID:AB_2265913
Anti-RPS6 Cell Signaling Technology Cat# 2217
RRID:AB_331355
Anti-phospho-RPS6-Ser235/236 Cell Signaling Technology Cat# 2211
RRID:AB_331679
Anti-eIF4B Cell Signaling Technology Cat# 3592
RRID:AB_2293388
Anti-phospho-eIF4B-Ser422 Cell Signaling Technology Cat# 3591
RRID:AB_2097522
Anti-eEF2K Cell Signaling Technology Cat# 3692
RRID:AB_2231040
Anti-phospho-eEF2K-Ser366 Cell Signaling Technology Cat# 3691
RRID:AB_2097313
Anti-mTOR Cell Signaling Technology Cat# 2972
RRID:AB_330978
Anti-phospho-mTOR-Ser2448 Cell Signaling Technology Cat# 2971
RRID:AB_330970
Anti-phospho-mTOR-Ser2481 Cell Signaling Technology Cat# 2974
RRID:AB_2231885
Anti-P-Ser PXSP or SPXR/K Cell Signaling Technology Cat# 2325
RRID:AB_331820
Anti-P-Thr PXTP motif Cell Signaling Technology Cat# 4391
RRID:AB_331247
Anti-Myc Cell Signaling Technology Cat# 2278
RRID:AB_490778
Anti-Myc tag Sepharose-conjugated Cell Signaling Technology Cat# 3400;
RRID:AB_1281297
Anti-Flag Cell Signaling Technology Cat# 2368;
RRID:AB_2217020
Anti-phospho-Thr Cell Signaling Technology Cat# 9386;
RRID:AB_331239
Anti-phospho-Ser Abcam Cat# ab9332
RRID:AB_307184
Anti-phospho-Tyr Millipore/Calbiochem Cat# 525322
Anti-Cdk5 Santa Cruz Biotechnology Cat# sc-173;
RRID:AB_631224
Anti-β-Actin Santa Cruz Biotechnology Cat# sc-1615;
RRID:AB_630835
Anti-His Santa Cruz Biotechnology Cat# sc-803;
RRID:AB_631655
Anti-COASY Novus Biologicals Cat# NBP2-15934
Anti-BPIFB1 GeneTex Cat# GTX60659
Anti-WDR77 GeneTex Cat# GTX49156
Anti-MIB1 R&D Systems Cat# AF7305;
RRID:AB_11127218
Anti-HP R&D Systems Cat# AF4409;
RRID:AB_1964600
Anti-MPO R&D Systems Cat# MAB3174;
RRID:AB_2250873
Anti-IARS Bethyl Cat# A304-748A;
RRID:AB_2620943
Anti-LCN2 R&D Systems Cat# AF1857;
RRID:AB_355022
Anti-CTTN Santa Cruz Biotechnology Cat# sc-55579;
RRID:AB_831187
Anti-S6K1 (490-502) Abcam Cat# ab14708;
RRID:AB_301427
Anti-EPRS (Ray and Fox, 2007) N/A
Anti-phospho-EPRS-Ser886 (Arif et al., 2009) N/A
Anti-phospho-EPRS-Ser999 (Arif et al., 2009) N/A
Bacterial and Virus Strains
MAX Efficiency Dh5a Competent Cells ThermoFisher Cat# 18258012
One Shot BL21(DE3)pLysS Competent Cells ThermoFisher Cat# C606003
Chemicals, Peptides, and Recombinant Proteins
Recombinant, purified His-tagged, wild-type and Ser886-to-Ala (S886A) mutant linker proteins spanning Pro683 to Asn1023 of human EPRS (Arif et al., 2011)
(Arif et al., 2009)
N/A
S6K1 peptide (KRRRLASLR) substrate SignalChem Cat# S05-58
Cdk5 peptide (PKTPKKAKKL) substrate Santa Cruz Biotechnology Cat# sc-3066
Recombinant, active, full-length S6K1 (rS6K1FL, 1-502 aa plus NLS) R&D Systems; Cell Signaling Technology (Kozma et al., 1993) Cat# 896-KS
Cat# 7684
Recombinant active S6K1 truncate lacking CTD (rS6K11-398, 1-398 aa plus NLS) EMD Millipore Cat# 14-486
Recombinant active Cdk5/p35 EMD Millipore Cat# 14-477
Recombinant full-length RPS6 Abcam Cat# ab91724
NSAP1 protein (Jia et al., 2008) N/A
Phosphorylated-L13a protein (Jia et al., 2008) N/A
GAPDH protein (Jia et al., 2008) Cat# G6019
Amaxa cell line nucleofector kit V Lonza VCA-1003
Amaxa cell line nucleofector kit L Lonza VCA-1005
Roscovitine Calbiochem Cat# 557360
Prolyl endopeptidase Abcam Cat# ab80376
Protein A-Sepharose Sigma-Aldrich Cat# P3391
Insulin Sigma-Aldrich Cat# I9278
LPS Sigma-Aldrich Cat# L8274
IFN-α Pestka Biomedical Lab. Cat# 11200-1
IFN-β Pestka Biomedical Lab. Cat# 11415-1
IFN-γ R&D Systems Cat# 285IF
Insulin (for 3T3-L1 cell culture) Cayman Chemical Cat# 10008979
Dexamethasone (for 3T3-L1 cell culture) Cayman Chemical Cat# 10008980
3-isobutyl-1-methylxanthine (for 3T3-L1 cell culture) Cayman Chemical Cat# 10008978
Phosphosafe extraction buffer (Novagen) EMD Millipore Cat# 71296
Laemmli gel-loading dye Bio-Rad Cat# 1610737
Imperial blue stain ThermoFisher Cat# 24615
[γ-32P]ATP Perkin-Elmer Cat# BLU502A250UC
32P-orthophosphoric acid Perkin-Elmer
MP Biomedicals
Cat# NEX053H005MC
Cat# 016401405
ATP Cell Signaling Technology Cat# 9804
P81-phosphocellulose squares EMD Millipore Cat# 20-134
Protease inhibitor cocktail Roche Diagnostics Cat# 11836170001
Phosphatase inhibitor cocktail ThermoFisher Cat# 78420
Polyvinyl difluoride membranes EMD Millipore Cat# IPVH00010
ECL prime western blotting detection reagent GE Healthcare Cat# RPN2232
Biacore Sensor chip CM5 Biacore Life Sciences (GE Healthcare) Cat# BR100012
ECL-Anti-Rabbit HRP-conjugated secondary antibody GE Healthcare NA9340V
ECL-Anti-Mouse HRP-conjugated secondary antibody GE Healthcare NA931V
[35S]Met Perkin-Elmer NEG709A500UC
Critical Commercial Assays
GENEART Site-Directed Mutagenesis System Life Technologies Cat# A1382
SilverQuest Silver Staining Kit Life Technologies Cat# LC6070
Fugene 6 transfection reagent Promega Cat# E2691
Wheat germ extract system Promega Cat# L4380
6xHis Fusion Protein Spin Purification Kit (Ni- affinity chromatography) ThermoFisher (Pierce) Cat# 78300
EndoFree Plasmid Maxi Kit Qiagen Cat# 12362
Express Plus Tris-MOPS-SDS PAGE GenScript Cat# M00138
QBT fatty acid uptake assay kit Molecular Devices Cat# R8132
mMESSAGE mMachine T7 Transcription Kit Ambion (Jia et al., 2008) Cat# AM1344
mMESSAGE mMachine SP6 Transcription Kit Ambion (Jia et al., 2008) Cat# AM1340
Enterokinase cleavage capture kit (Novagen) EMD Millipore Cat# 69067-3
Experimental Models: Cell Lines
Human: U937 monocytic cells ATCC CRL-1593.2
Human: HepG2 hepatoma ATCC HB-8065
Human: peripheral blood monocytes (PBM) Cleveland Clinic IRB-approved protocol N/A
Human: HEK293T ATCC CRL-3216
Mouse: 3T3-L1 ATCC CL-173
Oligonucleotides
siRNA targeting mouse EPRS
5’-UGAUACGAAGAUCUUCUCAG-3’
5’-GCCUAAAUUAACAGUGGAA-3’
Origene Technologies N/A
siRNA targeting human and mouse S6K1 (3’UTR-specific) Origene Technologies Cat# SR304164
siRNA targeting mouse S6K1 Origene Technologies Cat# SR416706
siRNA targeting mouse CTTN Origene Technologies Cat# SR427044
siRNA targeting mouse LCN2 Origene Technologies Cat# SR405112
siRNA targeting mouse COASY Origene Technologies Cat# SR417447
siRNA targeting mouse CDK5 Santa Cruz Biotechnology Cat# sc-35047
siRNA targeting human Cdk5 Dharmacon Santa Cruz Biotechnology Cat# L-003239-00-0050
Cat# sc-29263
Control siRNA Santa Cruz Biotechnology Cat# sc-37007
Recombinant DNA
Full-length human S6K1 cDNA in pCMV6-Entry Origene Technologies Cat# RC217324
Myc-tagged ORFs of wild -type and various mutant S6K1 (N-terminus 6-His, C-terminus Myc-tag S6K1 without 23-aa N-terminus NLS in pcDNA3) This study N/A
Flag-EPRS linker in pcDNA3 (Arif et al., 2009) N/A
Software and Algorithms
GraphPad Prism 5 www.graphpad.com N/A
BIAevaluation Software www.biacore.com N/A
NIH Image J https://imagej.nih.gov/ij/ N/A

Highlights.

  • Insulin-stimulated phosphorylation of EPRS Ser999 by S6K1 requires mTORCI and Cdk5

  • C-terminus phosphorylation of S6K1 by Cdk5 is required for EPRS phosphorylation

  • S6K1 exhibits a phospho-site dependent, target-selective phospho-code

  • Multisite phosphorylated S6K1 induces a metabolon driving adipocyte lipid metabolism

ACKNOWLEDGEMENTS

This work was supported by NIH grants P01 HL029582, P01 HL076491, and R01 GM086430 (to P.L.F.). A.A. was supported by a Scientist Development Grant from the AHA, National Affiliate. Shared Instrument Grant, S10 OD023436 from the National Institutes of Health (to B.W.) was used to purchase the mass spectrometer. None of the authors have any financial conflict of interest with the information in this manuscript.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

None of the material has been published or is under consideration elsewhere.

SUPPLEMENTAL INFORMATION

Supplemental Information includes four figures and one Excel spreadsheet

DECLARATION OF INTERESTS

The authors declare no competing financial interests.

REFERENCES

  1. Alonso A, Sasin J, Bottini N, Friedberg I, Friedberg I, Osterman A, Godzik A, Hunter T, Dixon J, and Mustelin T (2004). Protein tyrosine phosphatases in the human genome. Cell 117, 699–711. [DOI] [PubMed] [Google Scholar]
  2. Arif A, Jia J, Moodt RA, DiCorleto PE, and Fox PL (2011). Phosphorylation of glutamyl-prolyl tRNA synthetase by cyclin-dependent kinase 5 dictates transcript-selective translational control. Proc. Natl. Acad. Sci. USA 108, 1415–1420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Arif A, Jia J, Mukhopadhyay R, Willard B, Kinter M, and Fox PL (2009). Two-site phosphorylation of EPRS coordinates multimodal regulation of noncanonical translational control activity. Mol. Cell 35, 164–180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Arif A, Terenzi F, Potdar AA, Jia J, Sacks J, China A, Halawani D, Vasu K, Li X, Brown JM, et al. (2017). EPRS is a critical mTORC1-S6K1 effector that influences adiposity in mice. Nature 542, 357–361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Arif A, Yao P, Terenzi F, Jia J, Ray PS, and Fox PL (2018). The GAIT translational control system. Wiley Interdiscip. Rev. RNA 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Banks AS, McAllister FE, Camporez JP, Zushin PJ, Jurczak MJ, Laznik-Bogoslavski D, Shulman GI, Gygi SP, and Spiegelman BM (2015). An ERK/Cdk5 axis controls the diabetogenic actions of PPARgamma. Nature 517, 391–395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Beenstock J, Mooshayef N, and Engelberg D (2016). How do protein kinases take a selfie (autophosphorylate)? Trends Biochem. Sci 41, 938–953. [DOI] [PubMed] [Google Scholar]
  8. Cheatham L, Monfar M, Chou MM, and Blenis J (1995). Structural and functional analysis of pp70S6k. Proc. Natl. Acad. Sci. USA 92, 11696–11700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Chen F, and Studzinski GP (2001). Expression of the neuronal cyclin-dependent kinase 5 activator p35Nck5a in human monocytic cells is associated with differentiation. Blood 97, 3763–3767. [DOI] [PubMed] [Google Scholar]
  10. Chen F, Wang Q, Wang X, and Studzinski GP (2004). Up-regulation of Egr1 by 1,25-dihydroxyvitamin D3 contributes to increased expression of p35 activator of cyclin-dependent kinase 5 and consequent onset of the terminal phase of HL60 cell differentiation. Cancer Res 64, 5425–5433. [DOI] [PubMed] [Google Scholar]
  11. Cohen P (2000). The regulation of protein function by multisite phosphorylation–a 25 year update. Trends Biochem. Sci 25, 596–601. [DOI] [PubMed] [Google Scholar]
  12. de Oliveira PS, Ferraz FA, Pena DA, Pramio DT, Morais FA, and Schechtman D (2016). Revisiting protein kinase-substrate interactions: Toward therapeutic development. Sci. Signal 9, re3. [DOI] [PubMed] [Google Scholar]
  13. Dennis PB, Pullen N, Pearson RB, Kozma SC, and Thomas G (1998). Phosphorylation sites in the autoinhibitory domain participate in p70(s6k) activation loop phosphorylation. J. Biol. Chem 273, 14845–14852. [DOI] [PubMed] [Google Scholar]
  14. Dhavan R, and Tsai LH (2001). A decade of CDK5. Nat. Rev. Mol. Cell Biol 2, 749–759. [DOI] [PubMed] [Google Scholar]
  15. Dibble CC, and Cantley LC (2015). Regulation of mTORC1 by PI3K signaling. Trends Cell Biol 25, 545–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Dorand RD, Nthale J, Myers JT, Barkauskas DS, Avril S, Chirieleison SM, Pareek TK, Abbott DW, Stearns DS, Letterio JJ, et al. (2016). Cdk5 disruption attenuates tumor PD-L1 expression and promotes antitumor immunity. Science 353, 399–403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Edelmann HM, Kuhne C, Petritsch C, and Ballou LM (1996). Cell cycle regulation of p70 S6 kinase and p42/p44 mitogen-activated protein kinases in Swiss mouse 3T3 fibroblasts. J. Biol. Chem 271, 963–971. [DOI] [PubMed] [Google Scholar]
  18. Endicott JA, Noble ME, and Johnson LN (2012). The structural basis for control of eukaryotic protein kinases. Annu. Rev. Biochem 81, 587–613. [DOI] [PubMed] [Google Scholar]
  19. Ferrari S, Bannwarth W, Morley SJ, Totty NF, and Thomas G (1992). Activation of p70s6k is associated with phosphorylation of four clustered sites displaying Ser/Thr-Pro motifs. Proc. Natl. Acad. Sci. USA 89, 7282–7286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Ferrell JE Jr. (1996). Tripping the switch fantastic: how a protein kinase cascade can convert graded inputs into switch-like outputs. Trends Biochem. Sci 21, 460–466. [DOI] [PubMed] [Google Scholar]
  21. Fontana L, Partridge L, and Longo VD (2010). Extending healthy life span--from yeast to humans. Science 328, 321–326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Gargiulo CE, Stuhlsatz-Krouper SM, and Schaffer JE (1999). Localization of adipocyte long-chain fatty acyl-CoA synthetase at the plasma membrane. J. Lipid Res 40, 881–892. [PubMed] [Google Scholar]
  23. Goldsmith EJ, Akella R, Min X, Zhou T, and Humphreys JM (2007). Substrate and docking interactions in serine/threonine protein kinases. Chem. Rev 107, 5065–5081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Holland PM, and Cooper JA (1999). Protein modification: docking sites for kinases. Curr. Biol 9, R329–331. [DOI] [PubMed] [Google Scholar]
  25. Holmberg CI, Tran SE, Eriksson JE, and Sistonen L (2002). Multisite phosphorylation provides sophisticated regulation of transcription factors. Trends Biochem. Sci 27, 619–627. [DOI] [PubMed] [Google Scholar]
  26. Hornbeck PV, Kornhauser JM, Tkachev S, Zhang B, Skrzypek E, Murray B, Latham V, and Sullivan M (2012). PhosphoSitePlus: a comprehensive resource for investigating the structure and function of experimentally determined post-translational modifications in man and mouse. Nucleic Acids Res 40, D261–270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Hotamisligil GS, and Erbay E (2008). Nutrient sensing and inflammation in metabolic diseases. Nat. Rev. Immunol 8, 923–934. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Hou Z, He L, and Qi RZ (2007). Regulation of s6 kinase 1 activation by phosphorylation at ser-411. J. Biol. Chem 282, 6922–6928. [DOI] [PubMed] [Google Scholar]
  29. Humphrey SJ, Azimifar SB, and Mann M (2015). High-throughput phosphoproteomics reveals in vivo insulin signaling dynamics. Nat. Biotechnol 33, 990–995. [DOI] [PubMed] [Google Scholar]
  30. Jia J, Arif A, Ray PS, and Fox PL (2008). WHEP domains direct noncanonical function of glutamyl-prolyl tRNA synthetase in translational control of gene expression. Mol. Cell 29, 679–690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Johnson LN, and Lewis RJ (2001). Structural basis for control by phosphorylation. Chem. Rev 101, 2209–2242. [DOI] [PubMed] [Google Scholar]
  32. Kang SA, Pacold ME, Cervantes CL, Lim D, Lou HJ, Ottina K, Gray NS, Turk BE, Yaffe MB, and Sabatini DM (2013). mTORC1 phosphorylation sites encode their sensitivity to starvation and rapamycin. Science 341, 1236566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Khoury GA, Baliban RC, and Floudas CA (2011). Proteome-wide post-translational modification statistics: frequency analysis and curation of the swiss-prot database. Sci. Rep 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Kozma SC, McGlynn E, Siegmann M, Reinhard C, Ferrari S, and Thomas G (1993). Active baculovirus recombinant p70s6k and p85s6k produced as a function of the infectious response. J. Biol. Chem 268, 7134–7138. [PubMed] [Google Scholar]
  35. Lai KO, Liang Z, Fei E, Huang H, and Ip NY (2015). Cyclin-dependent kinase 5 (Cdk5)-dependent phosphorylation of p70 ribosomal S6 kinase 1 (S6K) Is required for dendritic spine morphogenesis. J. Biol. Chem 290, 14637–14646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Lalioti V, Muruais G, Dinarina A, van Damme J, Vandekerckhove J, and Sandoval IV (2009). The atypical kinase Cdk5 is activated by insulin, regulates the association between GLUT4 and E-Syt1, and modulates glucose transport in 3T3-L1 adipocytes. Proc. Natl. Acad. Sci. USA 106, 4249–4253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Law IK, Xu A, Lam KS, Berger T, Mak TW, Vanhoutte PM, Liu JT, Sweeney G, Zhou M, Yang B, et al. (2010). Lipocalin-2 deficiency attenuates insulin resistance associated with aging and obesity. Diabetes 59, 872–882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Li P, Fan W, Xu J, Lu M, Yamamoto H, Auwerx J, Sears DD, Talukdar S, Oh D, Chen A, et al. (2011). Adipocyte NCoR knockout decreases PPARgamma phosphorylation and enhances PPARgamma activity and insulin sensitivity. Cell 147, 815–826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Liao J, Sportsman R, Harris J, and Stahl A (2005). Real-time quantification of fatty acid uptake using a novel fluorescence assay. J. Lipid Res 46, 597–602. [DOI] [PubMed] [Google Scholar]
  40. Lobo S, Wiczer BM, and Bernlohr DA (2009). Functional analysis of long-chain acyl-CoA synthetase 1 in 3T3-L1 adipocytes. J. Biol. Chem 284, 18347–18356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Ma XM, and Blenis J (2009). Molecular mechanisms of mTOR-mediated translational control. Nat. Rev. Mol. Cell Biol 10, 307–318. [DOI] [PubMed] [Google Scholar]
  42. Magnuson B, Ekim B, and Fingar DC (2012). Regulation and function of ribosomal protein S6 kinase (S6K) within mTOR signalling networks. Biochem. J 441, 1–21. [DOI] [PubMed] [Google Scholar]
  43. Mahalingam M, and Templeton DJ (1996). Constitutive activation of S6 kinase by deletion of amino-terminal autoinhibitory and rapamycin sensitivity domains. Mol. Cell. Biol 16, 405–413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Manning G, Whyte DB, Martinez R, Hunter T, and Sudarsanam S (2002). The protein kinase complement of the human genome. Science 298, 1912–1934. [DOI] [PubMed] [Google Scholar]
  45. Miller CJ, and Turk BE (2018). Homing in: Mechanisms of Substrate Targeting by Protein Kinases. Trends Biochem. Sci 43, 380–394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Mok J, Kim PM, Lam HY, Piccirillo S, Zhou X, Jeschke GR, Sheridan DL, Parker SA, Desai V, Jwa M, et al. (2010). Deciphering protein kinase specificity through large-scale analysis of yeast phosphorylation site motifs. Sci. Signal 3, ra12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Moser BA, Dennis PB, Pullen N, Pearson RB, Williamson NA, Wettenhall RE, Kozma SC, and Thomas G (1997). Dual requirement for a newly identified phosphorylation site in p70s6k. Mol. Cell. Biol 17, 5648–5655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Mukhopadhyay NK, Price DJ, Kyriakis JM, Pelech S, Sanghera J, and Avruch J (1992). An array of insulin-activated, proline-directed serine/threonine protein kinases phosphorylate the p70 S6 kinase. J. Biol. Chem 267, 3325–3335. [PubMed] [Google Scholar]
  49. Murray-Zmijewski F, Slee EA, and Lu X (2008). A complex barcode underlies the heterogeneous response of p53 to stress. Nat. Rev. Mol. Cell Biol 9, 702–712. [DOI] [PubMed] [Google Scholar]
  50. Mylona A, Theillet FX, Foster C, Cheng TM, Miralles F, Bates PA, Selenko P, and Treisman R (2016). Opposing effects of Elk-1 multisite phosphorylation shape its response to ERK activation. Science 354, 233–237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Nash P, Tang X, Orlicky S, Chen Q, Gertler FB, Mendenhall MD, Sicheri F, Pawson T, and Tyers M (2001). Multisite phosphorylation of a CDK inhibitor sets a threshold for the onset of DNA replication. Nature 414, 514–521. [DOI] [PubMed] [Google Scholar]
  52. Nazari H, Khaleghian A, Takahashi A, Harada N, Webster NJ, Nakano M, Kishi K, Ebina Y, and Nakaya Y (2011). Cortactin, an actin binding protein, regulates GLUT4 translocation via actin filament remodeling. Biochemistry (Mosc.) 76, 1262–1269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Nemazanyy I, Panasyuk G, Zhyvoloup A, Panayotou G, Gout IT, and Filonenko V (2004). Specific interaction between S6K1 and CoA synthase: a potential link between the mTOR/S6K pathway, CoA biosynthesis and energy metabolism. FEBS Lett 578, 357–362. [DOI] [PubMed] [Google Scholar]
  54. Nussinov R, Tsai CJ, Xin F, and Radivojac P (2012). Allosteric post-translational modification codes. Trends Biochem. Sci 37, 447–455. [DOI] [PubMed] [Google Scholar]
  55. Olsson AK, Dimberg A, Kreuger J, and Claesson-Welsh L (2006). VEGF receptor signalling - in control of vascular function. Nat. Rev. Mol. Cell Biol 7, 359–371. [DOI] [PubMed] [Google Scholar]
  56. Papst PJ, Sugiyama H, Nagasawa M, Lucas JJ, Maller JL, and Terada N (1998). Cdc2-cyclin B phosphorylates p70 S6 kinase on Ser411 at mitosis. J. Biol. Chem 273, 15077–15084. [DOI] [PubMed] [Google Scholar]
  57. Pareek TK, Lam E, Zheng X, Askew D, Kulkarni AB, Chance MR, Huang AY, Cooke KR, and Letterio JJ (2010). Cyclin-dependent kinase 5 activity is required for T cell activation and induction of experimental autoimmune encephalomyelitis. J. Exp. Med 207, 2507–2519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Park KS, Mohapatra DP, Misonou H, and Trimmer JS (2006). Graded regulation of the Kv2.1 potassium channel by variable phosphorylation. Science 313, 976–979. [DOI] [PubMed] [Google Scholar]
  59. Paton CM, Rogowski MP, Kozimor AL, Stevenson JL, Chang H, and Cooper JA (2013). Lipocalin-2 increases fat oxidation in vitro and is correlated with energy expenditure in normal weight but not obese women. Obesity (Silver Spring) 21, E640–648. [DOI] [PubMed] [Google Scholar]
  60. Pawson T, and Scott JD (2005). Protein phosphorylation in signaling--50 years and counting. Trends Biochem. Sci 30, 286–290. [DOI] [PubMed] [Google Scholar]
  61. Pozo K, and Bibb JA (2016). The Emerging Role of Cdk5 in Cancer. Trends Cancer 2, 606–618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Pullen N, Dennis PB, Andjelkovic M, Dufner A, Kozma SC, Hemmings BA, and Thomas G (1998). Phosphorylation and activation of p70s6k by PDK1. Science 279, 707–710. [DOI] [PubMed] [Google Scholar]
  63. Ray PS, and Fox PL (2007). A post-transcriptional pathway represses monocyte VEGF-A expression and angiogenic activity. EMBO J 26, 3360–3372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Richards MR, Harp JD, Ory DS, and Schaffer JE (2006). Fatty acid transport protein 1 and long-chain acyl coenzyme A synthetase 1 interact in adipocytes. J. Lipid Res 47, 665–672. [DOI] [PubMed] [Google Scholar]
  65. Rubin SM (2013). Deciphering the retinoblastoma protein phosphorylation code. Trends Biochem. Sci 38, 12–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Ruvinsky I, Sharon N, Lerer T, Cohen H, Stolovich-Rain M, Nir T, Dor Y, Zisman P, and Meyuhas O (2005). Ribosomal protein S6 phosphorylation is a determinant of cell size and glucose homeostasis. Genes Dev 19, 2199–2211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Sampath P, Mazumder B, Seshadri V, Gerber CA, Chavatte L, Kinter M, Ting SM, Dignam JD, Kim S, Driscoll DM, et al. (2004). Noncanonical function of glutamyl-prolyl-tRNA synthetase: gene-specific silencing of translation. Cell 119, 195–208. [DOI] [PubMed] [Google Scholar]
  68. Sarker KP, and Lee KY (2004). L6 myoblast differentiation is modulated by Cdk5 via the PI3K-AKT-p70S6K signaling pathway. Oncogene 23, 6064–6070. [DOI] [PubMed] [Google Scholar]
  69. Saxton RA, and Sabatini DM (2017). mTOR signaling in growth, metabolism, and disease. Cell 169, 361–371. [DOI] [PubMed] [Google Scholar]
  70. Shah OJ, Ghosh S, and Hunter T (2003). Mitotic regulation of ribosomal S6 kinase 1 involves Ser/Thr, Pro phosphorylation of consensus and non-consensus sites by Cdc2. J. Biol. Chem 278, 16433–16442. [DOI] [PubMed] [Google Scholar]
  71. Sims RJ 3rd, and Reinberg D (2008). Is there a code embedded in proteins that is based on post-translational modifications? Nat. Rev. Mol. Cell Biol 9, 815–820. [DOI] [PubMed] [Google Scholar]
  72. Tunduguru R, Zhang J, Aslamy A, Salunkhe VA, Brozinick JT, Elmendorf JS, and Thurmond DC (2017). The actin-related p41ARC subunit contributes to p21-activated kinase-1 (PAK1)-mediated glucose uptake into skeletal muscle cells. J. Biol. Chem 292, 19034–19043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Ubeda M, Kemp DM, and Habener JF (2004). Glucose-induced expression of the cyclin-dependent protein kinase 5 activator p35 involved in Alzheimer’s disease regulates insulin gene transcription in pancreatic beta-cells. Endocrinology 145, 3023–3031. [DOI] [PubMed] [Google Scholar]
  74. Ubersax JA, and Ferrell JE Jr. (2007). Mechanisms of specificity in protein phosphorylation. Nat. Rev. Mol. Cell Biol 8, 530–541. [DOI] [PubMed] [Google Scholar]
  75. Walsh CT, Garneau-Tsodikova S, and Gatto GJ Jr. (2005). Protein posttranslational modifications: the chemistry of proteome diversifications. Angew Chem. Int. Ed. Engl 44, 7342–7372. [DOI] [PubMed] [Google Scholar]
  76. Wei FY, Nagashima K, Ohshima T, Saheki Y, Lu YF, Matsushita M, Yamada Y, Mikoshiba K, Seino Y, Matsui H, et al. (2005). Cdk5-dependent regulation of glucose-stimulated insulin secretion. Nat. Med 11, 1104–1108. [DOI] [PubMed] [Google Scholar]
  77. Weng QP, Kozlowski M, Belham C, Zhang A, Comb MJ, and Avruch J (1998). Regulation of the p70 S6 kinase by phosphorylation in vivo. Analysis using site-specific anti-phosphopeptide antibodies. J. Biol. Chem 273, 16621–16629. [DOI] [PubMed] [Google Scholar]
  78. Yan QW, Yang Q, Mody N, Graham TE, Hsu CH, Xu Z, Houstis NE, Kahn BB, and Rosen ED (2007). The adipokine lipocalin 2 is regulated by obesity and promotes insulin resistance. Diabetes 56, 2533–2540. [DOI] [PubMed] [Google Scholar]
  79. Zhyvoloup A, Nemazanyy I, Babich A, Panasyuk G, Pobigailo N, Vudmaska M, Naidenov V, Kukharenko O, Palchevskii S, Savinska L, et al. (2002). Molecular cloning of CoA Synthase. The missing link in CoA biosynthesis. J. Biol. Chem 277, 22107–22110. [DOI] [PubMed] [Google Scholar]
  80. Zhyvoloup A, Nemazanyy I, Panasyuk G, Valovka T, Fenton T, Rebholz H, Wang ML, Foxon R, Lyzogubov V, Usenko V, et al. (2003). Subcellular localization and regulation of coenzyme A synthase. J. Biol. Chem 278, 50316–50321. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

2
3

Table S1. Proteomic analysis of S6K1*-interacting proteins, Related to Figure 6.

(A) Summary of results obtained from each lane divided into eight regions. (B) Summary of results by lane type, i.e., control, S6K1 and S6K1*. (C) Comparison of proteins in control, S6K1, and S6K1* lanes.

Data Availability Statement

Comprehensive data from the proteomic analysis of S6K1*-interacting proteins is included as a single Excel spreadsheet.

RESOURCES