Abstract
Many serine/threonine protein kinases discriminate between serine and threonine substrates as a filter to control signaling output. Among these, the p21-activated kinase (PAK) group strongly favors phosphorylation of Ser over Thr residues. PAK4, a group II PAK, almost exclusively phosphorylates its substrates on serine residues. The only well documented exception is LIM domain kinase 1 (LIMK1), which is phosphorylated on an activation loop threonine (Thr508) to promote its catalytic activity. To understand the molecular and kinetic basis for PAK4 substrate selectivity we compared its mode of recognition of LIMK1 (Thr508) with that of a known serine substrate, β-catenin (Ser675). We determined X-ray crystal structures of PAK4 in complex with synthetic peptides corresponding to its phosphorylation sites in LIMK1 and β-catenin to 1.9 Å and 2.2 Å resolution, respectively. We found that the PAK4 DFG+1 residue, a key determinant of phosphoacceptor preference, adopts a sub-optimal orientation when bound to LIMK1 compared to β-catenin. In peptide kinase activity assays, we find that phosphoacceptor identity impacts catalytic efficiency but does not affect the Km value for both phosphorylation sites. Although catalytic efficiency of wild-type LIMK1 and β-catenin are equivalent, T508S mutation of LIMK1 creates a highly efficient substrate. These results suggest suboptimal phosphorylation of LIMK1 as a mechanism for controlling the dynamics of substrate phosphorylation by PAK4.
Keywords: Kinase, crystallography, LIM domain kinase, β-catenin, Ste20 kinase, substrate specificity
Graphical Abstract
Introduction
For protein kinases to correctly propagate signals, they must selectively target their downstream substrates. This can occur through an array of mechanisms including co-localization, docking interactions (Ubersax and Ferrell, 2007), and direct binding of the kinase catalytic domain to the substrate (Miller and Turk, 2018). Among these interactions, short linear motifs comprising amino acids that flank the substrate residue can significantly contribute to selectivity for both serine/threonine and tyrosine kinases (Miller and Turk, 2018; Turk, 2008). The phosphoacceptor residue itself is also a key component of substrate specificity; for example, tyrosine kinases are structurally distinct from serine/threonine kinases (Taylor et al., 1995; Ubersax and Ferrell, 2007). Similarly, serine/threonine kinases often discriminate between serine and threonine residues, which is driven by the residue immediately following the conserved activation segment DFG motif, termed the DFG+1 residue (Chen et al., 2014). While most protein kinases display strong phosphorylation site preferences in vitro, many bona fide protein substrates are phosphorylated at sites would appear to be disfavored by the kinase (Chen et al., 2014; Pinna and Ruzzene, 1996). The mechanisms of this apparent disfavored phosphorylation are not generally well understood, so in this study we sought to understand an important disfavored kinase-substrate pair.
The p21-activated kinases (PAKs), are members of the sterile-20 family of serine-threonine kinases and are major downstream effectors of the Rho family of small GTPases (Arias-Romero and Chernoff, 2008; Ha et al., 2015; Hofmann et al., 2004; King et al., 2014; Kumar et al., 2017; Wells and Jones, 2010). The six members of the PAK group are classified by their domain organization into two sub-groups, the type I (PAK1-3) and type II (PAK4-6) PAKs, all of which play roles in cytoskeletal remodeling, cell motility, inhibition of apoptosis and transcription regulation (Ha et al., 2015; Kumar et al., 2017; Li et al., 2012). Of the type II PAKs, PAK4 is the best studied and has demonstrated roles in embryonic and neural development (Dart and Wells, 2013; Wells et al., 2010); its overexpression is also linked to tumorigenesis and metastasis, particularly in prostate and non-small cell lung cancers (Ahmed et al., 2008; Cai et al., 2015; Ha et al., 2015; Lu et al., 2017; Minden, 2012; Thillai et al., 2018; Wells et al., 2010).
All six PAKs strongly favor phosphorylation of serine over threonine, with type II PAKs including PAK4 appearing more selective than type I PAKs (Rennefahrt et al., 2007). Although PAK4 has multiple validated direct substrates (Table 1) (Bompard et al., 2010; Callow et al., 2002; Dan et al., 2001; Guo et al., 2014; Li et al., 2012; Li et al., 2010; Wells et al., 2010; Xu et al., 2016; Zhao et al., 2017; Zhuang et al., 2015), only one of these substrates is phosphorylated on a threonine residue, Thr508 of LIM domain kinase 1 (LIMK1) (Dan et al., 2001). This residue is phosphorylated by PAK4 (and other Rho-effector kinases) (Bernard, 2007; Scott and Olson, 2007), and its phosphorylation is required for full activation of LIMK1, which in turn phosphorylates and inactivates cofilin proteins - actin binding proteins that facilitate actin filament severing (Dan et al., 2001; Hamill et al., 2016). PAK4 therefore has an unusual exception to its strong preference for phosphorylation of serine substrates. The molecular mechanisms for this exception are unclear, especially because previous studies suggested the large and hydrophobic phenylalanine residue at PAK4’s DFG+1 position should strongly disfavor phosphorylation of LIMK1-Thr508 (Chen et al., 2014).
Table 1. Documented human PAK4 substrate sequences.
PAK4 Substrate | Phosphosite Sequence | Phosphosite | Citation |
---|---|---|---|
PAK4 | KEVPRRKSLVGTPYW | S474 | Callow et al. (2002) |
β-catenin | QDYKKRLSVELTSSL | S675 | Li et al. (2012) |
GEF-H1 (ARHGEF2) | PVDPRRRSLPAGDAL | S810 | Callow et al. (2005) |
Paxillin | ELDELMASLSDFKFM | S272 | Wells et al. (2010) |
Estrogen Receptor- α | IKRSKKNSLALSLTA | S305 | Zhuang et al. (2015) |
N-WASP | KRSKAIHSSDEDEDE | S484, | Zhao et al. (2017) |
RSKAIHSSDEDEDED | S485 | ||
P53 | DRNTFRHSVVVPYEP | S215 | Xu et al. (2016) |
Ran | DRKVKAKSIVFHRKK | S135 | Bompard et al. (2010) |
STMN2 | KQINKRASGQAFELI | S50 | Guo et al. (2014) |
BAD | EIRSRHSSYPAGTED | S75 | Gnesutta et al (2001) |
Raf-1 | RPRGQRDSSYYWEIE | S338 | Cammarano et al. (2005) |
p120-catenin | GLEDDQRSMGYDDLD | S288 | Wong et al. (2010) |
Integrin β5 | REFAKFQSERSRARY | S759 | Li et al. (2010) |
AKFQSERSRARYEMA | S762 | ||
LIMK1 | PDRKKRYTVVGNPYW | T508 | Dan et al. (2001) |
To understand the mechanisms of PAK4’s atypical phosphorylation of LIMK1, we conducted a structural and functional study. We determined a 1.9 Å crystal structure of PAK4 catalytic domain in complex with a synthesized peptide corresponding to LIMK1’s substrate sequence (Thr508), and we compared this to a 2.2 Å crystal structure of PAK4 catalytic domain in complex with a synthesized peptide corresponding to a well characterized serine substrate of PAK4, β-catenin (Ser675). We find that similar to optimized ‘PAKtide’ substrates (Chen et al., 2014), the DFG+1 residue is re-oriented to accommodate the unusual Thr phosphoacceptor. We then assessed the kinetics of kinase activity for PAK4 against LIMK and β-catenin substrate peptides with either Ser or Thr as the phosphoacceptor residue. We find that phosphoacceptor identity increases catalytic efficiency but does not affect Km, and although catalytic efficiency of wild-type LIMK1 and β-catenin are equivalent, T508S mutation of the LIMK1 peptide creates a highly efficient substrate.
Materials and Methods
Protein Expression and Purification
PAK4 expression and purification was conducted as previously described (Chen et al., 2014; Ha et al., 2012; Zhang et al., 2018). The catalytic domain of PAK4 (residues 109-426) (UniProt ID: O96013-2) was expressed using a modified pET28 vector with hexa-histidine (6xHis) tag, removable by TEV protease. Following nickel-affinity chromatography using a HisTrap chelating column (GE Healthcare) and gel filtration using a Superdex 200 10/300GL (GE Healthcare), purified PAK4 catalytic domain was concentrated to 4 mg/mL for co-crystallization, or to 0.98 μg/mL for kinase activity assays.
Peptide Synthesis.
Peptides for crystallization and kinase assays were synthesized on the 0.1 mM scale with N-terminal acetyl and C-terminal amide and HPLC purified at Tufts University Core Facility. Peptides were resuspended in distilled water and concentration monitored using a Nanodrop spectrophotometer (GE). For kinase assays, peptides of equal length were used for LIMK1 and β-catenin: LIMK1 peptide (RKKRYTVVGN), LIMK1T508S (RKKRYSVVGN), β-catenin (YKKRLSVELT), β-cateninS6757 (YKKRLTVELT). Peptides correspond to residues 503 to 512 of LIMK1 (UniProt ID: P53667) and residues 670-679 of β-catenin (UniProt UD: P35222). The PAKtide peptide was one amino acid residue longer at the C-terminus (RKRRNSLAYKK). For structure determination, a shorter β-catenin peptide (KKRLSVE) was used.
Structure determination
PAK4 catalytic domain crystals were grown against 0.1 M Tris-HCL (pH 7.5) and 2 M sodium acetate at room temperature using the hanging drop vapor diffusion method (well: 500 μL; drop: 1 μL protein mixed with 1 μL precipitant). Prior to mixture with precipitant, the protein solution contained 1 mM AMP-PNP and 5 mM MgCl2. Following crystal growth, 0.2 μL of peptide solution was added to a final concentration of 2 mM LIMK1 peptide, or a final concentration of 5 mM β-catenin peptide. Soaking was conducted for 1 day at room temperature. Crystals were looped and cryoprotected in 3 M sodium acetate. Crystallographic data were collected on beamline 24-ID-E of the Advanced Photon Source (APS) and were processed using the HKL-2000 package (Otwinowski and Minor, 1997) and molecular replacement using Phaser (McCoy et al., 2007) to obtain an initial solution with a previous structure of PAK4 alone as the search model (PDB ID: 4FIJ (Ha et al., 2012)). Iterative refinement and model building were conducted using the Phenix package (Adams et al., 2010) and Coot (Emsley et al., 2010). For both structures, after refinement of the kinase domain alone, clear unbiased difference density was observed in the substrate binding site allowing bound peptide to be built. In both structures, there was insufficient electron density to build bound AMP-PNP. Final placement of peptide position was validated using simulated annealing omit maps. Model validation was conducted using MolProbity (Chen et al., 2010).
Kinase activity assays
PAK4 activity was determined by a filter binding assay performed in triplicate using radiolabeled [γ-32P]ATP, modified from (Chen et al., 2014; Hastie et al., 2006). PAK4 kinase domain (25 nM) was incubated with peptides at varying concentrations in kinase buffer (50 mM HEPES, pH 7.4, 10 mM MgCl2, 12.5 mM NaCl, 1 mM MnCl2, 1 mM DTT, 0.1 mM EGTA). Peptide concentrations were titrated in two-fold increments across the following concentration ranges: LIMK1, 25 μM - 1600 μM; LIMK1T508S 2.5 μM - 320 μM; β-catenin, 56.3 μM - 3600 μM; β-cateninS675T, 56.3 μM - 1800 μM; PAKtide, 6.25 μM - 400 μM (Miller et al., 2019). Reactions were initiated by addition of ATP to reach a final concentration of 20 μM with 0.025 μCi/μL [γ-32P]ATP incubated at 30°C for 10 min and 20 min. Aliquots of each reaction mixture (5 μL) were pipetted onto P81 phosphocellulose filter paper that was then quenched in 75 mM phosphoric acid. Filters were washed twice for 5 min in 75 mM phosphoric acid, washed once briefly in acetone, air-dried, suspended in 10 mL Optifluor scintillation fluid, and analyzed with a liquid scintillation counter. The concentration of product formed was calculated using standards made from unwashed filters. Initial reaction velocities were found by fitting the [product] vs time data to a line with Excel (Microsoft). Catalytic constants were calculated by fitting the initial reaction velocities to the Michaelis-Menten equation using GraphPad Prism (version 8).
Results
To date, no structures have been determined of PAK family members in complex with regions of physiologically relevant substrates. Almost all type II PAK substrates are phosphorylated on a serine residue, with the sole exception of LIM domain kinase 1, which is phosphorylated on its activation loop threonine (Thr508) (Table 1, Figure 1A). This exception to the strong preference of type II PAKs towards serine as a substrate caused us to wonder whether there is a structural basis for this difference. We therefore determined crystal structures of the PAK4 catalytic domain in complex with synthetic peptide substrates derived from its physiological substrates LIMK1 and β-catenin.
The crystal structure of PAK4 catalytic domain (residues 109-426) in complex with the LIMK1 peptide substrate (RKKRYT508VVGN) was determined to 1.9 Å resolution (Table 2), with good electron density observed for the bound substrate (Figure 1B,C). Similarly, the crystal structure of PAK4 catalytic domain in complex with a synthetic peptide corresponding to its major phosphorylation site in β-catenin (KKRLS695VE) was determined to 2.2 Å resolution (Table 2) with good electron density observed for the bound substrate (Figure 1D,E). Both structures exhibit the typical two lobe kinase fold, consisting of an N-lobe with five β-sheets and one α-helix and a C-lobe with mostly α-helices (Figure 1B,D). PAK4 in both structures is in the active state ‘DFG-in’ conformation, and the activation loop is phosphorylated on Ser474, similarly to other E.coli expressed PAK4 crystal structures (Baskaran et al., 2015; Chen et al., 2014; Ha and Boggon, 2018; Ha et al., 2012; Zhang et al., 2018). Root-mean-deviation between PAK4 in the structures is 0.2 Å over 292 residues (McNicholas et al., 2011).
Table 2. Data collection and refinement statistics.
Complex | PAK4 with LIMK1 peptide | PAK4 with β-catenin peptide |
---|---|---|
PDB accession code | 6WLY | 6WLX |
Data collection | ||
Space group | P41212 | P41212 |
PAK4 Substrate | Phosphosite Sequence | Phosphosite | Citation |
---|---|---|---|
PAK4 | KEVPRRKSLVGTPYW | S474 | Callow et al. (2002) |
β-catenin | QDYKKRLSVELTSSL | S675 | Li et al. (2012) |
GEF-H1 (ARHGEF2) | PVDPRRRSLPAGDAL | S810 | Callow et al. (2005) |
Paxillin | ELDELMASLSDFKFM | S272 | Wells et al. (2010) |
Estrogen Receptor- α | IKRSKKNSLALSLTA | S305 | Zhuang et al. (2015) |
N-WASP | KRSKAIHSSDEDEDE RSKAIHSSDEDEDED | S484, S485 | Zhao et al. (2017) |
p53 | DRNTFRHSVVVPYEP | S215 | Xu et al. (2016) |
Ran | DRKVKAKSIVFHRKK | S135 | Bompard et al. (2010) |
STMN2 | KQINKRASGQAFELI | S50 | Guo et al. (2014) |
BAD | EIRSRHSSYPAGTED | S75 | Gnesutta et al (2001) |
Raf-1 | RPRGQRDSSYYWEIE | S338 | Cammarano et al. (2005) |
p120-catenin | GLEDDQRSMGYDDLD | S288 | Wong et al. (2010) |
Integrin β5 | REFAKFQSERSRARY AKFQSERSRARYEMA | S759 S762 | Li et al. (2010) |
LIMK1 | PDRKKRYTVVGNPYW | T508 | Dan et al. (2001) |
X-ray source | APS NE-CAT-E | APS NE-CAT-E | |
Detector | Eiger 16M PAD | Eiger 16M PAD | |
Wavelength (Å) | 0.97918 | 0.97918 | |
Unit cell | |||
a, b, c, (Å) | 61.6, 61.6, 179.6 | 61.6, 61.6, 179.4 | |
α, β, γ, (°) | 90.0, 90.0, 90.0 | 90.0, 90.0, 90.0 | |
Resolution range (Å)* | 50.00-1.90 (1.97-1.90) | 50.00-2.20 (2.28-2.20) | |
No. of unique reflections | 28404 (2747) | 18577 (1799) | |
Multiplicity* | 20.3 (15.3) | 11.2 (11.8) | |
Completeness (%)* | 99.9 (99.9) | 99.8 (100) | |
Mean I/σ(I)* | 20.8 (1.8) | 20.7 (2.0) | |
Wilson B factor (Å2) | 38.5 | 47.5 | |
Rpim (%)* | 2.7 (38.6) | 3.2 (34.7) | |
CC 1/2* | 0.999 (0.615) | 0.977 (0.701) | |
Refinement statistics | |||
Reflections used in refinement | 28228 (2621) | 18366 (2419) | |
Resolution range* (Å) | 43.58-1.90 (1.97-1.90) | 44.85-2.20 (2.32-2.20) | |
Rwork (%)* | 18.6 (31.5) | 19.7 (29.5) | |
Rfree (%)* | 20.5 (32.1) | 22.9 (35.2) | |
Free R reflections (%)* | 5.0 (5.0) | 5.0 (5.0) | |
Free R reflections, no.* | 1412 (138) | 919 (88) | |
Residue range built | PAK4 / 299-590, LIMK1 / 504-512 | PAK4 / 299-590, β-catenin / 672-676 | |
Soaked peptide / Built peptide | RKKRYTVVGN / KKRYTVVGN | KKRLSVE / KRLSV | |
No. water molecules | 135 | 54 | |
Mean B factor, Å2 | |||
All atoms | 45.9 | 57.42 | |
PAK4 (chain A) | 45.1 | 57.16 | |
Peptide (chain B) | 66.3 | 82.06 | |
H2O | 47.7 | 49.88 | |
Model statistics | |||
RMSD bond lengths (Å) | 0.005 | 0.010 | |
RMSD bond angles (°) | 0.776 | 1.094 | |
Ramachandran plot: favored/allowed/outliers (%) | 98.6 / 1.4 / 0.00 | 97.59 / 2.41 / 0.00 | |
MolProbity clashscore | 3.21 (98th percentile) | 5.24 (98th percentile) |
Parentheses indicate highest resolution shell.
Crystallographic data and refinement statistics indicate good quality diffraction data and refined crystal structures.
The mode of peptide binding in both structures is broadly similar. Each peptide is posed in the canonical kinase peptide substrate binding site (Figure 1B,D), similar to other PAK-substrate or PAK-pseudosubstrate structures (Baskaran et al., 2015; Chen et al., 2014; Ha and Boggon, 2018; Ha et al., 2012), and the positioning of the substrate residues is broadly conserved (Figure 2A). Concordantly, both peptides share similar peptide backbone conformations, indicating that the identity of the phosphoacceptor residue does not significantly impact peptide backbone conformation (Figure 2B,C). A common feature of the previously determined PAK-substrate and PAK-pseudosubstrate structures is PAK recognition of a basic residue N-terminal to the phosphorylation site (Ha et al., 2015; Rennefahrt et al., 2007). In PAK4 this is achieved through an acidic cradle defined by residues Asp444 and Glu507 (Ha et al., 2015; Rennefahrt et al., 2007). For both PAK4-LIMK1 and PAK4-β-catenin, the acidic cradle makes salt bridge interactions with the headgroup of an Arg residue two residues N-terminal to the phosphorylation site (the P-2 Arg) (Figure 2D,E).
The phosphoacceptor residue hydroxyl group for both LIMK1 and β-catenin peptide substrates is poised for phosphotransfer, although nucleotide is not visible in either structure. As with synthetic optimized peptides (Chen et al., 2014), the consequences of correct orientation of the phosphoacceptor residue are structurally observed in the orientation of the residue immediately C-terminal to the conserved DFG motif, the DFG+1 residue (Phe461). In the β-catenin bound structure, the phenyl ring is oriented approximately perpendicular to the peptide backbone, but in the LIMK1 bound structure Phe461 rotates to accommodate the methyl group on Thr508 (Figure 2B,C). Similar orientation changes were observed for peptides optimized for PAK phosphorylation (PAKtides) but have not previously been observed for peptides corresponding to physiologically relevant substrates. These structures therefore illustrate that peptides derived from physiological substrates follow similar recognition rules to those observed for optimized substrates (Chen et al., 2014).
To assess the enzymatic impact of the unusual threonine phosphoacceptor in LIM kinases compared to the strong preference of PAK4 for serine, we conducted kinase activity assays. Using radiolabeled [γ-32P]ATP we determined Michaelis-Menten kinetics and obtained Km and kcat values for PAK4 catalytic domain phosphorylation of 5 distinct peptides (Figure 3, Table 3). We first compared phosphorylation of synthetic peptides corresponding to β-catenin (YKKRLS675VELT) and LIMK1 (RKKRYT508VVGN). We found that unexpectedly the catalytic efficiency for PAK4 phosphorylation of the LIMK1 and β-catenin peptides was broadly equivalent (kcat/Km ≈ 300) (Figure 3A, inset i, Table 3).
Table 3. Catalytic constants for phosphorylation of peptide substrates by PAK4 and exchange mutants.
Peptide | Sequence | kcat (S−1) | Km (μM) (95% CI) | kcat/Km (M−1S−1) |
---|---|---|---|---|
LIMK1 | RKKRYTVVGN | 0.038 ± 0.004 | 150 ± 60a* (70 to 320) | 300 ± 100+ |
LIMK1T508S | RKKRYSVVGN | 0.29 ± 0.04* | 60 ± 20ab+ (25 to 135) | 5000 ± 2000* |
β-catenin | YKKRLSVELT | 0.15 ± 0.02a | 600 ± 300c* (260 to 1400) | 300 ±100+ |
β-cateninS675T | YKKRLTVELT | 0.010 ± 0.003 | 1100 ± 600ac+# (400 to 5300) | 9 ± 6 |
PAKtide | RKRRNSLAYKK | 0.20 ± 0.02a* | 60 ± 20ab# (25 to 125) | 4000 ± 1000* |
We then assessed the impact of switching the phosphorylation site residue for both β-catenin and LIMK1 derived peptides. For β-catenin we generated a Thr-substrate peptide (YKKRLT675VELT) which we term β-cateninS675T, and for LIMK1 we generated a Ser-substrate peptide (RKKRYS508VVGN) which we term LIMK1T508S. Comparison of the kinetics for β-catenin and β-cateninS675T indicates the Km values are not significantly different, but the kcat and catalytic efficiency are reduced by approximately an order of magnitude for the non-wild type Thr-substrate peptide (Figure 3A, inset ii). Likewise, comparison of the kinetics for LIMK1 and LIMK1T508S indicates that a Thr-substrate is disfavored. Again, we find the Km values are not significantly different but kcat and catalytic efficiency increase by approximately an order of magnitude for the non-wild type Ser-substrate peptide (Figure 3A, inset iii). These results demonstrate that Thr as a substrate residue is highly disfavored, primarily by reduction of catalytic efficiency.
Lastly, we compared the catalytic efficiency of LIMK1T508S with an optimized PAKtide peptide (RKRRNSLAYKK) that was previously designed based on linear peptide motif analysis (Miller et al., 2019). We find that catalytic efficiency of LIMK1T508S is broadly similar to the optimized PAKtide peptide (Figure 3A, inset iv). This suggests that LIM kinase with a serine phosphorylation site on its activation loop would represent a highly efficient substrate for PAK4.
Discussion
PAK4 phosphorylates all of its known substrates on serine with the sole known exception of LIM domain kinase 1, which it phosphorylates on threonine. In this study we sought to understand the molecular and enzymatic reasons for these exceptions by studying PAK4 phosphorylation of a well-known serine substrate, β-catenin, and comparing this to its unusual substrate, LIMK1. We reveal the molecular basis for selectivity by showing that the DFG+1 residue is important for discriminating between serine and threonine phosphoacceptors, but LIMK1 phosphorylation by PAK4 proceeds at a similar catalytic efficiency as β-catenin. We also assessed the impact of a LIMK1 optimizing mutation, T508S, and find that this yields a good PAK4 substrate with catalytic efficiency equivalent to a designed PAKtide peptide. These studies therefore provide new insights into PAK phosphorylation of physiologically relevant substrates.
At the molecular level, we find that PAK4’s interaction with physiologically relevant Ser/Thr phosphoacceptors seems to be associated with its differences in enzymatic activity towards these substrates. The DFG+1 residue, immediately C-terminal to the conserved DFG motif, was previously found in conformations that seem to depend on the identity of the substrate residue. For designed peptide substrates that are optimized as PAKtide peptides, a mutation from Ser to Thr as the phosphoacceptor results in reorientation of Phe461 around its β-γ bond. This is thought to interfere with placement of ATP γ-phosphate and to hinder phosphate transfer (Chen et al., 2014). We show that similar conformational movements in the DFG+1 residue occur for the physiologically relevant substrates β-catenin and LIMK1, indicating that molecular level discrimination between Ser and Thr substrates occurs at the DFG+1 residue for these substrates.
At the enzymatic level, our comparison of PAK4’s physiological substrates shows that despite possessing a Thr phosphoacceptor residue LIMK1 is a favorable substrate phosphorylated by PAK4 with similar catalytic efficiency to β-catenin, a Ser phosphoacceptor substrate. It is not immediately clear which residues of the linear motif drive the tighter Km observed for LIMK1 compared to β-catenin (Table 3). Nevertheless, recent studies have shown that interactions between the kinase β3-αC loop and substrate residues C-terminal to the phosphoacceptor are important for controlling catalytic efficiency (Miller et al., 2019), implying that Val-Gly at positions +2 and +3 in LIMK1 are more preferable than Glu-Leu in β-catenin. This position preference is further supported by peptide array profiling (Rennefahrt et al., 2007).
The effect of switching the phosphoacceptor residue in both the β-catenin and LIMK1 peptide yields results of similar magnitude. For β-catenin, mutation of the phosphoacceptor Ser to Thr results in an approximately 10-fold reduction in kcat and commensurate change in catalytic efficiency, and similarly, for LIMK1, mutation of the phosphoacceptor Thr to Ser results in an approximately 10-fold increase in kcat and change in catalytic efficiency. For LIMK1, however, this increase in catalytic efficiency results in a peptide that compares favorably with a substrate peptide designed to be optimal for phosphorylation by PAKs based on peptide array profiling (Miller et al., 2019; Rennefahrt et al., 2007). Compared to PAK2, PAK4’s phosphorylation of a LIMK1 peptide has a low kcat (Wu and Wang, 2003). This may be because PAK4, unlike PAK2, has a noncanonical KPEN motif containing a Ser instead of an Asn residue (Taylor et al., 1992). Mutation of this Ser to an Asn increases PAK4 activity (Miller et al., 2019; Qu et al., 2001) implying that the noncanonical KPEN motif hinders PAK4 activity. It is interesting to hypothesize that the weak kinase activity of PAK4 may increase the importance of substrate recognition for allowing efficient catalysis and consequent downstream signaling.
The presence of a Thr phosphoacceptor site on LIMKs is highly conserved evolutionarily, and many of their upstream kinases preferentially phosphorylate threonine over serine, including the ROCK and MRCK families (Amano et al., 2001; Maekawa et al., 1999; Ohashi et al., 2000; Sumi et al., 2001b). Although our study does not establish disfavor for the PAK-LIMK pathway in a cellular context, in other systems the presence of a suboptimal phosphoacceptor residue confers sensitivity to inhibitors and tempers response to activating stimuli (Kang et al., 2013). It is tempting to speculate on the impact of optimizing PAK-LIMK by altering LIM kinase phosphoacceptor identity and tuning phosphosite specificity. This may allow simultaneous enhancement of PAK-driven signals to the cytoskeleton from RAC/CDC42 and suppression of ROCK-driven signals from RHO. If, however, LIMK1 requires Thr in its activation loop for full activity, substituting of the phosphoacceptor residue for Ser could negatively impact its function (Kaldis et al., 2000). Considering LIMK1’s and LIMK2’s distinct subcellular localization (Acevedo et al., 2006; Sumi et al., 2006) and that ROCK phosphorylates LIMK2 (Sumi et al., 2001a), the possibility that LIMK1 and LIMK2 have distinct upstream pathways remains open, and we note that while PAK phosphorylation of LIMK1 is well-established (Dan et al., 2001; Edwards et al., 1999), to date, PAK phosphorylation of LIMK2 has not been directly shown.
In summary, we find that PAK4 phosphorylation of peptide substrates is significantly impacted by the identity of the phosphoacceptor residue, but both β-catenin and LIMK1 fall into a similar range of catalytic efficiency. Although catalytic efficiency of wild-type LIMK1 and β-catenin are equivalent, T508S mutation of LIMK1 creates a highly efficient substrate suggesting that inefficient phosphorylation of LIMK1 is a mechanism for controlling normal substrate dynamics.
Highlights.
Molecular basis for PAK4 phosphorylation of physiological substrates
Phosphoacceptor identity impacts catalytic efficiency but does not affect the Km
Suggests suboptimal phosphorylation of LIMK1 as a mechanism for controlling the dynamics of substrate phosphorylation by PAK4
Acknowledgements
We thank Chad Miller for preparation of PAK4 protein used in this study. Staff at beamline 24-ID-E (NE-CAT-E) at the Advanced Photon Source, Argonne National Laboratory are thanked. This work is based upon research conducted at the Northeastern Collaborative Access Team beamlines, which are funded by National Institutes of Health grant P41GM103403. This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02- 06CH11357. AKC was partially funded by the Yale College First-Year Summer Research Fellowship in the Sciences & Engineering. NIH Grants R01GM102262 to TJB and BET, T32GM007324 to JAS, and S10OD018007 and American Heart Association Grant 19IPLOI34740007 to TJB funded this research.
Footnotes
Data availability
Coordinates and structure factors have been deposited in the Protein Data Bank under accession codes 6WLX and 6WLY. X-ray diffraction images are available online at SBGrid Data (Meyer et al., 2016): doi: 10.15785/SBGRID/781 (6WLX) and doi: 10.15785/SBGRID/782 (6WLY).
Competing Interests
The authors declare no competing financial interests.
Declaration of interests
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
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 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.
References
- Acevedo K, Moussi N, Li R, Soo P, Bernard O, 2006. LIM kinase 2 is widely expressed in all tissues. J Histochem Cytochem 54, 487–501. [DOI] [PubMed] [Google Scholar]
- Adams PD, Afonine PV, Bunkoczi G, Chen VB, Davis IW, Echols N, Headd JJ, Hung LW, Kapral GJ, Grosse-Kunstleve RW, McCoy AJ, Moriarty NW, Oeffner R, Read RJ, Richardson DC, Richardson JS, Terwilliger TC, Zwart PH, 2010. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr D Biol Crystallogr 66, 213–221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahmed T, Shea K, Masters JR, Jones GE, Wells CM, 2008. A PAK4-LIMK1 pathway drives prostate cancer cell migration downstream of HGF. Cell Signal 20, 1320–1328. [DOI] [PubMed] [Google Scholar]
- Amano T, Tanabe K, Eto T, Narumiya S, Mizuno K, 2001. LIM-kinase 2 induces formation of stress fibres, focal adhesions and membrane blebs, dependent on its activation by Rho-associated kinase-catalysed phosphorylation at threonine-505. Biochem J 354, 149–159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arias-Romero LE, Chernoff J, 2008. A tale of two Paks. Biol Cell 100, 97–108. [DOI] [PubMed] [Google Scholar]
- Baskaran Y, Ang KC, Anekal PV, Chan WL, Grimes JM, Manser E, Robinson RC, 2015. An in cellulo-derived structure of PAK4 in complex with its inhibitor Inka1. Nature communications 6, 8681. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bernard O, 2007. Lim kinases, regulators of actin dynamics. Int J Biochem Cell Biol 39, 1071–1076. [DOI] [PubMed] [Google Scholar]
- Bompard G, Rabeharivelo G, Frank M, Cau J, Delsert C, Morin N, 2010. Subgroup II PAK-mediated phosphorylation regulates Ran activity during mitosis. J Cell Biol 190, 807–822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cai S, Ye Z, Wang X, Pan Y, Weng Y, Lao S, Wei H, Li L, 2015. Overexpression of P21-activated kinase 4 is associated with poor prognosis in non-small cell lung cancer and promotes migration and invasion. J Exp Clin Cancer Res 34, 48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Callow MG, Zozulya S, Gishizky ML, Jallal B, Smeal T, 2005. PAK4 mediates morphological changes through the regulation of GEF-H1. J Cell Sci 118, 1861–1872. [DOI] [PubMed] [Google Scholar]
- Callow MG, Clairvoyant F, Zhu S, Schryver B, Whyte DB, Bischoff JR, Jallal B, Smeal T, 2002. Requirement for PAK4 in the anchorage-independent growth of human cancer cell lines. J Biol Chem 277, 550–558. [DOI] [PubMed] [Google Scholar]
- Cammarano MS, Nekrasova T, Noel B, Minden A, 2005. Pak4 induces premature senescence via a pathway requiring p16INK4/p19ARF and mitogen-activated protein kinase signaling. Mol Cell Biol 25, 9532–9542. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen C, Ha BH, Thevenin AF, Lou HJ, Zhang R, Yip KY, Peterson JR, Gerstein M, Kim PM, Filippakopoulos P, Knapp S, Boggon TJ, Turk BE, 2014. Identification of a major determinant for serine-threonine kinase phosphoacceptor specificity. Mol Cell 53, 140–147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen VB, Arendall WB 3rd, Headd JJ, Keedy DA, Immormino RM, Kapral GJ, Murray LW, Richardson JS, Richardson DC, 2010. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr D Biol Crystallogr 66, 12–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dan C, Kelly A, Bernard O, Minden A, 2001. Cytoskeletal changes regulated by the PAK4 serine/threonine kinase are mediated by LIM kinase 1 and cofilin. J Biol Chem 276, 32115–32121. [DOI] [PubMed] [Google Scholar]
- Dart AE, Wells CM, 2013. P21-activated kinase 4--not just one of the PAK. Eur J Cell Biol 92, 129–138. [DOI] [PubMed] [Google Scholar]
- Edwards DC, Sanders LC, Bokoch GM, Gill GN, 1999. Activation of LIM-kinase by Pak1 couples Rac/Cdc42 GTPase signalling to actin cytoskeletal dynamics. Nat Cell Biol 1, 253–259. [DOI] [PubMed] [Google Scholar]
- Emsley P, Lohkamp B, Scott WG, Cowtan K, 2010. Features and development of Coot. Acta Crystallogr D Biol Crystallogr 66, 486–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gnesutta N, Qu J, Minden A, 2001. The serine/threonine kinase PAK4 prevents caspase activation and protects cells from apoptosis. J Biol Chem 276, 14414–14419. [DOI] [PubMed] [Google Scholar]
- Guo Q, Su N, Zhang J, Li X, Miao Z, Wang G, Cheng M, Xu H, Cao L, Li F, 2014. PAK4 kinase-mediated SCG10 phosphorylation involved in gastric cancer metastasis. Oncogene 33, 3277–3287. [DOI] [PubMed] [Google Scholar]
- Ha BH, Boggon TJ, 2018. CDC42 binds PAK4 via an extended GTPase-effector interface. Proc Natl Acad Sci U S A 115, 531–536. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ha BH, Morse EM, Turk BE, Boggon TJ, 2015. Signaling, Regulation, and Specificity of the Type II p21-activated Kinases. J Biol Chem 290, 12975–12983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ha BH, Davis MJ, Chen C, Lou HJ, Gao J, Zhang R, Krauthammer M, Halaban R, Schlessinger J, Turk BE, Boggon TJ, 2012. Type II p21-activated kinases (PAKs) are regulated by an autoinhibitory pseudosubstrate. Proc Natl Acad Sci U S A 109, 16107–16112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hamill S, Lou HJ, Turk BE, Boggon TJ, 2016. Structural basis for non-canonical substrate recognition of cofilin/ADF proteins by LIM kinases. . Mol Cell 62, 397–408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hastie CJ, McLauchlan HJ, Cohen P, 2006. Assay of protein kinases using radiolabeled ATP: a protocol. Nature Protocols 1, 968–971. [DOI] [PubMed] [Google Scholar]
- Hofmann C, Shepelev M, Chernoff J, 2004. The genetics of Pak. J Cell Sci 117, 4343–4354. [DOI] [PubMed] [Google Scholar]
- Kaldis P, Cheng A, Solomon MJ, 2000. The effects of changing the site of activating phosphorylation in CDK2 from threonine to serine. J Biol Chem 275, 32578–32584. [DOI] [PubMed] [Google Scholar]
- Kang SA, Pacold ME, Cervantes CL, Lim D, Lou HJ, Ottina K, Gray NS, Turk BE, Yaffe MB, Sabatini DM, 2013. mTORC1 phosphorylation sites encode their sensitivity to starvation and rapamycin. Science 341, 1236566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- King H, Nicholas NS, Wells CM, 2014. Role of p-21-activated kinases in cancer progression. Int Rev Cell Mol Biol 309, 347–387. [DOI] [PubMed] [Google Scholar]
- Kumar R, Sanawar R, Li X, Li F, 2017. Structure, biochemistry, and biology of PAK kinases. Gene 605, 20–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y, Shao Y, Tong Y, Shen T, Zhang J, Li Y, Gu H, Li F, 2012. Nucleo-cytoplasmic shuttling of PAK4 modulates beta-catenin intracellular translocation and signaling. Biochim Biophys Acta 1823, 465–475. [DOI] [PubMed] [Google Scholar]
- Li Z, Zhang H, Lundin L, Thullberg M, Liu Y, Wang Y, Claesson-Welsh L, Stromblad S, 2010. p21-activated kinase 4 phosphorylation of integrin beta5 Ser-759 and Ser-762 regulates cell migration. J Biol Chem 285, 23699–23710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lu H, Liu S, Zhang G, Bin W, Zhu Y, Frederick DT, Hu Y, Zhong W, Randell S, Sadek N, Zhang W, Chen G, Cheng C, Zeng J, Wu LW, Zhang J, Liu X, Xu W, Krepler C, Sproesser K, Xiao M, Miao B, Liu J, Song CD, Liu JY, Karakousis GC, Schuchter LM, Lu Y, Mills G, Cong Y, Chernoff J, Guo J, Boland GM, Sullivan RJ, Wei Z, Field J, Amaravadi RK, Flaherty KT, Herlyn M, Xu X, Guo W, 2017. PAK signalling drives acquired drug resistance to MAPK inhibitors in BRAF-mutant melanomas. Nature. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maekawa M, Ishizaki T, Boku S, Watanabe N, Fujita A, Iwamatsu A, Obinata T, Ohashi K, Mizuno K, Narumiya S, 1999. Signaling from Rho to the actin cytoskeleton through protein kinases ROCK and LIM-kinase. Science 285, 895–898. [DOI] [PubMed] [Google Scholar]
- McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, Read RJ, 2007. Phaser crystallographic software. Journal of applied crystallography 40, 658–674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McNicholas S, Potterton E, Wilson KS, Noble ME, 2011. Presenting your structures: the CCP4mg molecular-graphics software. Acta Crystallogr D Biol Crystallogr 67, 386–394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meyer PA, Socias S, Key J, Ransey E, Tjon EC, Buschiazzo A, Lei M, Botka C, Withrow J, Neau D, Rajashankar K, Anderson KS, Baxter RH, Blacklow SC, Boggon TJ, Bonvin AM, Borek D, Brett TJ, Caflisch A, Chang CI, Chazin WJ, Corbett KD, Cosgrove MS, Crosson S, Dhe-Paganon S, Di Cera E, Drennan CL, Eck MJ, Eichman BF, Fan QR, Ferre-D'Amare AR, Fromme JC, Garcia KC, Gaudet R, Gong P, Harrison SC, Heldwein EE, Jia Z, Keenan RJ, Kruse AC, Kvansakul M, McLellan JS, Modis Y, Nam Y, Otwinowski Z, Pai EF, Pereira PJ, Petosa C, Raman CS, Rapoport TA, Roll-Mecak A, Rosen MK, Rudenko G, Schlessinger J, Schwartz TU, Shamoo Y, Sondermann H, Tao YJ, Tolia NH, Tsodikov OV, Westover KD, Wu H, Foster I, Fraser JS, Maia FR, Gonen T, Kirchhausen T, Diederichs K, Crosas M, Sliz P, 2016. Data publication with the structural biology data grid supports live analysis. Nature communications 7, 10882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller CJ, 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]
- Miller CJ, Lou HJ, Simpson C, van de Kooij B, Ha BH, Fisher OS, Pirman NL, Boggon TJ, Rinehart J, Yaffe MB, Linding R, Turk BE, 2019. Comprehensive profiling of the STE20 kinase family defines features essential for selective substrate targeting and signaling output. PLoS Biol 17, e2006540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Minden A, 2012. The pak4 protein kinase in breast cancer. ISRN oncology 2012, 694201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ohashi K, Nagata K, Maekawa M, Ishizaki T, Narumiya S, Mizuno K, 2000. Rho-associated kinase ROCK activates LIM-kinase 1 by phosphorylation at threonine 508 within the activation loop. J Biol Chem 275, 3577–3582. [DOI] [PubMed] [Google Scholar]
- Otwinowski Z, Minor W, 1997. Processing of X-ray diffraction data collected in oscillation mode, p. 307–326, in: Carter CW and Sweet RM, Eds.), Methods in Enzymology, Academic Press (New York), San Diego. [DOI] [PubMed] [Google Scholar]
- Pinna LA, Ruzzene M, 1996. How do protein kinases recognize their substrates? Biochim Biophys Acta 1314, 191–225. [DOI] [PubMed] [Google Scholar]
- Qu J, Cammarano MS, Shi Q, Ha KC, de Lanerolle P, Minden A, 2001. Activated PAK4 regulates cell adhesion and anchorage-independent growth. Mol Cell Biol 21, 3523–3533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rennefahrt UE, Deacon SW, Parker SA, Devarajan K, Beeser A, Chernoff J, Knapp S, Turk BE, Peterson JR, 2007. Specificity profiling of Pak kinases allows identification of novel phosphorylation sites. J Biol Chem 282, 15667–15678. [DOI] [PubMed] [Google Scholar]
- Scott RW, Olson MF, 2007. LIM kinases: function, regulation and association with human disease. Journal of molecular medicine 85, 555–568. [DOI] [PubMed] [Google Scholar]
- Sumi T, Matsumoto K, Nakamura T, 2001a. Specific activation of LIM kinase 2 via phosphorylation of threonine 505 by ROCK, a Rho-dependent protein kinase. J Biol Chem 276, 670–676. [DOI] [PubMed] [Google Scholar]
- Sumi T, Matsumoto K, Shibuya A, Nakamura T, 2001b. Activation of LIM kinases by myotonic dystrophy kinase-related Cdc42-binding kinase alpha. J Biol Chem 276, 23092–23096. [DOI] [PubMed] [Google Scholar]
- Sumi T, Hashigasako A, Matsumoto K, Nakamura T, 2006. Different activity regulation and subcellular localization of LIMK1 and LIMK2 during cell cycle transition. Exp Cell Res 312, 1021–1030. [DOI] [PubMed] [Google Scholar]
- Taylor SS, Radzio-Andzelm E, Hunter T, 1995. How do protein kinases discriminate between serine/threonine and tyrosine? Structural insights from the insulin receptor protein-tyrosine kinase. FASEB J 9, 1255–1266. [DOI] [PubMed] [Google Scholar]
- Taylor SS, Knighton DR, Zheng J, Ten Eyck LF, Sowadski JM, 1992. Structural framework for the protein kinase family. Annu Rev Cell Biol 8, 429–462. [DOI] [PubMed] [Google Scholar]
- Thillai K, Sarker D, Wells C, 2018. PAK4 pathway as a potential therapeutic target in pancreatic cancer. Future Oncol 14, 579–582. [DOI] [PubMed] [Google Scholar]
- Turk BE, 2008. Understanding and exploiting substrate recognition by protein kinases. Curr Opin Chem Biol 12, 4–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ubersax JA, Ferrell JE Jr., 2007. Mechanisms of specificity in protein phosphorylation. Nat Rev Mol Cell Biol 8, 530–541. [DOI] [PubMed] [Google Scholar]
- Wells CM, Jones GE, 2010. The emerging importance of group II PAKs. Biochem J 425, 465–473. [DOI] [PubMed] [Google Scholar]
- Wells CM, Whale AD, Parsons M, Masters JR, Jones GE, 2010. PAK4: a pluripotent kinase that regulates prostate cancer cell adhesion. J Cell Sci 123, 1663–1673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wong LE, Reynolds AB, Dissanayaka NT, Minden A, 2010. p120-catenin is a binding partner and substrate for Group B Pak kinases. Journal of cellular biochemistry 110, 1244–1254. [DOI] [PubMed] [Google Scholar]
- Wu H, Wang ZX, 2003. The mechanism of p21-activated kinase 2 autoactivation. J Biol Chem 278, 41768–41778. [DOI] [PubMed] [Google Scholar]
- Xu HT, Lai WL, Liu HF, Wong LL, Ng IO, Ching YP, 2016. PAK4 Phosphorylates p53 at Serine 215 to Promote Liver Cancer Metastasis. Cancer Res 76, 5732–5742. [DOI] [PubMed] [Google Scholar]
- Zhang EY, Ha BH, Boggon TJ, 2018. PAK4 crystal structures suggest unusual kinase conformational movements. Biochim Biophys Acta 1866, 356–365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao M, Spiess M, Johansson HJ, Olofsson H, Hu J, Lehtio J, Stromblad S, 2017. Identification of the PAK4 interactome reveals PAK4 phosphorylation of N-WASP and promotion of Arp2/3-dependent actin polymerization. Oncotarget 8, 77061–77074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhuang T, Zhu J, Li Z, Lorent J, Zhao C, Dahlman-Wright K, Stromblad S, 2015. p21-activated kinase group II small compound inhibitor GNE-2861 perturbs estrogen receptor alpha signaling and restores tamoxifen-sensitivity in breast cancer cells. Oncotarget 6, 43853–43868. [DOI] [PMC free article] [PubMed] [Google Scholar]