Significance
Molecular recognition via protein–ligand binding is essential for biomolecular functions. Current understanding of ligand recognition is mostly brought from a bound state structure resolved by X-ray crystallography. Using molecular dynamics simulation with enhanced sampling techniques, we here characterize multiple bound poses and transient encounter complexes of an inhibitor kinase on the statistically converged free-energy landscapes. Our simulations propose parallel binding pathways connected through encounter complexes, highlighting the importance of early binding processes. Determination of transient encounter complexes is still difficult experimentally, and the present finding helps to understand the process and widens the possibility of drug compound design.
Keywords: protein–ligand interaction, protein kinase, free-energy landscape, molecular dynamics simulation, replica-exchange molecular dynamics
Abstract
Modern drug discovery increasingly focuses on the drug-target binding kinetics which depend on drug (un)binding pathways. The conventional molecular dynamics simulation can observe only a few binding events even using the fastest supercomputer. Here, we develop 2D gREST/REUS simulation with enhanced flexibility of the ligand and the protein binding site. Simulation (43 μs in total) applied to an inhibitor binding to c-Src kinase covers 100 binding and unbinding events. On the statistically converged free-energy landscapes, we succeed in predicting the X-ray binding structure, including water positions. Furthermore, we characterize hidden semibound poses and transient encounter complexes on the free-energy landscapes. Regulatory residues distant from the catalytic core are responsible for the initial inhibitor uptake and regulation of subsequent bindings, which was unresolved by experiments. Stabilizing/blocking of either the semibound poses or the encounter complexes can be an effective strategy to optimize drug-target residence time.
Drug compounds have been traditionally designed by optimizing their equilibrium binding affinity (e.g., dissociation constant Kd) to a target protein. Many designed compounds still lack in vivo efficacy, raising the importance of kinetics-based compound design (1, 2). In the past decade, the “residence time concept” has been extensively tested, where the drug compound is design to reside in the vicinity of the target protein (3). This kinetics-based design can potentially increase the selectivity to the target protein and also better integrate with pharmacodynamics and pharmacokinetics design (4). In addition to the bound state structures, obtaining an exhaustive mapping of the drug binding landscape is also desired. However, obtaining such mapping experimentally at the atomic level still remains a significant challenge.
Protein kinases (PKs) regulate signal transductions in cells and their dysfunction can cause several diseases, making them an important drug target, particularly for cancer treatment (5, 6). Designing a selective PK inhibitor is a significant challenge because the inhibitor competitively binds to a highly conserved ATP pocket (Fig. 1) (7). The regulatory residues forming the pocket involve the hinge and the gatekeeper residues (catalytic core), where ATP directly binds, as well as a Gly-rich loop (G-loop or P-loop), an α-helix (αC-helix), and an activation loop (A-loop). The current inhibitor design heavily relies on the atomic structures of the ATP binding site, particularly those of the catalytic core, at bound states obtainable from X-ray crystallography (8, 9). However, the inhibitor selectivity is often a result of the combined contributions of the regulatory residues (10, 11), and the regulation mechanism remains unknown. The residues distant from the catalytic core could play an important role during the binding processes besides the bound state stabilization. Recent experiments using NMR and fast kinetics measurements propose a 2-step binding model consisting of a fast binding followed by a slow induced fit process (12). Elucidating the structural features of binding processes is critical for understanding the regulation mechanism at the atomic level. Yet, structures of transient intermediates such as encounter complexes are still unknown.
Molecular dynamics (MD) simulations, which provide atomic-level details of dynamical processes, are increasingly used for calculating drug-target binding pathways and kinetics, as well as thermodynamics (13–16). Nowadays, MD simulations are able to reach microsecond timescales and observe binding events. Binding pathways and the kinetics extracted from these simulations, however, suffer from large statistical uncertainty, because it is difficult to observe multiple binding and unbinding events within the available computational time, even when using specialized hardware (17, 18). In overcoming the difficulty, parallel trajectory approaches predict binding pathways and kinetics at longer timescales using short-time simulation data (19–22), while various enhanced sampling methods characterize drug-target binding, with a particular focus on unbinding kinetics, by effectively crossing high energy barriers (23–31). Currently, unbinding kinetics and their molecular determinants can be reasonably estimated and directly compared with experiments, though obtaining converged free-energy landscapes to map all of the key intermediates along the pathway remains a challenge. In this work, we focus on the 2D replica-exchange method—REST/REUS, a combined replica-exchange umbrella sampling (REUS) (32, 33) and replica exchange with solute tempering (REST) (34, 35). This combined approach complements drawbacks in each method to further enhance the sampling and can precisely predict protein–ligand bound structures ab initio (36). Here, we extend the method to introduce the flexibility of the protein binding site as well as that of the ligand by utilizing the generalized REST (gREST) method (37). The approach, which we call gREST/REUS, realizes extensive sampling of binding and unbinding events and provides converged free-energy landscapes, which allows us to characterize binding poses on multiple binding pathways with unprecedented accuracy.
We apply the method to investigate the binding of PP1 (ATP-competitive inhibitor) to c-Src kinase, a key signaling kinase in the cancer process (38, 39). A previous microsecond timescale unbiased MD simulation correctly predicted the structure of c-Src kinase complexed with either PP1 or dasatinib as observed in the X-ray crystal structures (17). The simulation shows that PP1 resides for a long time in the vicinity of the N-lobe before binding through a water-separated intermediate. A subsequent free-energy calculation shows that desolvation of the binding pocket is the main contributor to the free-energy barrier (40). More recently, a metadynamics-based unbinding simulation gave further insight into the protein–water coupled dasatinib binding (24). The study reports the koff value and relates it to cleavages of the evolutionally conserved Lys-Glu salt bridge and the dasatinib-hinge hydrogen bonds. In this work, we characterize the structures and interactions of multiple binding poses, transient intermediates, encounter complexes, and pathways connecting them, on the free-energy landscapes. The calculated energetics and kinetics agree with experiments as well as with previous computations. We integrate those findings and propose a parallel binding model. We highlight the role of G-loop residues distant from the catalyst core at the early stage of the binding process, i.e., the formation of encounter complexes, which was unresolved by a mutagenesis experiment and X-ray crystal structures (10). We believe that this work provides an effective strategy to optimize drug efficacy with prolonged target residence times.
Results
Efficiency of gREST/REUS Simulation.
The details of gREST/REUS simulation are described in Materials and Methods and SI Appendix. In brief, gREST/REUS simulation conducts replica exchanges in 2 dimensions: exchanges of the PP1-Src distance umbrella potentials (REUS dimension), and exchanges of solute temperatures (gREST dimension). The solute is defined as the dihedral angle and the nonbonded energy terms of both PP1 and 10 binding site residues, based on the X-ray crystal structure (Fig. 1 and SI Appendix, Fig. S1). The current choice of the solute enhances the flexibility of protein binding site residues (SI Appendix, Table S1) and significantly increases the chance of binding events to occur. In 78% of the replicas, the minimum heavy atoms rsmd between the ligand PP1 and the X-ray crystal structure is <1 Å (Fig. 2A, SI Appendix, Table S2 and Figs. S2–S6, and Movies S1 and S2). Simulating 100 binding and unbinding events, obtaining convergence of the protein structure (SI Appendix, Fig. S7), we are able to characterize the bound state and transient intermediates on the statistically converged free-energy landscapes (Fig. 2B).
Multiple Binding Poses of the Inhibitor and Water Molecules.
First we identify multiple binding poses by standard k-means clustering analysis, which uses Euclidean distance function to classify the structures in the trajectory data at 310 K into a predefined number of clusters (10 in the present study) (SI Appendix, Fig. S8). In Fig. 2 C and D, we show 5 representative poses of the bound region (ξ < 6 Å). The major pose I, which dominates the bound region of ξ < 4 Å, is in excellent agreement with the X-ray structure (PDBID: 1QCF) (41). PP1 occupies both the adenine and the back pockets and forms highly conserved hydrogen bonds with the hinge residues: 1 with the side chain of the gatekeeper residue (Thr80), and an additional 2 with the main chain of the hinge residues (Glu81and Met83) (8). We also find 4 semibound poses in the region of ξ = 5 ∼ 6 Å, which are not resolved in the X-ray structure. Pose IV has the same PP1 orientation as pose I, while the other poses (II, III, and V) have different orientations. In the semibound poses, PP1 occupies only the adenine pocket, and the hydrogen bonds with the hinge residues are mostly absent.
The volmap feature in the Visual Molecular Dynamics (VMD) software (isosurface at 20% occupancy) provides water distribution in the binding pocket in each pose (Fig. 2E) (42). The water distribution in pose I overlaps with water positions in the X-ray structure. In the other poses, a layer of water molecules in the back pocket separates PP1 and the binding site residues and breaks the hydrogen bonds with Thr80 and Glu81. The semibound pose IV agrees well with a prebound pose found in the earlier conventional MD simulation (17). The previous free energy analysis shows that desolvation of the binding pocket mainly contributes to the free-energy barrier for the dasatinib binding to the Src kinase (40). Given that the water molecules deep inside the pocket are absent only in the pose I, desolvation of the pocket likely relates to the molecular orientation of the inhibitor.
Encounter Complexes in Parallel Pathways.
We can confidently characterize the binding pathways with multiple intermediates, including encounter complexes, by constructing free-energy landscapes from the simulation data. For this purpose, a set of 2D free-energy landscapes was constructed along the reaction coordinate, chosen as the PP1-Src distance (ξ), and either the PP1 position or orientation with respect to the protein, defined using 6 anchor atoms (P1, P2, and P3 from the protein and L1, L2, and L3 from PP1). For example, P1 is defined as the center of mass (COM) of the protein backbone heavy atoms of the residue closest to the COM of the protein (SI Appendix for details). The position and orientation of PP1 are defined by polar angles, θ (P2-P1-L1) and φ (P3-P2-P1-L1), and Euler angles, α (P1-L1-L2), β (P2-P1-L1-L2), and γ (P1-L1-L2-L3), respectively (Fig. 3A). The free-energy landscapes along the PP1-Src distance (ξ) and position, (θ, ξ) and (φ, ξ), are smooth and the predominant binding pathway is well defined (Fig. 3B). In the (θ, ξ) landscape, we also find a secondary path originating from the large θ value (∼120°), in addition to the dominant path from the small θ value (∼60°). On the contrary, 3 landscapes along the PP1-Src distance (ξ) and orientation, (α/β/γ, ξ), are relatively rugged (Fig. 3C). This means that PP1 enters the pocket through a unique azimuth angle but in various molecular orientations.
In order to characterize the binding pathways, we classify the bound poses in Fig. 2D (Poses I ∼ V) by PP1 orientation: Bzz (around β = 0° and γ = 0°), Bzp (around β = 0° and γ = 180°), Bpz (around β = 180° and γ = 0°), and Bpp (around β = 180° and γ = 180°) (SI Appendix, Fig. S9). Given that pose IV, having the same orientation as pose I, is an intermediate toward pose I, there are actually 4 distinctive poses (thus we denote pose IV as Bzz*). In the free-energy landscapes in Fig. 3C, the canonical pose Bzz (= pose I) is located around ξ = 3 Å, whereas the semibound poses (Bzp, Bpz, and Bpp) and the intermediate pose (Bzz*) are found in the region of ξ = 5 ∼ 6 Å. Because of the energy barriers dividing these poses in Fig. 3C, multiple poses may not be directly exchangeable. To confirm this, we perform independent microsecond-long conventional MD simulations (total 15 μs), where PP1 is bound after 4 μs and 3.2 μs in 2 of the 3 simulations (SI Appendix). The analysis of simulation trajectories supports that the multiple poses are not mutually exchanged in the bound region. For example, the trajectory leading to the semibound pose, Bpz, dissociates without changing to other poses after reaching the bound state (SI Appendix, Fig. S10). The plausible parallel binding pathways to multiple poses are indicated in the free-energy landscapes in Fig. 3C. We also find distinct pathways to multiple poses in the 3D probability distribution (Movies S3 and S4). The projection of free-energy landscapes onto the pairs of reaction coordinates is useful to understand molecular mechanisms underlying the parallel binding pathways. The combined kinetic network models and flux analysis could help to rigorously characterize the pathways based on kinetic flux (43), although more extensive conformational sampling is necessary. Parallel binding pathways are observed for the ligand in other proteins, such as p38 MAP kinase (28), bromodomain (20), and LAO protein (22) and may be a general feature of the ligand binding process (44).
Using the k-means clustering analysis of the regions excluding the bound region (ξ > 6 Å) (SI Appendix, Fig. S8), we extract the representative intermediate and encounter complex structures along the pathways to different poses, and their positions are superposed on the free-energy landscapes in Fig. 3C. Multiple bound states (Bzz, Bzp, Bpz, and Bpp) are connected only through their own encounter states. The 4 parallel binding pathways are sketched in Fig. 4A. In 3 of them, including the major pathway (Ezz → Bzz), PP1 enters the binding pocket from the G-loop side. In the Epz → Bpz pathway on the contrary, PP1 directly accesses the hinge region. The latter corresponds to the path approaching from the large θ value (∼120°) in the free-energy landscape for the PP1 position (Fig. 3B).
Fig. 4B illustrates the binding site interactions along the pathways. In the encounter complexes (Ezz, Ezp, and Epp), PP1 interacts with the highly conserved GxGxxG motif in the G-loop (8). We find an additional hydrogen bond with the side chain of Asp146 of the conserved Asp-Phe-Gly (DFG) motif in Ezz, which likely fixes PP1 to the canonical pose orientation. Only in Epz, PP1 directly accesses the pocket by forming hydrogen bonds with the hinge residues (the side chain of Tyr82 and the main chain of Ser84). In a subsequent step toward the canonical pose (Ezz → Bzz), PP1–G-loop interactions are replaced with hydrogen bonds to the hinge residues (the side chain of Tyr82 and the main chain of Met83), where Val23 appears to assist in retaining PP1 inside the pocket. Accompanied with coupled desolvation of the binding pocket, these hydrogen bonds finally rearrange to produce the well-conserved hydrogen bonds with the gatekeeper (Thr80) and the hinge residues (Glu81and Met83), stabilizing the canonical bound state Bzz (= pose I). An interaction switch from PP1–G-loop to PP1–hinge residues is observed in the other pathways (Ezp → Bzp and Epp → Bpp), but these PP1 orientations preclude hydrogen bond formation. Hence, multiple bound states are commonly stabilized by interaction with the hinge residues, but the interaction mechanism is different. Evidently, the G-loop residues play a major role in stabilizing the encounter complexes. The G-loop displays a remarkably high flexibility, but its fluctuations are largely suppressed by the formation of encounter complexes (SI Appendix, Fig. S11). In addition, upon going from the encounter complex to the bound state, Phe20 reorients to fill the space originally occupied by PP1 (SI Appendix, Fig. S12). Thus, the G-loop also contributes to the secondary stabilization of the bound state.
Along the main pathway (Ezz → Bzz), we find the highest free-energy region in between Ezz and Izz (ξ = 8 ∼ 9 Å). We create an ensemble of the transition state structures by extracting 1,600 structures that satisfy the following conditions from the simulation trajectory at 310 K: 8 Å < ξ < 9 Å, 60° < θ < 100°, 60° < α < 120°, 60° < β < 100°, and 60° < γ < 100°. The k-means clustering, with the input number of clusters equal to 5, provides 2 representative structures at ξ ∼ 8.9 Å (77.1%) and ξ ∼ 8.2 Å (22.9%) (Fig. 4C). In the former, the conserved Lys37-Glu52 salt bridge is maintained (Lys37:Cδ − Glu52:Cε distance is 4.4 Å), while in the latter, the salt bridge is broken and instead Lys37 forms a hydrogen bond with Asp146, releasing PP1. Concomitantly, the αC-helix transiently rotates outward, opening the access to the binding pocket. This feature of the transition state region is consistent with the molecular switch mechanism proposed for the dasatinib-Src kinase binding (24). The mutagenesis experiment shows that the Src residues stabilizing the inactive αC-helix-out conformation is sensitive to the inhibitor binding (10). The coupled salt bridge and αC-helix motion is a key feature of the transition state for inhibitor-kinase binding.
Binding Energy and Kinetics.
The present simulation describes half of the binding process: i.e., from the encounter to the bound state (intrusion into the binding pocket). The corresponding free-energy change, ΔGintrusion, is −3.6 kcal/mol, obtained by integrating the bound (ξ = 2 ∼ 10 Å) and unbound (ξ = 10 ∼ 16 Å) regions of the free-energy profile in Fig. 2B and taking their ratio. The value of ΔGintrusion displays the increase to −4.3 kcal/mol when increasing the boundary distance up to ξ = 14 Å (bound region: ξ = 2 ∼ 14 Å, unbound region: ξ = 14 ∼ 16 Å). The estimated ΔGintrusion corresponds to 2,900 μM, which is significantly larger than the experimental IC50 (170 nM) (45). An independent free-energy perturbation calculation produces a binding free energy (ΔGbind) value of −8.6 kcal/mol (Kd = 808 nM), which reproduces the experimental IC50 to a reasonable extent (SI Appendix, Table S3). The difference between ΔGbind and ΔGintrusion is about −5 kcal/mol, which corresponds to the free-energy change associated with the encounter complex formation. The result is consistent with the finding on Gleevec for which the physical binding step contributes only micromolar affinities (−6.5 kcal/mol) for both Abl and Src (12). However, the PP1 is small and its intrusion is accompanied by subtle changes in the binding site: the transient breaking of the Lys37-Glu52 salt bridge (Fig. 4C) and the reorientation of Phe20 (SI Appendix, Fig. S12). This may contrast to the Gleevec-Abl kinase binding for which a large conformational change (a slow induced-fit process) is suggested (12).
To link with binding kinetics, we analyzed the data of independently performed microsecond-long conventional MD simulations (SI Appendix). The estimated kon value is 4.6 μM−1 agreeing with the previous computational estimate (17) and the experimental value (∼5 μM−1) (46). This value, combined with the calculated Kd (808 nM) produces a koff value of 3.7 s−1 using the Kd = koff/kon relation, resulting in a residence time (1/koff) of 0.3 s. The reported residence time for dasatinib-Src binding, with a dissociation constant of 11 nM, is 21 ± 10 s as estimated by simulation (24) and 18 s to 900 s by experiments (47, 48). Considering that the binding of PP1 to the Src kinase is weaker than that of dasatinib, the estimated residence time is reasonable.
Discussion
Precision Drug-Target Binding Simulations.
Drug-target binding simulations have evolved from predicting binding affinities to mapping binding pathways and kinetics. The present gREST/REUS simulations realize multiple binding and unbinding events, achieved by enhanced flexibilities of the inhibitor and the protein binding site. The simulations display faster statistical convergence than in the original method, REST/REUS and provide reliable free-energy landscapes for understanding molecular mechanisms. Currently, simulations are limited to sampling from the encounter region to the bound state, and thus direct estimation of kinetic parameters is not feasible. However, the reversible sampling of the binding and unbinding events using gREST/REUS serves as the basis for accurate estimation of thermodynamics and kinetics, complementary with other methods. For example, the gREST/REUS may be integrated with Markov state modeling (MSM) with the transition-based reweighting analysis method (TRAM) (49, 50) to significantly improve the efficiency of estimating drug-target unbinding kinetics. Alternatively, gREST/REUS simulations can be combined with the Brownian dynamics approach (51, 52) or multiple independent MD, such as Parallel Cascade Selection Molecular Dynamics (19) to provide both an atomic level description of the binding pathway and kinetic parameters at high accuracy. The present 2D replica exchange approach requires large computational resources, which prevent routine usages. For a relatively fast binding process, 1-dimensional gREST may be successful. Another possibility is to replace the gREST with a nonreplica type enhanced sampling method (53–55), maintaining sampling efficiency while reducing the total number of replicas. The simplicity of the present scheme could serve as a basis for future complementary approaches using various types of enhanced sampling methods for precision drug-target binding simulations.
Encounter Complexes with G-Loop.
Traditional drug compound design has relied on the structures and interactions at the bound state resolved by X-ray crystallography. The analysis of free-energy landscapes illuminates the roles of regulatory residues in the binding process, which are not obtainable from the bound state structure. Here, we succeeded in predicting the encounter complexes on the multiple binding pathways for PP1-Src binding. The formation of an encounter complex is generally considered as a key step guiding the ligand toward a final bound state (56), while it is difficult to observe and characterize the transient encounter complexes experimentally (57). Our simulation shows that the G-loop grabs PP1 by forming encounter complexes, and the interaction between the G-loop and PP1 decides whether to enter the pocket in the right orientation. In previous studies, the interaction with the G-loop was found to modulate the inhibitor selectivity (10, 11, 58, 59). The mutagenesis experiment shows the combined contribution of the G-loop and other regulatory residues to the observed selectivity (10). Our simulation shows that the conserved G-loop is preferentially responsible for the initial uptake of the inhibitor and the regulation of subsequent bindings. This mechanism is applicable to several inhibitor-kinase bindings. Whether the encounter states have some influence on the binding selectivity or not is an open question for future study.
Implication for Design.
The encounter complexes and parallel binding pathways characterized in this work can provide new perspectives for the inhibitor design. Our results suggest that the modulation of encounter complexes can be an effective strategy for controlling the inhibitor binding, in addition to control of the rate-limiting step thorough the conserved salt bridge (24). The stabilization of encounter complexes can extend the residence time (lower the koff) without affecting the association rate (kon). The enhanced stability of the encounter complex may also retard the dissociation from the semibound poses and can help extend the residence time. The validity of this idea can be seen in the recently designed Mps1/TTK kinase inhibitors, where the shift of the G-loop position is related to the slow binding kinetics (60). It is likely that the G-loop traps the inhibitor at an encounter state, resulting in the observed slow kinetics. Conversely, by designing the compound so as to destabilize the encounter complexes against off-target proteins, it is possible to effectively increase the binding to the target protein. Thus, the possibility of inhibitor design would be widened by targeting the encounter complexes. The encounter complexes characterized in this work are only those directly leading to the bound states, but in fact, the inhibitor can form encounter complexes at multiple sites on the protein surface (17). Further analysis of inhibitor dynamics including these encounter complexes, using methods such as MSM, may provide yet another strategy for extending the residence time. Our simulation predicts the existence of semibound poses that are not resolved in the X-ray crystal structure. The interactions with the surrounding residues of the semibound poses are different from that of the canonical pose. It would be possible to control the inhibitor binding by stabilizing/blocking the semibound poses. As the binding process becomes explored, the possibility of ATP-competitive inhibitor design will greatly expand and help to overcome the difficulty in the traditional design based on the X-ray crystal structures.
Materials and Methods
MD Simulation.
The initial configuration of the PP1-Src complex was constructed using the X-ray structures of the unphosphorylated c-Src in active conformation (PDBID: 1Y57) (61) and of the autoinhibited form of Hck in complex with PP1 (PDBID: 1QCF) (41). The PP1-Src complex was solvated by water molecules and neutralized by 6 sodium cations. All simulations were performed using a development version of the GENESIS program package (62, 63). We used the AMBER ff99SB-ILDN (64, 65) and TIP3P (66) parameters for the protein and water molecules, respectively. The ligand parameters were obtained by using GAFF with AM1-BCC (67).
Two-Dimensional gREST/REUS Simulation.
We defined the solute region in gREST as the dihedral angle and nonbonded energy terms of an inhibitor PP1 and 10 binding site residues, based on the X-ray crystal structure. Eight replicas were used to cover the solute temperature range of 310 ∼ 3,100 K (gREST dimension). The reaction coordinate (ξ) of the umbrella sampling simulation (REUS dimension) was taken as the distance between COMs of PP1 and the backbone heavy atoms of 2 binding site residues (Ala35 and Leu135). Eighteen replicas were used to cover the ξ values in the range of 3.0 Å ∼ 15.0 Å. The total number of replicas in gREST/REUS simulation is 18 × 8 = 144. Replica exchange was attempted every 2 ps alternatively for the REUS and gREST dimensions. Following 1 ns of equilibration run for each replica, gREST/REUS for 300 ns per replica (43.2 μs in total) was performed for the purpose of analyzing the PP1-Src binding.
Additional details are described in SI Appendix.
Supplementary Material
Acknowledgments
We thank Dr. Ai Shinobu for her critical reading of the manuscript. We acknowledge the computing time granted by RIKEN Advanced Center for Computing and Communication (HOKUSAI GreatWave and BigWaterfall), RIKEN Center for Computational Science (K Computer), and the High Performance Computing Infrastructure system (Project ID: hp150270, hp160207, hp170115, hp170254, hp180201, hp180274, and hp190181). This research was supported by MEXT as “Priority Issue on Post-K Computer” (Building Innovative Drug Discovery Infrastructure Through Functional Control of Biomolecular Systems) (to Y.S.), MEXT Grant-in-Aid for Scientific Research on Innovative Areas Grant No. 26119006 (to Y.S.), MEXT/JSPS KAKENHI Grant Nos. 26220807 and 19H05645 (to Y.S.) and 19K12229 (to S.R.), RIKEN pioneering project “Dynamic Structural Biology” (to Y.S.), and Center of Innovation Program from Japan Science and Technology Agency (to Y.S.).
Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission. Y.S. is a guest editor invited by the Editorial Board.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1904707116/-/DCSupplemental.
References
- 1.Copeland R. A., The drug-target residence time model: A 10-year retrospective. Nat. Rev. Drug Discov. 15, 87–95 (2016). [DOI] [PubMed] [Google Scholar]
- 2.Folmer R. H. A., Drug target residence time: A misleading concept. Drug Discov. Today 23, 12–16 (2018). [DOI] [PubMed] [Google Scholar]
- 3.Schuetz D. A., et al. , Kinetics for drug discovery: An industry-driven effort to target drug residence time. Drug Discov. Today 22, 896–911 (2017). [DOI] [PubMed] [Google Scholar]
- 4.Dahl G., Akerud T., Pharmacokinetics and the drug-target residence time concept. Drug Discov. Today 18, 697–707 (2013). [DOI] [PubMed] [Google Scholar]
- 5.Zhang J., Yang P. L., Gray N. S., Targeting cancer with small molecule kinase inhibitors. Nat. Rev. Cancer 9, 28–39 (2009). [DOI] [PubMed] [Google Scholar]
- 6.Ferguson F. M., Gray N. S., Kinase inhibitors: The road ahead. Nat. Rev. Drug Discov. 17, 353–377 (2018). [DOI] [PubMed] [Google Scholar]
- 7.Wu P., Nielsen T. E., Clausen M. H., Small-molecule kinase inhibitors: An analysis of FDA-approved drugs. Drug Discov. Today 21, 5–10 (2016). [DOI] [PubMed] [Google Scholar]
- 8.van Linden O. P. J., Kooistra A. J., Leurs R., de Esch I. J. P., de Graaf C., KLIFS: A knowledge-based structural database to navigate kinase-ligand interaction space. J. Med. Chem. 57, 249–277 (2014). [DOI] [PubMed] [Google Scholar]
- 9.Müller S., Chaikuad A., Gray N. S., Knapp S., The ins and outs of selective kinase inhibitor development. Nat. Chem. Biol. 11, 818–821 (2015). [DOI] [PubMed] [Google Scholar]
- 10.Seeliger M. A., et al. , c-Src binds to the cancer drug imatinib with an inactive Abl/c-Kit conformation and a distributed thermodynamic penalty. Structure 15, 299–311 (2007). [DOI] [PubMed] [Google Scholar]
- 11.Brandvold K. R., Steffey M. E., Fox C. C., Soellner M. B., Development of a highly selective c-Src kinase inhibitor. ACS Chem. Biol. 7, 1393–1398 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Agafonov R. V., Wilson C., Otten R., Buosi V., Kern D., Energetic dissection of Gleevec’s selectivity toward human tyrosine kinases. Nat. Struct. Mol. Biol. 21, 848–853 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Bruce N. J., Ganotra G. K., Kokh D. B., Sadiq S. K., Wade R. C., New approaches for computing ligand-receptor binding kinetics. Curr. Opin. Struct. Biol. 49, 1–10 (2018). [DOI] [PubMed] [Google Scholar]
- 14.Dickson A., Tiwary P., Vashisth H., Kinetics of ligand binding through advanced computational approaches: A review. Curr. Top. Med. Chem. 17, 2626–2641 (2017). [DOI] [PubMed] [Google Scholar]
- 15.De Vivo M., Masetti M., Bottegoni G., Cavalli A., Role of molecular dynamics and related methods in drug discovery. J. Med. Chem. 59, 4035–4061 (2016). [DOI] [PubMed] [Google Scholar]
- 16.Bernetti M., Cavalli A., Mollica L., Protein-ligand (un)binding kinetics as a new paradigm for drug discovery at the crossroad between experiments and modelling. MedChemComm 8, 534–550 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Shan Y., et al. , How does a drug molecule find its target binding site? J. Am. Chem. Soc. 133, 9181–9183 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Dror R. O., et al. , Pathway and mechanism of drug binding to G-protein-coupled receptors. Proc. Natl. Acad. Sci. U.S.A. 108, 13118–13123 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Tran D. P., Kitao A., Dissociation process of a MDM2/p53 complex investigated by parallel cascade selection molecular dynamics and the Markov state model. J. Phys. Chem. B 123, 2469–2478 (2019). [DOI] [PubMed] [Google Scholar]
- 20.Dickson A., Mapping the ligand binding landscape. Biophys. J. 115, 1707–1719 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Plattner N., Noé F., Protein conformational plasticity and complex ligand-binding kinetics explored by atomistic simulations and Markov models. Nat. Commun. 6, 7653 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Silva D. A., Bowman G. R., Sosa-Peinado A., Huang X., A role for both conformational selection and induced fit in ligand binding by the LAO protein. PLoS Comput. Biol. 7, e1002054 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Patel J. S., Berteotti A., Ronsisvalle S., Rocchia W., Cavalli A., Steered molecular dynamics simulations for studying protein-ligand interaction in cyclin-dependent kinase 5. J. Chem. Inf. Model. 54, 470–480 (2014). [DOI] [PubMed] [Google Scholar]
- 24.Tiwary P., Mondal J., Berne B. J., How and when does an anticancer drug leave its binding site? Sci. Adv. 3, e1700014 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Schuetz D. A., et al. , Predicting residence time and drug unbinding pathway through scaled molecular dynamics. J. Chem. Inf. Model. 59, 535–549 (2019). [DOI] [PubMed] [Google Scholar]
- 26.Miao Y., McCammon J. A., Mechanism of the G-protein mimetic nanobody binding to a muscarinic G-protein-coupled receptor. Proc. Natl. Acad. Sci. U.S.A. 115, 3036–3041 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Guo D., et al. , Molecular basis of ligand dissociation from the adenosine A2A receptor. Mol. Pharmacol. 89, 485–491 (2016). [DOI] [PubMed] [Google Scholar]
- 28.Casasnovas R., Limongelli V., Tiwary P., Carloni P., Parrinello M., Unbinding kinetics of a p38 MAP kinase type II inhibitor from metadynamics simulations. J. Am. Chem. Soc. 139, 4780–4788 (2017). [DOI] [PubMed] [Google Scholar]
- 29.Spitaleri A., Decherchi S., Cavalli A., Rocchia W., Fast dynamic docking guided by adaptive electrostatic bias: The MD-binding approach. J. Chem. Theory Comput. 14, 1727–1736 (2018). [DOI] [PubMed] [Google Scholar]
- 30.Gobbo D., et al. , Investigating drug-target residence time in kinases through enhanced sampling simulations. J. Chem. Theory Comput., 10.1021/acs.jctc.9b00104 (2019). [DOI] [PubMed] [Google Scholar]
- 31.Lin Y.-L., Meng Y., Jiang W., Roux B., Explaining why Gleevec is a specific and potent inhibitor of Abl kinase. Proc. Natl. Acad. Sci. U.S.A. 110, 1664–1669 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Sugita Y., Kitao A., Okamoto Y., Multidimensional replica-exchange method for free-energy calculations. J. Chem. Phys. 113, 6042–6051 (2000). [Google Scholar]
- 33.Fukunishi H., Watanabe O., Takada S., On the Hamiltonian replica exchange method for efficient sampling of biomolecular systems: Application to protein structure prediction. J. Chem. Phys. 116, 9058–9067 (2002). [Google Scholar]
- 34.Liu P., Kim B., Friesner R. A., Berne B. J., Replica exchange with solute tempering: A method for sampling biological systems in explicit water. Proc. Natl. Acad. Sci. U.S.A. 102, 13749–13754 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Wang L., Friesner R. A., Berne B. J., Replica exchange with solute scaling: A more efficient version of replica exchange with solute tempering (REST2). J. Phys. Chem. B 115, 9431–9438 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Kokubo H., Tanaka T., Okamoto Y., Two-dimensional replica-exchange method for predicting protein-ligand binding structures. J. Comput. Chem. 34, 2601–2614 (2013). [DOI] [PubMed] [Google Scholar]
- 37.Kamiya M., Sugita Y., Flexible selection of the solute region in replica exchange with solute tempering: Application to protein-folding simulations. J. Chem. Phys. 149, 072304 (2018). [DOI] [PubMed] [Google Scholar]
- 38.Thomas S. M., Brugge J. S., Cellular functions regulated by Src family kinases. Annu. Rev. Cell Dev. Biol. 13, 513–609 (1997). [DOI] [PubMed] [Google Scholar]
- 39.Martin G. S., The hunting of the Src. Nat. Rev. Mol. Cell Biol. 2, 467–475 (2001). [DOI] [PubMed] [Google Scholar]
- 40.Mondal J., Friesner R. A., Berne B. J., Role of desolvation in thermodynamics and kinetics of ligand binding to a kinase. J. Chem. Theory Comput. 10, 5696–5705 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Schindler T., et al. , Crystal structure of Hck in complex with a Src family-selective tyrosine kinase inhibitor. Mol. Cell 3, 639–648 (1999). [DOI] [PubMed] [Google Scholar]
- 42.Humphrey W., Dalke A., Schulten K., VMD: Visual molecular dynamics. J. Mol. Graph. 14, 33–38, 27–28 (1996). [DOI] [PubMed] [Google Scholar]
- 43.Gu S., Silva D.-A., Meng L., Yue A., Huang X., Quantitatively characterizing the ligand binding mechanisms of choline binding protein using Markov state model analysis. PLoS Comput. Biol. 10, e1003767 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Vakser I. A., Low-resolution docking: Prediction of complexes for underdetermined structures. Biopolymers 39, 455–464 (1996). [DOI] [PubMed] [Google Scholar]
- 45.Hanke J. H., et al. , Discovery of a novel, potent, and Src family-selective tyrosine kinase inhibitor. Study of Lck- and FynT-dependent T cell activation. J. Biol. Chem. 271, 695–701 (1996). [DOI] [PubMed] [Google Scholar]
- 46.Shan Y., et al. , A conserved protonation-dependent switch controls drug binding in the Abl kinase. Proc. Natl. Acad. Sci. U.S.A. 106, 139–144 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Kwarcinski F. E., et al. , Conformation-selective analogues of dasatinib reveal insight into kinase inhibitor binding and selectivity. ACS Chem. Biol. 11, 1296–1304 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Getlik M., et al. , Hybrid compound design to overcome the gatekeeper T338M mutation in cSrc. J. Med. Chem. 52, 3915–3926 (2009). [DOI] [PubMed] [Google Scholar]
- 49.Wang W., Cao S., Zhu L., Huang X., Constructing Markov State Models to elucidate the functional conformational changes of complex biomolecules. Wiley Interdiscip. Rev. Comput. Mol. Sci. 8, e1343 (2018). [Google Scholar]
- 50.Wu H., Paul F., Wehmeyer C., Noé F., Multiensemble Markov models of molecular thermodynamics and kinetics. Proc. Natl. Acad. Sci. U.S.A. 113, E3221–E3230 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Mereghetti P., Gabdoulline R. R., Wade R. C., Brownian dynamics simulation of protein solutions: Structural and dynamical properties. Biophys. J. 99, 3782–3791 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Votapka L. W., Amaro R. E., Multiscale estimation of binding kinetics using brownian dynamics, molecular dynamics and milestoning. PLoS Comput. Biol. 11, e1004381 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Marinari E., Parisi G., Simulated tempering: A new Monte Carlo scheme. Europhys. Lett. 19, 451–458 (1992). [Google Scholar]
- 54.Miao Y., Feher V. A., McCammon J. A., Gaussian accelerated molecular dynamics: Unconstrained enhanced sampling and free energy calculation. J. Chem. Theory Comput. 11, 3584–3595 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Lyubartsev A. P., Martsinovski A. A., Shevkunov S. V., Vorontsov-Velyaminov P. N., New approach to Monte Carlo calculation of the free energy: Method of expanded ensembles. J. Chem. Phys. 96, 1776–1783 (1992). [Google Scholar]
- 56.Ubbink M., The courtship of proteins: Understanding the encounter complex. FEBS Lett. 583, 1060–1066 (2009). [DOI] [PubMed] [Google Scholar]
- 57.Kinz-Thompson C. D., Gonzalez R. L. Jr, smFRET studies of the ‘encounter’ complexes and subsequent intermediate states that regulate the selectivity of ligand binding. FEBS Lett. 588, 3526–3538 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Patel R. Y., Doerksen R. J., Protein kinase-inhibitor database : Structural variability of and inhibitor interactions with the protein kinase P-Loop. J. Proteome. Res. 9, 4433–4442 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Shah N. P., et al. , Multiple BCR-ABL kinase domain mutations confer polyclonal resistance to the tyrosine kinase inhibitor imatinib (STI571) in chronic phase and blast crisis chronic myeloid leukemia. Cancer Cell 2, 117–125 (2002). [DOI] [PubMed] [Google Scholar]
- 60.Uitdehaag J. C. M., et al. , Target residence time-guided optimization on TTK kinase results in inhibitors with potent anti-proliferative activity. J. Mol. Biol. 429, 2211–2230 (2017). [DOI] [PubMed] [Google Scholar]
- 61.Cowan-Jacob S. W., et al. , The crystal structure of a c-Src complex in an active conformation suggests possible steps in c-Src activation. Structure 13, 861–871 (2005). [DOI] [PubMed] [Google Scholar]
- 62.Kobayashi C., et al. , GENESIS 1.1: A hybrid-parallel molecular dynamics simulator with enhanced sampling algorithms on multiple computational platforms. J. Comput. Chem. 38, 2193–2206 (2017). [DOI] [PubMed] [Google Scholar]
- 63.Jung J., et al. , GENESIS: A hybrid-parallel and multi-scale molecular dynamics simulator with enhanced sampling algorithms for biomolecular and cellular simulations. Wiley Interdiscip. Rev. Comput. Mol. Sci. 5, 310–323 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Lindorff-Larsen K., et al. , Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins 78, 1950–1958 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Hornak V., et al. , Comparison of multiple Amber force fields and development of improved protein backbone parameters. Proteins 65, 712–725 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Jorgensen W. L., Chandrasekhar J., Madura J. D., Impey R. W., Klein M. L., Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935 (1983). [Google Scholar]
- 67.Wang J., Wolf R. M., Caldwell J. W., Kollman P. A., Case D. A., Development and testing of a general amber force field. J. Comput. Chem. 25, 1157–1174 (2004). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.