Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Dec 14.
Published in final edited form as: Eur Phys J B. 2021 Oct 11;94(10):203. doi: 10.1140/epjb/s10051-021-00207-7

TYROSINE KINASES: COMPLEX MOLECULAR SYSTEMS CHALLENGING COMPUTATIONAL METHODOLOGIES

Trayder Thomas 1, Benoît Roux 1,*
PMCID: PMC9749240  NIHMSID: NIHMS1853822  PMID: 36524055

Abstract

Classical molecular dynamics (MD) simulations based on atomic models play an increasingly important role in a wide range of applications in physics, biology, and chemistry. Nonetheless, generating genuine knowledge about biological systems using MD simulations remains challenging. Protein tyrosine kinases are important cellular signaling enzymes that regulate cell growth, proliferation, metabolism, differentiation, and migration. Due to the large conformational changes and long timescales involved in their function, these kinases present particularly challenging problems to modern computational and theoretical frameworks aimed at elucidating the dynamics of complex biomolecular systems. Markov state models have achieved limited success in tackling the broader conformational ensemble and biased methods are often employed to examine specific long timescale events. Recent advances in machine learning continue to push the limitations of current methodologies and provide notable improvements when integrated with the existing frameworks. A broad perspective is drawn from a critical review of recent studies.

Introduction

Classical molecular dynamics (MD) simulations based on atomic models play an increasingly important role in a wide range of applications in physics, biology, and chemistry. The approach consists of constructing detailed atomic models of the macromolecular system and, having described the microscopic forces with a potential function, using Newton’s classical equation to literally “simulate” the dynamical motions of all the atoms as a function of time. The calculated trajectory, though an approximation to the real world, provides detailed information about the time course of the atomic motions, which is impossible to access experimentally. While great progress has been made, producing genuine knowledge about biological systems using MD simulations remains challenging. Among the most difficult problems are the characterization of large conformational changes occurring over long timescales that impact biological function.

In the last decade, a powerful theoretical and computational paradigm for studying complex biomolecular systems has emerged. Building on classical statistical mechanical techniques such as alchemical free energy perturbation (FEP)14 and umbrella sampling (US) in multiple dimensions,5,6 it consists in combining various strategies including, transition pathway determination711 and stochastic Markov modeling,1214 together with time-independent component analysis (TICA)15,16 and kinetic mapping.15 Theoretical frameworks relying on transition path theory (TPT),1720 and on a variational formulation of the transfer operator,2124 serve as foundations to expand many of these novel ideas toward various data-driven machine learning (ML) techniques.2430

Protein tyrosine kinases offer particularly challenging systems to advance and put to the test this emerging computational framework. After providing some background on the structure and function of kinases, we will broadly review previous computational studies of these important systems and highlight the main findings. We will conclude with an outlook on the considerable challenges that remain and the promise of the most recent theoretical advances in the treatment of biomolecular dynamical systems.

Some background on protein tyrosine kinases

Protein kinases are enzymes that catalyze the covalent transfer of the γ-phosphate of an adenosine triphosphate (ATP) molecule onto a tyrosine, serine, threonine, or histidine residue on targeted and specific substrate proteins (including kinases), thereby sending a “downstream” chemical signal throughout a network of proteins. Phosphorylation of a substrate protein can result in a functional change and modification of cellular localization, with a potential impact on numerous cellular processes. The human genome encodes more than 500 protein kinases, making the kinome one of the largest gene families.31 Kinases are important cellular signaling enzymes that regulate cell growth, proliferation, metabolism, differentiation, and migration. Unregulated kinase activity and loss of function are often involved in a wide range of human diseases. For this reason, these enzymes are major therapeutic targets, and the discovery of kinase-specific inhibitors has been intensely pursued by the pharmaceutical industry for many years.3237 Kinases are highly validated drug targets with a vast amount of available structural data.

The members of the Src family of protein tyrosine kinases represent prototypical model systems of great interest to study the regulation mechanisms. All nine members of the Src family (Src, Yes, Fyn, Lyn, Lck, Blk, Hck, Fgr, and Yrk) are highly homologous proteins with a similar regulatory mechanism sharing a multidomain architecture illustrated in Figure 1 that comprises a catalytic tyrosine kinase domain (grey), preceded by two peptide-binding regulatory modules, the Src-homology domains SH2 (red) and SH3 (yellow).38 The kinase members of the Src family are allosteric molecular switches. Their catalytic activity can be modulated in response to cellular signals. Phosphorylation of the protein kinase is especially crucial to the regulation of the catalytic activity.

Figure. 1.

Figure. 1.

c-Src tyrosine kinase in the autoinhibitory assembled form (PDB 2SRC). The regulatory domains SH3 (yellow) and SH2 (red) are shown along with the linker (purple) to the kinase domain (white). Important structural features of the kinase domain have also been highlighted: P-loop (blue), activation loop (green), αC-helix (orange), and c-tail (cyan). An analogue of ATP (grey sticks) is also shown bound to the kinase domain.

Molecular dynamics studies

Molecular dynamics (MD) simulations provide one important avenue to better understand protein kinases. Many computational studies focused on the conformational transition taking place within the catalytic domain, which is from an inactive to its catalytically competent active state.39,40 Characterizing the properties of the isolated catalytic domain without the SH2 and SH3 regulatory domains has also been of considerable interest,41 as it is known to be constitutively active in solution according to experiment.39 Smaller conformational transitions have also attracted interest, in particular, a 3-residue motif comprised of Asp-Phe-Gly (DFG) near the N-terminus of the activation (A-) loop covering the catalytic site is involved in the activation process and its conformation is known to be critical for binding some kinase inhibitors.4244 An understanding of how the conformational ensemble of kinases is influenced by ligand binding, phosphorylation, autoinhibition, and drug resistant mutations is of keen interest, and there is a long history of MD studies investigating the effects of these phenomena on the conformational landscape of kinases. In one of the earliest studies, Young et al. examined the effect of mutations on the structural fluctuations within the multi-domain c-Src kinase.45 Suenaga et al. observed the effect of phosphorylation on the structural flexibility of Shc kinase.46 Kuriyan and co-workers observed a Src-like inactive conformation in the Abl tyrosine kinase domain.47 Dixit et al. examined the activation mechanisms in the catalytic domain of Abl and EGFR kinases.48 In time, computational studies became ever more ambitious. Cambran et al. studied the effect of myristoylation, phosphorylation, and ligand binding on the conformational ensemble of protein kinase A (PKA),49,50 whilst Boras et al. conducted a similar study focusing on the autoinhibition by the regulatory domains.51 Lopez et al. used an array of methods to assess the impact of phosphorylation on the conformation of ERK2.52 In a study of the Janus kinase 2 (JAK2) tyrosine kinase,53 the DESRES team led by David Shaw used unbiased MD simulations to investigate autoinhibition. Similar characterizations have also been conducted on the conformation of the epidermal growth factor receptor (EGFR) and its changes due to dimerization and mutations.5459Yan et al. studied the hepatocyte growth factor receptor (HGFR, also known as c-Met),60 and found that the allosteric inhibitor tivantinib caused a transition from the DFG-in to DFG-out state in a sub-microsecond timescale allowing them to observe the DFG-flip during several 250 ns simulations.

Conformation of the DFG motif

As mentioned above, the orientation of the small DFG motif plays a role in the inhibition mechanism of the Abelson tyrosine kinase (Abl) by the anticancer drug imatinib.43 Imatinib displays a much lower inhibitory effect on c-Src, even though these two kinases display a high level of sequence identity (47%) and similar structural features.43,61,62 The similarity of binding site residues between the two kinases, combined with the idea that imatinib binds exclusively to a DFG-out conformation, led to the view that there was a “conformational selection” mechanism of binding specificity. But the crystal structure of imatinib in complex with the c-Src kinase domain subsequently showed that it also adopted the same inactive DFG-out conformation.61,63,64 This led to the view that there must be some “thermodynamic penalty” associated with the DFG-out conformation.65 Because of its great importance, the conformational transition of the DFG motif and its impact on the binding specificity of inhibitors has been the subject of many computational studies.2,60,62,6670 Researchers from DESRES characterized the conformation of Abl kinase,62 identifying multiple factors that contributed to the stability of the DFG-in or DFG-out poses, including a pH dependence in the conformation of the DFG motif that they attributed to electrostatic changes during the catalytic cycle. An analysis of the kinetics of imatinib binding in this same study reinforced that imatinib selectively bound to the DFG-out conformation and estimated the timescale of the DFG-flip to be in the order of 10s of milliseconds.

The DFG-flip conformational transition and its thermodynamic impact specifically regarding the selectivity of imatinib has also been the subject of several computational studies.24,48,62,66,6872 Lovera et al. calculated a free energy landscape of the DFG-flip transition in c-Abl and c-Src using meta-dynamics MD simulations with the AMBER force field,68 yielding free energy differences 4.0 and 6.0 kcal/mol in Abl and c-Src, respectively. Roux and co-workers estimated the same quantity using umbrella sampling with the CHARMM force field and obtained 1.4 and 5.4 kcal/mol in Abl and c-Src, respectively.2,69 Furthermore, computations based on alchemical free energy perturbation demonstrate that different hidden factors (e.g. DFG-flip, intrinsic affinity, steric clashes with the P-loop) that are not readily detectable by experiments partly cancel one another to explain the observed Kd of imatinib and of a related inhibitor G6G.4 Lovera et al. investigated the free energy landscape of several widespread drug-resistant mutations of Abl kinase, finding a category of mutations that reduced imatinib affinity by favoring the DFG-in conformations of Abl.67 All these studies indicate that the DFG-flip can happen for both Abl and c-Src, but the latter simply incurs a larger “thermodynamic penalty”. Such findings are not unique to Abl, and similar studies have been conducted across the kinome.73,74 Nonetheless, the issue is not settled, as an alternative model to explain these observations was recently proposed by Kern and co-workers, who found from their kinetic experiments that imatinib bound in a two-step process which differed significantly in kinetics between Abl and Src.75

Activation pathways and string method

Because the timescales for the conformational changes leading to activation and inactivation are very long, many of the early studies relied on some type of non-equilibrium perturbation to accelerate transitions of interest, such as targeted-MD driven by the root-mean-square deviations,45,7679 the half quadratic biasing for positive selection of spontaneous fluctuations,8082 or chemical perturbation (protonation of titratable residues).62 While these efforts provided some useful mechanistic insight, the path generated by these biased methods must be considered with caution because it generally depends on the choice of progress variable used to move the system between the initial and final states.83

While such driven (biased) simulations based on non-equilibrium trahectories can generate useful information to characterize conformational transition pathways, they are often insufficient to tackle slow conformational transitions occurring on long timescales. These issues are resolved with the more rigorous computational framework provided by the string method.7,8 The string method represents the pathway as a chain of M “images” (copies of the full system) in the multidimensional space of the set of collective variables Z, e.g., the pathway is given as{Z(1), Z(2), …, Z(M)}. In the string method with swarms-of-trajectories,9 the transition pathway is iteratively refined starting from an initial guess by using the mean drift of short unbiased simulations launched from each of the M images. Because all these simulations are uncorrelated, the procedure can be efficiently scaled up to a very large number of CPUs using the general multiple-copy algorithm framework implemented in NAMD.84 It is also possible to accelerate the convergence of the iterative refinement using a multi-scale pre-conditioner.85

The first application of the string method with swarms-of-trajectories was to determine the activation pathway in the isolated catalytic domain of Src kinase.86 A subsequent study reported a very ambitious characterization of the activation pathway of the catalytic domain in the presence of the SH2 and SH3 regulatory domains.87 Alternative approaches to the string method have also been used to study conformational transition in kinases, including the adaptively biased path optimization (ABPO) method to determine the principal curve defining a conformational transition between two known end states.10 Post and co-workers used ABPO to examine the transitions within the catalytic domain of c-Src and CDK2 kinase.8890 This analysis highlighted the importance of a previously identified switched electrostatic network81,82 with respect to the αC helix and the activation loop motions.8890 The string method was also used to determine the transition pathway of the DFG-flip.69 Another pathway algorithm developed by Levy and Elber, the called “Rock-Climbing” method, was to characterize the DFG-flip transition in Abl kinase.11,91

Umbrella sampling and free energy landscape

Umbrella sampling simulations were used to calculate the potential of mean force (PMF) along selected collective coordinates to examine the free energy landscape associated with various processes.10,69,77,87,9295 Studies included the conformational equilibrium of the N-terminus of the catalytic domain,77 the DFG-flip,69 and the complete transition from the inactive to the active state in the catalytic domain.10,87,92,93 In a number of cases, the transition pathway was exploited as a basis to map the free energy landscape for these conformational changes. In one of the earliest efforts, Yang et al. constructed a low resolution free energy landscape for the activation of Hck kinase from unbiased equilibrium MD simulations initiated along a TMD transition pathway, revealing the existence of intermediate conformations along the activation pathway.92 Sampling strategies relying on the curvilinear transition pathways also include the Milestoning method,96 with a recent application to the DFG-flip transition in Abl kinase.91 To enhance the sampling, many of these PMF calculations were carried out using a multiple-copies replica-exchange algorithm.84 The ABPO method was used compute the PMF along the curvilinear activation transition pathway of Src.90 Metadynamics is also a possible avenue that has been used to enhance the sampling capability of MD simulations in studies of kinases.54,68,97

Alchemical FEP

Alchemical free energy perturbation (FEP) MD simulations (FEP/MD) is a method that can be used to calculate the equilibrium binding free energy of ligands to their target receptor.98 The method is generally considered to be accurate to approximately 1 kcal/mol, and the cancellation of errors between similar binding sites has been shown to result in increased accuracy when predicting selectivity for kinases.99 Alchemical FEP/MD simulations were used to examine the binding free energy of kinase inhibitors.24 While these calculations are mainly focused on the binding affinity of a ligand to a given protein conformation, the critical issue of conformational diversity cannot be avoided, even if it does not appear directly in alchemical FEP/MD simulations. For example, the “in” or “out” conformation of the DFG motif is critical to ascertain the binding specificity of imatinib to Abl and c-Src kinases.2 The general take-home message of alchemical FEP/MD is that the method is useful, but the underlying issue of conformational diversity must be resolved, at least partially, in order to draw meaningful conclusions.

Coarse-grained models

In part to circumvent the nearly insurmountable sampling problems associated with a large, complex protein, a number of studies have resorted to a simpler coarse-grained (CG) representation of the kinase. Yang and Roux used a multistate coarse-grained Gõ-like model to simulate the activation of the catalytic domain of Hck kinase.100 A similar multistate Gõ-like model of Lyn kinase was exploited to search the optimal path for the same transition using the maximum flux transition path (MFTP) method.88 Simplified CG models have also been used to test algorithms to determine conformational pathways,10 and examine the computational feasibility of Markov State Models (MSMs).100 While it is clear that detailed all-atom models are more realistic, simplified CG models have played an important role to test and illustrate new algorithms in the context of kinase structures. This type of CG model also played a useful role in the interpretation of experimental data from small angle X-ray scattering (SAXS) of Hck kinase in solution.92,101

Markov State Models

Markov State Models (MSMs) provide a powerful computational strategy combining the information accumulated from a large number of MD simulations.12 Building an MSM involves splitting up a system’s conformational space into a set of microstates, running independent trajectories visiting each of those regions, and then tabulating the observed state-to-state transitions into a Markov matrix of transition probabilities.1214,102104 Traditionally, MSMs have been constructed from aggregate MD simulation data from multiple trajectories. With the tremendous progress in accelerating classical MD simulations on inexpensive graphics processing units (GPUs)105108 the simulation of tens of thousands of short, independent trajectories can be executed with relative ease using OpenMM105107 or AMBER-GPU.108 Sophisticated analysis software packages MSMBuilder (http://msmbuilder.org)13,109 and PyEMMA110 play an important role for a sustained progress in the field. The approach was used to characterize the activating transition in the catalytic domain of c-Src kinase.94 The qualitative conclusions confirmed the existence of a putative intermediate state along the activation pathway, as previously revealed.92 The MSMs thus constructed, enabled a Transition Path Theory (TPT)19 analysis of the nature of the “reaction tube” supporting the conformational change for the activation of c-Src.111 Sultan et al. used MSMs to analyze 1.7 ms of unbiased simulations of BTK kinase to construct a MSM that captured the DFG-flip.112 This methodology was later expanded with a further aggregate 4ms of MD data across 6 Src-family kinases that was combined into single MSM, establishing a common conformational ensemble.113

It is clear that the large number of conformations accessible to kinases poses a special challenge to efforts aimed at designing inhibitors binding to a given kinase with high specificity and affinity. Conventional docking algorithms can produce native like complex structures when a number of conditions are satisfied (e.g., if a deep binding pocket has a single dominant conformation).114 However, they often fall short when a protein is flexible and can adopt multiple conformations. Analysis based on a single non-dominant structure, whether it is from X-ray crystallography or from homology modeling, can lead to an erroneous SBDD strategy. For kinases structural elements near the binding pocket (A-loop, P-loop, DFG motif, catalytic spine, etc.) are able to adopt multiple conformations (see Figure 2). These structural elements affect the shape and properties of the binding pocket. As a consequence, it becomes highly difficult to ascertain which of those conformations should be considered in a drug discovery project. For example, the DFG motif can point inward (DFG-in) or outward (DFG-out), which strongly affects the type of inhibitors that can bind to the pocket; the A-loop and P-loop can adopt a variety of conformations, creating a large number of combinatorial possibilities. In an effort to try to consider multiple conformations, one can try to harvest the possible conformations from similar proteins in the PDB, or by computational sampling (such as Rosetta) to generate alternative structures. Nevertheless, such knowledge-based heuristic approaches are not guaranteed to discover all accessible conformations of the various structural elements for the protein of interest. More importantly, it does not provide information about the relative importance of multiple different conformations of the target protein (the frequency of conformations in the PDB is not a measure of energetic stability). In large part to explore this issue, the MSM framework was also used to examine the conformational diversity of Abl kinase in the absence of any ligand.115 The efforts showed that it is feasible to discover and identify most of the most relevant conformations of the catalytic domain of a kinase, although ranking them in terms of their equilibrium population is harder and more uncertain. While these states are likely to be metastable, a characterization of the kinetic timescales is extremely difficult.116 Even with a little over 100 μs of aggregate MD data, the eigenvalues of the transition matrix remain highly uncertain. Similar issues are also present when trying to characterize the binding of a specific ligand such as imatinib to Abl kinase.117

Figure. 2.

Figure. 2.

Four main structural elements in Abl kinase. (Left) Crystal structure of Abl kinase (PDB 2HYY) highlighting the P-loop (blue), DFG motif (purple), activation loop (green), αC-helix (orange), and the bound ligand imatinib (yellow). (Right) Alternative conformations of structural elements from three Abl kinase experimental structures (PDB 2HYY, 2G1T, 6XR6), illustrating the broad conformational space of the binding site.

Relating computational studies to functional assays

Relating the results of computational studies to experimental observation about kinases also presents some special challenges. The activity of signaling enzymes is often measured indirectly via cellular assays that reveal small but biologically meaningful trends on macroscopic timescales.45 For this reason, validating the results of computational studies with respect to kinase function is not always straightforward. Generally, it is clear that microscopic factors that stabilize the down-regulated inactive conformation of the kinase should reduce activity.40 Correspondingly, factors that destabilize the inactive state should appear as increased kinase activity in experimental assays. But this naïve understanding rapidly encounters difficulties. For example, the isolated catalytic domain (without its regulatory domains) is known to be constitutively active according to experiments.39 Presumably, the active catalytic domain is trans-phosphorylated via a bimolecular encounter with another kinase in order to stabilize the fully active state. Indeed, the calculated free energy landscape shows that phosphorylation of Y416 in the A-loop essentially “locks” the kinase into its catalytically competent conformation.93 Nonetheless, while this conceptual framework is reasonable, it is unclear whether it can be reconciled with experimental observations: the timescale of the interconversion between the inactive and active state from the MSM is on the order of 100 μs whereas the experimentally observed timescale for Src-family tyrosine kinase autophosphorylation at residue Tyr416 is on the order of minutes. It was possible to relate these vastly disparate timescales by constructing a simple kinetic model using the data extracted from atomistic simulations as well as a reasonable estimate for the bimolecular rate of kinase phosphorylation.94 A similar conceptual framework was used to explain the functional behavior of the activating W260A mutant.118 At first glance, the computational result seems to be inconsistent with experiments: while the W260A mutation in c-Src is markedly up-regulated,119 the free energy cost to reach the active-like unphosphorylated state A is only slightly smaller in the W260A mutant compared with WT.118 Nonetheless, such a small difference in stability accounts for the large increase in the activity of the mutant that is observed experimentally.118 Furthermore, umbrella sampling computations indicate that the configuration of the SH2 and SH3 regulatory domains, greatly affects the ability of the catalytic domain to adopt the active or inactive conformation.87 Conversely, individual inhibitors binding to the catalytic domain can have differing allosteric effects on whether the regulatory domains adopt a closed or open conformation.120 Computational methods based on simplified CG models have played a useful role to interpret SAXS data on Hck kinase in solution in terms of the configuration of the multi-domain kinase.92,101 For instance, MSM one can observe the organization of the SH2, SH3, and catalytic domain in response to different external factors,92 or mutations.101 This type of strategy has also been useful to interpret data about the membrane anchoring tail of Hck kinase from neutron reflectometry.121 Xie et al. used NMR to identify conformational states occupied by Abl kinase and were then able to fit their chemical exchange data to a model in order to predict populations and kinetics of those states for the wildtype and various mutants.122 MD simulations have also been important to develop a perspective on substrate recognition integrating information from NMR.123

Binding of inhibitors

There is also considerable interest in using computational methods to understand experimental data about the binding of specific kinase inhibitors. One of the most studied prototypical systems in this case is the binding of imatinib and how it is affected by the conformation of the DFG motif.43,75,124126 The tryptophan fluorescence quenching experiments of Agafonov et al.75 revealed that the ligand imatinib (L) binds to the Abl kinase protein (P) via a two-step process: P + LP ~ LPL. While the first step is a concentration-dependent bimolecular association leading to the formation of a protein-ligand intermediate, the second step corresponds to a slow concentration-independent unimolecular rearrangement that ultimately leads this intermediate to the strongly bound protein-ligand complex. The authors interpreted this as evidence for an induced-fit conformational change of the protein kinase that is directly responsible for the binding specificity. However, an extensive MSM study indicates that there exists a large number of metastable states corresponding to non-specific association of imatinib with Abl kinase along two main pathways leading imatinib to the binding pocket.117 The MSM analysis based on more than 500 μs of aggregate MD data strongly supports the notion that the interconversion from the metastable binding modes of the ligand to the fully bound state is extremely slow, suggesting that the set of metastable binding modes of imatinib could reasonably give rise to the long-lived intermediate state revealed by the tryptophan fluorescence data.75 In that sense, the picture emerging from the MSM analysis is consistent with the actual experimental observation of a slow concentration-independent unimolecular rearrangement leading to the final complex. This picture departs from the traditional view of the induced-fit mechanism, where an initial weakly bound protein-ligand complex induces some conformational changes in the protein that leads to a strongly bound protein-ligand complex, while the computational results rather point to the slow rearrangement of the ligand toward its correct binding pocket from a multitude of long-lived metastable states that underlies the slow step detected in the fluorescence experiment. More generally, kinases are viewed as highly flexible allosteric proteins with multiple moving parts that can adopt different conformations, which clearly impacts on our ability to design small molecule inhibitors binding with high specificity. In Figure 2, four structural elements in Abl kinase displaying multiple conformational variations—the P-loop (blue), the DFG motif (purple), the A-loop (green), and the αC-helix (orange)—are illustrated. Those conformational variants affect the binding pocket. While a large amount of attention has been devoted to the features of the isolated catalytic domain by itself, understanding the influence of the regulatory SH2 and SH3 domains is tremendously important for the biological cellular signaling function,87,101,127 and membrane co-localization.121

Overall assessment of computational methods

At the present time, Markov state models (MSM) perhaps represent the richest and most powerful computational framework to characterize the dynamics of complex biomolecular systems.1214 Yet, in spite of a huge amount of aggregate MD data, our own efforts with kinase proteins were only partial successes.94,115117 One of the fundamental difficulties and an important requirement for the construction of a reliable and accurate MSM is the definition of “states”, dynamically meaningful regions of conformational space that must correctly map out the behavior of the system. Several structural elements of kinases display great flexibility. Ultimately, it is important to realize that MSMs only provide a model of the explored conformational space. There are several slow processes that may not be well sampled, even with a millisecond of data. In our MSM analysis, dimensionality reduction of the full configurational space was carried out via the time-lagged independent component analysis (TICA),15,16 applied to pre-identified structural regions displaying high conformational variability, e.g., the A-loop, the αC-helix, the DFG motif, and the P-loop.115,116 The ideas of Taylor and co-workers to subdivide a kinase structure into contiguous “communities” that exhibit internally correlated motions provide some interesting directions worth exploring to improve the identification of relevant structural protein regions and select appropriate “features” for the MSM analysis.128 These limitations notwithstanding, the MSM framework was undoubtedly invaluable in exploring the conformational landscape of apo-Abl kinase,115 allowing us to identify and rank the metastability of novel states. Further analysis of these states by comparing them to kinase crystal structures available in the PDB116 revealed high similarity between our identified apo-Abl conformations and ligand-bound structures across other kinase families. This analysis provided remarkable insight into both the similarities in conformational behavior across kinase families, and the druggability of our identified conformations. Nevertheless, despite the 800 μs of MD simulation used in these studies, we still fell short of reproducing the correct kinetics or thermodynamics of the molecular system. MSMs are prone to discretization error arising during the clustering steps that causes underestimation of energy, and error due to under-sampling of rare transitions that can cause overestimation of the same barriers. Shütte and collaborators have developed approaches for estimating the discretization error in MSM and address this issue.129 130,131

Understanding the binding of specific kinase inhibitors is likely to remain an important objective for many years to come. In some cases, when the binding process is fairly fast, it has been possible to construct an accurate MSM from the aggregate data of unbiased MD simulations.132,133 For example, to study the binding of trypsin and its competitive inhibitor benzamidine, Buch et al.132 used 495 unbiased MD of 100 ns each for a total of 49 μs. For slower binding processes where slow induced-fit conformational changes occur concertedly during the binding process, this direct strategy would not be practical. It is unlikely that brute-force MD could simulate the complete unbinding (koff) of a slow kinase inhibitor such as imatinib, which takes place over ~40 ms. From a more mundane point of view, improving the accuracy of the atomic force field will also be needed. For reliable computational studies, the force field for the ligand inhibitors must be obtained in an objective manner, using a protocol with no direct manipulation by the users, e.g. GAFF,134 CGENFF,135 or built from the General Automated Atomic Model Parameterization (GAAMP) protocol.136 Another new frontier to tackle is the binding of covalent inhibitors, which are the focus of increased interest following the recent clinical success of drugs such as ibrutinib.137 Properly optimized covalent inhibition offers potential for longer on target residency time and selectivity. The design of accurate computational treatments is very challenging because the action of covalent inhibitors involves a variety of microscopic processes that take place on timescales that are far beyond the reach of standard MD simulations. Furthermore, because the formation of a covalent bond is a quantum mechanical (QM) process it is necessary to go beyond classical force fields. While simulating the formation of the covalent bond requires a quantum mechanical/molecular mechanical (QM/MM) representation of the system, the rate for the process itself can naturally be incorporated into the MSM framework. In fact, MSM is arguably the only framework by which the covalent bond formation can be incorporated into the kinetics of the slower diffusion processes.

Future Outlook

Many of the theoretical frameworks used to extract kinetics from MD simulations find their underpinning in transition path theory (TPT),1720 which considers a reaction across a rugged energy landscape via the pathways of highest flux of probability between two states. This method has been expanded upon through a variational approach to molecular kinetics,2124 which has been more suitable to machine learning approaches and led to methods such as VAMPnets.26 Klus et al. used an alternative kernel-based approach and achieved a competitive result.138 Boltzmann generators are a different strategy that bypasses the required sampling of rare events by drawing statistically independent samples from a prior distribution and reversibly mapping them to a “latent space” where pathways can be linearly interpolated between minima and then mapped back into configuration space with an associated probability.29 The length of most simulations, which is orders of magnitudes smaller than that of the processes of interest, is likely to remain a key limitation. Previous applications show that efficient sampling of rare events remain the greatest challenge in extracting useful kinetic information from MD simulations; these events are the most relevant to the kinetics but also the most likely to be undersampled. The information from unbiased and biased MD trajectories may be combined to improve the statistical robustness and convergence of the MSM, either using the dynamical weighted histogram analysis (DHAM),139 or the discrete transition-based reweighting analysis method (dTRAM).140,141

Some promising advancements in this area may be provided by Machine learning (ML). As an example, Wang et al. approached the sampling problem by utilizing machine learning to design a coarse-grain force field for their system.142 The recent success of ML techniques suggests that they may be capable of approximating high-dimensional functions with controllably small errors.27,28,30,143146 They may assist in solving the sampling problem by combining these approaches with some aspects of importance sampling methods for data acquisition to focus the computational resources on the critical part of a problem. One approach has been to leverage information from the string method to construct models focused on individual pathways.96,147

The problem of accurately defining a state space for the MSM also has promising solutions in machine learning. VAMPnets26 combine the optimization of hyperparameters with the featurization, dimensionality reduction, and clustering of the system into a single machine learning pipeline, thereby eliminating many sources of error. We have applied this methodology to the imatinib-Abl system,117 and demonstrated its ability to identify metastable states of the protein, although it does not solve undersampling problems. Ung et al. used machine learning to classify the conformation of all kinase structures in the PDB and a similar approach could be applied to simulation data.148

While these methods, particularly in the area of ML, represent a significant advancement in the amounts of simulation data we can effectively utilize, and the accuracy of the analysis we can perform, they are not yet closing the gap between the timescales of what we can access, and the events we are interested in. This is largely because the methods endeavor to be unbiased and undirected, if we wish to close the gap faster than waiting for the gradual improvement of hardware, we will need methods that are more integrated with data generation. Fortunately, there is every indication that this is an active field with a huge potential for innovation.

Acknowledgment

This work is supported by the National Cancer Institute (NCI) of the National Institutes of Health (NIH) through grant R01-CAO93577.

References

  • 1.Deng Y & Roux B. Computations of Standard Binding Free Energies with Molecular Dynamics Simulations, The journal of physical chemistry 113, 2234–2246, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Lin YL, 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]
  • 3.Lin YL & Roux B. Computational Analysis of the Binding Specificity of Gleevec to Abl, c-Kit, Lck, and c-Src Tyrosine Kinases, J. Am. Chem. Soc 135, 14741–14753, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Lin YL, Meng Y, Huang L & Roux B. Computational Study of Gleevec and G6G Reveals Molecular Determinants of Kinase Inhibitor Selectivity, J. Am. Chem. Soc 136, 14753–14762, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Jiang W, Luo Y, Maragliano L & Roux B. Calculation of Free Energy Landscape in Multi-Dimensions with Hamiltonian-Exchange Umbrella Sampling on Petascale Supercomputer, J. Chem. Theo. Comp 8, 4672–4680, (2012). [DOI] [PubMed] [Google Scholar]
  • 6.Wojtas-Niziurski W, Meng Y, Roux B & Berneche S. Self-Learning Adaptive Umbrella Sampling Method for the Determination of Free Energy Landscapes in Multiple Dimensions, J. Chem. Theo. Comp 9, 1885–1895, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Weinan E, Ren W & Eijnden E. String method for the study of rare events, Physical Review B 66, 052301, (2002). [Google Scholar]
  • 8.Maragliano L, Fischer A, Vanden-Eijnden E & Ciccotti G. String method in collective variables: minimum free energy paths and isocommittor surfaces, The Journal of chemical physics 125, 24106, (2006). [DOI] [PubMed] [Google Scholar]
  • 9.Pan AC, Sezer D & Roux B. Finding transition pathways using the string method with swarms of trajectories, The journal of physical chemistry 112, 3432–3440, (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Dickson BM, Huang H & Post CB. Unrestrained Computation of Free Energy along a Path, J. Phys. Chem. B 116, 11046–11055, (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Templeton C, Chen SH, Fathizadeh A & Elber R. Rock climbing: A local-global algorithm to compute minimum energy and minimum free energy pathways, J. Chem. Phys 147, 152718, (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Bowman GR, Pande VS & Noé F. An Introduction to Markov State Models and Their Application to Long Timescale Molecular Simulation Advances in Experimental Medicine and Biology 797, (2014). [Google Scholar]
  • 13.Pande VS, Beauchamp K & Bowman GR. Everything you wanted to know about Markov State Models but were afraid to ask, Methods 52, 99–105, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Prinz JH, Wu H, Sarich M, Keller B, Senne M, Held M, Chodera JD, Schutte C & Noe F. Markov models of molecular kinetics: generation and validation, The Journal of chemical physics 134, 174105, (2011). [DOI] [PubMed] [Google Scholar]
  • 15.Noe F & Clementi C. Kinetic Distance and Kinetic Maps from Molecular Dynamics Simulation, J. Chem. Theory Comput 11, 5002–5011, (2015). [DOI] [PubMed] [Google Scholar]
  • 16.Perez-Hernandez G, Paul F, Giorgino T, De Fabritiis G & Noe F. Identification of slow molecular order parameters for Markov model construction, J. Chem. Phys 139, 015102, (2013). [DOI] [PubMed] [Google Scholar]
  • 17.Metzner P, Schutte C & Vanden-Eijnden E. Illustration of transition path theory on a collection of simple examples, Journal of Chemical Physics 125, (2006). [DOI] [PubMed] [Google Scholar]
  • 18.Metzner P, Schutte C & Vanden-Eijnden E. Transition Path Theory for Markov Jump Processes, Multiscale Model Sim 7, 1192–1219, (2009). [Google Scholar]
  • 19.E WN & Vanden-Eijnden E. Transition-Path Theory and Path-Finding Algorithms for the Study of Rare Events, Annual Review of Physical Chemistry, Vol 61 61, 391–420, (2010). [DOI] [PubMed] [Google Scholar]
  • 20.Vanden-Eijnden E. Transition Path Theory, Adv Exp Med Biol 797, 91–100, (2014). [DOI] [PubMed] [Google Scholar]
  • 21.Nuske F, Keller BG, Perez-Hernandez G, Mey AS & Noe F. Variational Approach to Molecular Kinetics, Journal of chemical theory and computation 10, 1739–1752, (2014). [DOI] [PubMed] [Google Scholar]
  • 22.Thiede EH, Giannakis D, Dinner AR & Weare J. Galerkin approximation of dynamical quantities using trajectory data, The Journal of chemical physics 150, 244111, (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Lorpaiboon C, Thiede EH, Webber RJ, Weare J & Dinner AR. Integrated Variational Approach to Conformational Dynamics: A Robust Strategy for Identifying Eigenfunctions of Dynamical Operators, The journal of physical chemistry. B 124, 9354–9364, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Bittracher A, Banisch R & Schutte C. Data-driven computation of molecular reaction coordinates, Journal of Chemical Physics 149, (2018). [DOI] [PubMed] [Google Scholar]
  • 25.Conrad ND, Weber M & Schutte C. Finding Dominant Structures of Nonreversible Markov Processes, Multiscale Model Sim 14, 1319–1340, (2016). [Google Scholar]
  • 26.Mardt A, Pasquali L, Wu H & Noe F. VAMPnets for deep learning of molecular kinetics, Nature communications 9, 5, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Noe F, De Fabritiis G & Clementi C. Machine learning for protein folding and dynamics, Curr Opin Struct Biol 60, 77–84, (2020). [DOI] [PubMed] [Google Scholar]
  • 28.Noe F, Tkatchenko A, Muller KR & Clementi C. Machine Learning for Molecular Simulation, Annu Rev Phys Chem 71, 361–390, (2020). [DOI] [PubMed] [Google Scholar]
  • 29.Noe F, Olsson S, Kohler J & Wu H. Boltzmann generators: Sampling equilibrium states of many-body systems with deep learning, Science 365, (2019). [DOI] [PubMed] [Google Scholar]
  • 30.Bittracher A & Schutte C. A probabilistic algorithm for aggregating vastly undersampled large Markov chains, Physica D 416, (2021). [Google Scholar]
  • 31.Manning G, Whyte DB, Martinez R, Hunter T & Sudarsanam S. The protein kinase complement of the human genome, Science 298, 1912–1934, (2002). [DOI] [PubMed] [Google Scholar]
  • 32.Fabbro D & Garcia-Echeverria C. Targeting protein kinases in cancer therapy, Curr Opin Drug Discov Devel 5, 701–712, (2002). [PubMed] [Google Scholar]
  • 33.Cohen P. Protein kinases--the major drug targets of the twenty-first century?, Nat. Rev. Drug Discov 1, 309–315, (2002). [DOI] [PubMed] [Google Scholar]
  • 34.Noble MEM, Endicott JA & Johnson LN. Protein kinase inhibitors: Insights into drug design from structure, Science 303, 1800–1805, (2004). [DOI] [PubMed] [Google Scholar]
  • 35.Zhang JM, Yang PL & Gray NS. Targeting cancer with small molecule kinase inhibitors, Nat. Rev. Cancer 9, 28–39, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Zhang SY & Yu DH. Targeting Src family kinases in anti-cancer therapies: turning promise into triumph, Trends Pharmacol. Sci 33, 122–128, (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Ferguson FM & Gray NS. Kinase inhibitors: the road ahead, Nat Rev Drug Discov 17, 353–377, (2018). [DOI] [PubMed] [Google Scholar]
  • 38.Boggon TJ & Eck MJ. Structure and regulation of Src family kinases, Oncogene 23, 7918–7927, (2004). [DOI] [PubMed] [Google Scholar]
  • 39.Yamaguchi H & Hendrickson WA. Structural basis for activation of human lymphocyte kinase Lck upon tyrosine phosphorylation, Nature 384, 484–489, (1996). [DOI] [PubMed] [Google Scholar]
  • 40.Sicheri F, Moarefi I & Kuriyan J. Crystal structure of the Src family tyrosine kinase Hck, Nature 385, 602–609, (1997). [DOI] [PubMed] [Google Scholar]
  • 41.Kornev AP, Haste NM, Taylor SS & Ten Eyck LF. Surface comparison of active and inactive protein kinases identifies a conserved activation mechanism, Proc. Natl. Acad. Sci. U.S.A 103, 17783–17788, (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Nagar B, Bornmann WG, Pellicena P, Schindler T, Veach DR, Miller WT, Clarkson B & Kuriyan J. Crystal structures of the kinase domain of c-Abl in complex with the small molecule inhibitors PD173955 and imatinib (STI-571), Cancer Res. 62, 4236–4243, (2002). [PubMed] [Google Scholar]
  • 43.Schindler T, Bornmann W, Pellicena P, Miller WT, Clarkson B & Kuriyan J. Structural mechanism for STI-571 inhibition of Abelson tyrosine kinase, Science 289, 1938–1942, (2000). [DOI] [PubMed] [Google Scholar]
  • 44.Vajpai N, Strauss A, Fendrich G, Cowan-Jacob SW, Manley PW, Grzesiek S & Jahnke W. Solution conformations and dynamics of ABL kinase-inhibitor complexes determined by NMR substantiate the different binding modes of imatinib/nilotinib and dasatinib, J. Biol. Chem 283, 18292–18302, (2008). [DOI] [PubMed] [Google Scholar]
  • 45.Young MA, Gonfloni S, Superti-Furga G, Roux B & Kuriyan J. Dynamic coupling between the SH2 and SH3 domains of c-Src and hck underlies their inactivation by C-terminal tyrosine phosphorylation, Cell 105, 115–126, (2001). [DOI] [PubMed] [Google Scholar]
  • 46.Suenaga A, Kiyatkin AB, Hatakeyama M, Futatsugi N, Okimoto N, Hirano Y, Narumi T, Kawai A, Susukita R, Koishi T, Furusawa H, Yasuoka K, Takada N, Ohno Y, Taiji M, Ebisuzaki T, Hoek JB, Konagaya A & Kholodenko BN. Tyr-317 phosphorylation increases Shc structural rigidity and reduces coupling of domain motions remote from the phosphorylation site as revealed by molecular dynamics simulations, J Biol Chem 279, 4657–4662, (2004). [DOI] [PubMed] [Google Scholar]
  • 47.Levinson NM, Kuchment O, Shen K, Young MA, Koldobskiy M, Karplus M, Cole PA & Kuriyan J. A Src-like inactive conformation in the Abl tyrosine kinase domain, PLoS Biology 4, 753–767, (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Dixit A & Verkhivker GM. Hierarchical Modeling of Activation Mechanisms in the ABL and EGFR Kinase Domains: Thermodynamic and Mechanistic Catalysts of Kinase Activation by Cancer Mutations, PLoS Comput. Biol 5, e1000487, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Cembran A, Masterson LR, McClendon CL, Taylor SS, Gao JL & Veglia G. Conformational Equilibrium of N-Myristoylated cAMP-Dependent Protein Kinase A by Molecular Dynamics Simulations, Biochemistry 51, 10186–10196, (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Masterson LR, Cembran A, Shi L & Veglia G. in Adv. Protein Chem. Struct. Biol Vol. 87 (eds Christov Christo & Karabencheva-Christova Tatyana) Ch. 12, 363–389 (Academic Press, 2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Boras BW, Kornev A, Taylor SS & McCulloch AD. Using Markov State Models to Develop a Mechanistic Understanding of Protein Kinase A Regulatory Subunit RI alpha Activation in Response to cAMP Binding, J. Biol. Chem 289, 30040–30051, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Lopez ED, Burastero O, Arcon JP, Defelipe LA, Ahn NG, Marti MA & Turjanski AG. Kinase Activation by Small Conformational Changes, J Chem Inf Model 60, 821–832, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Shan YB, Gnanasambandan K, Ungureanu D, Kim ET, Hammaren H, Yamashita K, Silvennoinen O, Shaw DE & Hubbard SR. Molecular basis for pseudokinase-dependent autoinhibition of JAK2 tyrosine kinase, Nat Struct Mol Biol 21, 579–584, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Sutto L & Gervasio FL. Effects of oncogenic mutations on the conformational free-energy landscape of EGFR kinase, Proc. Natl. Acad. Sci. U. S. A 110, 10616–10621, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Shan YB, Eastwood MP, Zhang XW, Kim ET, Arkhipov A, Dror RO, Jumper J, Kuriyan J & Shaw DE. Oncogenic Mutations Counteract Intrinsic Disorder in the EGFR Kinase and Promote Receptor Dimerization, Cell 149, 860–870, (2012). [DOI] [PubMed] [Google Scholar]
  • 56.Arkhipov A, Shan YB, Das R, Endres NF, Eastwood MP, Wemmer DE, Kuriyan J & Shaw DE. Architecture and Membrane Interactions of the EGF Receptor, Cell 152, 557–569, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Arkhipov A, Shan YB, Kim ET, Dror RO & Shaw DE. Her2 activation mechanism reflects evolutionary preservation of asymmetric ectodomain dimers in the human EGFR family, Elife 2, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Endres NF, Das R, Smith AW, Arkhipov A, Kovacs E, Huang YJ, Pelton JG, Shan YB, Shaw DE, Wemmer DE, Groves JT & Kuriyan J. Conformational Coupling across the Plasma Membrane in Activation of the EGF Receptor, Cell 152, 543–556, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Shan YB, Arkhipov A, Kim ET, Pan AC & Shaw DE. Transitions to catalytically inactive conformations in EGFR kinase, Proc. Natl. Acad. Sci. U.S.A 110, 7270–7275, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Yan M, Wang H, Wang Q, Zhang Z & Zhang C. Allosteric inhibition of c-Met kinase in sub-microsecond molecular dynamics simulations induced by its inhibitor, tivantinib, Phys Chem Chem Phys 18, 10367–10374, (2016). [DOI] [PubMed] [Google Scholar]
  • 61.Seeliger MA, Ranjitkar P, Kasap C, Shan YB, Shaw DE, Shah NP, Kuriyan J & Maly DJ. Equally Potent Inhibition of c-Src and Abl by Compounds that Recognize Inactive Kinase Conformations, Cancer Res. 69, 2384–2392, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Shan YB, Seeliger MA, Eastwood MP, Frank F, Xu HF, Jensen MO, Dror RO, Kuriyan J & Shaw DE. 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]
  • 63.Dar AC, Lopez MS & Shokat KM. Small Molecule Recognition of c-Src via the lmatinib-Binding Conformation, Chem. Biol 15, 1015–1022, (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Seeliger MA, Nagar B, Frank F, Cao X, Henderson MN & Kuriyan J. 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]
  • 65.Cowan-Jacob SW, Mobitz H & Fabbro D. Structural biology contributions to tyrosine kinase drug discovery, Curr. Opin. Cell Biol 21, 280–287, (2009). [DOI] [PubMed] [Google Scholar]
  • 66.Aleksandrov A & Simonson T. Molecular Dynamics Simulations Show That Conformational Selection Governs the Binding Preferences of Imatinib for Several Tyrosine Kinases, J. Biol. Chem 285, 13807–13815, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Lovera S, Morando M, Pucheta-Martinez E, Martinez-Torrecuadrada JL, Saladino G & Gervasio FL. Towards a Molecular Understanding of the Link between Imatinib Resistance and Kinase Conformational Dynamics, PLoS Comput. Biol 11, e1004578, (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Lovera S, Sutto L, Boubeva R, Scapozza L, Dolker N & Gervasio FL. The Different Flexibility of c-Src and c-Abl Kinases Regulates the Accessibility of a Druggable Inactive Conformation, J. Am. Chem. Soc 134, 2496–2499, (2012). [DOI] [PubMed] [Google Scholar]
  • 69.Meng Y, Lin YL & Roux B. Computational study of the “DFG-flip” conformational transition in c-Abl and c-Src tyrosine kinases, J. Phys. Chem. B 119, 1443–1456, (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Vashisth H, Maragliano L & Abrams CF. “DFG-Flip” in the Insulin Receptor Kinase Is Facilitated by a Helical Intermediate State of the Activation Loop, Biophys. J 102, 1979–1987, (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Filomia F, De Rienzo F & Menziani MC. Insights into MAPK p38 alpha DFG flip mechanism by accelerated molecular dynamics, Bioorgan. Med. Chem 18, 6805–6812, (2010). [DOI] [PubMed] [Google Scholar]
  • 72.Dixit A & Verkhivker GM. Computational Modeling of Allosteric Communication Reveals Organizing Principles of Mutation-Induced Signaling in ABL and EGFR Kinases, PLoS Comput. Biol 7, e1002179, (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Vijayan RSK, He P, Modi V, Duong-Ly KC, Ma HC, Peterson JR, Dunbrack RL & Levy RM. Conformational Analysis of the DFG-Out Kinase Motif and Biochemical Profiling of Structurally Validated Type II Inhibitors, J Med Chem 58, 466–479, (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Haldane A, Flynn WF, He P, Vijayan RSK & Levy RM. Structural propensities of kinase family proteins from a Potts model of residue co-variation, Protein Sci. 25, 1378–1384, (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Agafonov RV, 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]
  • 76.Mendieta J & Gago F. In silico activation of Src tyrosine kinase reveals the molecular basis for intramolecular autophosphorylation, J Mol Graph Model 23, 189–198, (2004). [DOI] [PubMed] [Google Scholar]
  • 77.Banavali NK & Roux B. The N-terminal end of the catalytic domain of Src kinase Hck is a conformational switch implicated in long-range allosteric regulation, Structure 13, 1715–1723, (2005). [DOI] [PubMed] [Google Scholar]
  • 78.Banavali NK & Roux B. Anatomy of a structural pathway for activation of the catalytic domain of Src kinase Hck, Proteins-Structure Function and Bioinformatics 67, 1096–1112, (2007). [DOI] [PubMed] [Google Scholar]
  • 79.Banavali NK & Roux B. Flexibility and charge asymmetry in the activation loop of Src tyrosine kinases, Proteins 74, 378–389, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Paci E & Karplus M. Forced unfolding of fibronectin type 3 modules: an analysis by biased molecular dynamics simulations, Journal of molecular biology 288, 441–459, (1999). [DOI] [PubMed] [Google Scholar]
  • 81.Ozkirimli E & Post CB. Src kinase activation: A switched electrostatic network, Protein Sci. 15, 1051–1062, (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Ozkirimli E, Yadav SS, Miller WT & Post CB. An electrostatic network and long-range regulation of Src kinases, Protein Sci. 17, 1871–1880, (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Huang X, Yao Y, Bowman GR, Sun J, Guibas LJ, Carlsson G & Pande VS. Constructing multi-resolution Markov State Models (MSMs) to elucidate RNA hairpin folding mechanisms, Pac Symp Biocomput, 228–239, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Jiang W, Phillips JC, Huang L, Fajer M, Meng Y, Gumbart JC, Luo Y, Schulten K & Roux B. Generalized Scalable Multiple Copy Algorithms for Molecular Dynamics Simulations in NAMD, Comp. Phys. Comm 185, 908–916, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Tempkin JO, Qi B, Saunders MG, Roux B, Dinner AR & Weare J. Using multiscale preconditioning to accelerate the convergence of iterative molecular calculations, J. Chem. Phys 140, 184114, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Gan W, Yang S & Roux B. Atomistic view of the conformational activation of Src kinase using the string method with swarms-of-trajectories, Biophys J 97, L8–L10, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Fajer M, Meng Y & Roux B. The Activation of c-Src Tyrosine Kinase: Conformational Transition Pathway and Free Energy Landscape, J. Phys. Chem. B 121, 3352–3363, (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Huang H, Zhao RJ, Dickson BM, Skeel RD & Post CB. alpha C Helix as a Switch in the Conformational Transition of Src/CDK-like Kinase Domains, J. Phys. Chem. B 116, 4465–4475, (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Wu H & Post CB. Protein Conformational Transitions from All-Atom Adaptively Biased Path Optimization, Journal of chemical theory and computation 14, 5372–5382, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Wu H, Huang H & Post CB. All-atom adaptively biased path optimization of Src kinase conformational inactivation: Switched electrostatic network in the concerted motion of alphaC helix and the activation loop, J Chem Phys 153, 175101, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Narayan B, Fathizadeh A, Templeton C, He P, Arasteh S, Elber R, Buchete NV & Levy RM. The transition between active and inactive conformations of Abl kinase studied by rock climbing and Milestoning, Biochim Biophys Acta Gen Subj 1864, 129508, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Yang S, Banavali NK & Roux B. Mapping the conformational transition in Src activation by cumulating the information from multiple molecular dynamics trajectories, Proc. Natl. Acad. Sci. U.S.A 106, 3776–3781, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Meng Y & Roux B. Locking the Active Conformation of c-Src Kinase through the Phosphorylation of the Activation Loop, J. Mol. Biol 426, 423–435, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Shukla D, Meng Y, Roux B & Pande VS. Activation pathway of Src kinase reveals intermediate states as targets for drug design, Nat Commun 5, 3397, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Meng Y, Ahuja LG, Kornev AP, Taylor SS & Roux B. A Catalytically Disabled Double Mutant of Src Tyrosine Kinase Can Be Stabilized into an Active-Like Conformation, J. Mol. Biol 430, 881–889, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Maragliano L, Vanden-Eijnden E & Roux B. Free Energy and Kinetics of Conformational Transitions from Voronoi Tessellated Milestoning with Restraining Potentials., J. Chem. Theo. Comp 5, 2594–2589, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Morando MA, Saladino G, D’Amelio N, Pucheta-Martinez E, Lovera S, Lelli M, Lopez-Mendez B, Marenchino M, Campos-Olivas R & Gervasio FL. Conformational Selection and Induced Fit Mechanisms in the Binding of an Anticancer Drug to the c-Src Kinase, Sci Rep 6, 24439, (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Deng Y & Roux B. Computation of binding free energy with molecular dynamics and grand canonical Monte Carlo simulations, The Journal of chemical physics 128, 115103, (2008). [DOI] [PubMed] [Google Scholar]
  • 99.Albanese SK, Chodera JD, Volkamer A, Keng S, Abel R & Wang L. Is Structure-Based Drug Design Ready for Selectivity Optimization?, J Chem Inf Model 60, 6211–6227, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Yang S & Roux B. Src kinase conformational activation: thermodynamics, pathways, and mechanisms, PLoS computational biology 4, e1000047, (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Huang L, Wright M, Yang S, Blachowicz L, Makowski L & Roux B. Glycine substitution in SH3-SH2 connector of Hck tyrosine kinase causes population shift from assembled to disassembled state, Biochim. Biophys. Act. Gen. Sub 1864, 129604, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Bowman GR, Beauchamp KA, Boxer G & Pande VS. Progress and challenges in the automated construction of Markov state models for full protein systems, The Journal of chemical physics 131, 124101, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Bowman GR, Huang X & Pande VS. Using generalized ensemble simulations and Markov state models to identify conformational states, Methods 49, 197–201, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Bowman GR, Ensign DL & Pande VS. Enhanced modeling via network theory: Adaptive sampling of Markov state models, J Chem Theory Comput 6, 787–794, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Friedrichs MS, Eastman P, Vaidyanathan V, Houston M, Legrand S, Beberg AL, Ensign DL, Bruns CM & Pande VS. Accelerating molecular dynamic simulation on graphics processing units, Journal of computational chemistry 30, 864–872, (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Luttmann E, Ensign DL, Vaidyanathan V, Houston M, Rimon N, Oland J, Jayachandran G, Friedrichs M & Pande VS. Accelerating molecular dynamic simulation on the cell processor and Playstation 3, Journal of computational chemistry 30, 268–274, (2009). [DOI] [PubMed] [Google Scholar]
  • 107.Eastman P, Friedrichs MS, Chodera JD, Radmer RJ, Bruns CM, Ku JP, Beauchamp KA, Lane TJ, Wang LP, Shukla D, Tye T, Houston M, Stich T, Klein C, Shirts MR & Pande VS. OpenMM 4: A Reusable, Extensible, Hardware Independent Library for High Performance Molecular Simulation, J. Chem. Theo. Comp 9, 461–469, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Salomon-Ferrer R, Gotz AW, Poole D, Le Grand S & Walker RC. Routine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 2. Explicit Solvent Particle Mesh Ewald, J. Chem. Theory Comput 9, 3878–3888, (2013). [DOI] [PubMed] [Google Scholar]
  • 109.Beauchamp KA, Bowman GR, Lane TJ, Maibaum L, Haque IS & Pande VS. MSMBuilder2: Modeling Conformational Dynamics at the Picosecond to Millisecond Scale, J. Chem. Theory Comput 7, 3412–3419, (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Scherer MK, Trendelkamp-Schroer B, Paul F, Perez-Hernandez G, Hoffmann M, Plattner N, Wehmeyer C, Prinz JH & Noe F. PyEMMA 2: A Software Package for Estimation, Validation, and Analysis of Markov Models, J. Chem. Theory Comput 11, 5525–5542, (2015). [DOI] [PubMed] [Google Scholar]
  • 111.Meng Y, Shukla D, Pande VS & Roux B. Transition path theory analysis of c-Src kinase activation, Proc. Nat. Acad. Sci. U.S.A 113, 9193–9198, (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Harrigan MP, Sultan MM, Hernandez CX, Husic BE, Eastman P, Schwantes CR, Beauchamp KA, McGibbon RT & Pande VS. MSMBuilder: Statistical Models for Biomolecular Dynamics, Biophys. J 112, 10–15, (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Sultan MM, Kiss G & Pande VS. Towards simple kinetic models of functional dynamics for a kinase subfamily, Nat Chem 10, 903–909, (2018). [DOI] [PubMed] [Google Scholar]
  • 114.Keiser MJ, Irwin JJ & Shoichet BK. The chemical basis of pharmacology, Biochemistry 49, 10267–10276, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Meng Y, Gao C, Clawson DK, Atwell S, Russell M, Vieth M & Roux B. Predicting the Conformational Variability of Abl Tyrosine Kinase using Molecular Dynamics Simulations and Markov State Models, J. Chem. Theo. Comp 14, 2721–2732, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Paul F, Meng Y & Roux B. Identification of Druggable Kinase Target Conformations Using Markov Model Metastable States Analysis of apo-Abl, J. Chem. Theo. Comp 16, 1896–1912, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Paul F, Thomas T & Roux B. Diversity of Long-Lived Intermediates along the Binding Pathway of Imatinib to Abl Kinase Revealed by MD Simulations, J Chem Theo Comp 16, 7852–7865, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Meng Y & Roux B. Computational study of the W260A activating mutant of Src tyrosine kinase, Protein Science 25, 219–230, (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.LaFevre-Bernt M, Sicheri F, Pico A, Porter M, Kuriyano J & Miller WT. Intramolecular regulatory interactions in the Src family kinase Hck probed by mutagenesis of a conserved tryptophan residue, J. Biol. Chem 273, 32129–32134, (1998). [DOI] [PubMed] [Google Scholar]
  • 120.Fang L, Vilas-Boas J, Chakraborty S, Potter ZE, Register AC, Seeliger MA & Maly DJ. How ATP-Competitive Inhibitors Allosterically Modulate Tyrosine Kinases That Contain a Src-like Regulatory Architecture, ACS Chem Biol 15, 2005–2016, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Pond MP, Eells R, Treece BW, Heinrich F, Losche M & Roux B. Membrane Anchoring of Hck Kinase via the Intrinsically Disordered SH4-U and Length Scale Associated with Subcellular Localization, J. Mol. Biol 432, 2985–2997, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Xie T, Saleh T, Rossi P & Kalodimos CG. Conformational states dynamically populated by a kinase determine its function, Science 370, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Joshi MK, Burton RA, Wu H, Lipchik AM, Craddock BP, Mo H, Parker LL, Miller WT & Post CB. Substrate binding to Src: A new perspective on tyrosine kinase substrate recognition from NMR and molecular dynamics, Protein science: a publication of the Protein Society 29, 350–359, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Swendeman S, Nagar B, Wisniewski D, Strife A, Lambek C, Liu C, Veach D, Bornmann W, Kuriyan J & Clarkson B. Crystal structures of the c-Abl tyrosine kinase domain in complex with STI-571 and a novel Bcr-Abl inhibitor, PD1173955., Clin. Cancer Res 7, 3768s–3768s, (2001). [Google Scholar]
  • 125.Pontiggia F, Pachov DV, Clarkson MW, Villali J, Hagan MF, Pande VS & Kern D. Free energy landscape of activation in a signalling protein at atomic resolution, Nat. Commun 6, 7284, (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Wilson C, Agafonov RV, Hoemberger M, Kutter S, Zorba A, Halpin J, Buosi V, Otten R, Waterman D, Theobald DL & Kern D. Kinase dynamics. Using ancient protein kinases to unravel a modern cancer drug’s mechanism, Science 347, 882–886, (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Yang S, Blachowicz L, Makowski L & Roux B. Multidomain assembled states of Hck tyrosine kinase in solution, Proc. Natl. Acad. Sci. U.S.A 107, 15757–15762, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.McClendon CL, Kornev AP, Gilson MK & Taylor SS. Dynamic architecture of a protein kinase, Proc. Natl. Acad. Sci. U.S.A 111, E4623–E4631, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Prinz JH, Wu H, Sarich M, Keller B, Senne M, Held M, Chodera JD, Schutte C & Noe F. Markov models of molecular kinetics: generation and validation, J. Chem. Phys 134, 174105, (2011). [DOI] [PubMed] [Google Scholar]
  • 130.Nuske F, Wu H, Prinz JH, Wehmeyer C, Clementi C & Noe F. Markov state models from short non-equilibrium simulations-Analysis and correction of estimation bias, Journal of Chemical Physics 146, (2017). [Google Scholar]
  • 131.Weber M, Fackeldey K & Schutte C. Set-free Markov state model building, Journal of Chemical Physics 146, (2017). [DOI] [PubMed] [Google Scholar]
  • 132.Buch I, Giorgino T & De Fabritiis G. Complete reconstruction of an enzyme-inhibitor binding process by molecular dynamics simulations, Proc. Natl. Acad. Sci. U.S.A 108, 10184–10189, (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Plattner N & Noe 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]
  • 134.Wang J, Wolf RM, Caldwell JW, Kollman PA & Case DA. Development and testing of a general amber force field, J. Comp. Chem 25, 1157–1174, (2004). [DOI] [PubMed] [Google Scholar]
  • 135.Vanommeslaeghe K, Hatcher E, Acharya C, Kundu S, Zhong S, Shim J, Darian E, Guvench O, Lopes P, Vorobyov I & Mackerell AD Jr. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields, J. Comput. Chem 31, 671–690, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Huang L & Roux B. Automated Force Field Parameterization for Non-Polarizable and Polarizable Atomic Models Based on Target Data, J. Chem. Theo. Comp 9, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Singh J, Petter RC, Baillie TA & Whitty A. The resurgence of covalent drugs, Nat. Rev. Drug Discov 10, 307–317, (2011). [DOI] [PubMed] [Google Scholar]
  • 138.Klus S, Bittracher A, Schuster I & Schutte C. A kernel-based approach to molecular conformation analysis, Journal of Chemical Physics 149, (2018). [DOI] [PubMed] [Google Scholar]
  • 139.Rosta E & Hummer G. Free Energies from Dynamic Weighted Histogram Analysis Using Unbiased Markov State Model, J. Chem. Theory Comput 11, 276–285, (2015). [DOI] [PubMed] [Google Scholar]
  • 140.Mey ASJS, Wu H & Noe F. xTRAM: Estimating Equilibrium Expectations from Time-Correlated Simulation Data at Multiple Thermodynamic States, Phys. Rev. X 4, 041018, (2014). [Google Scholar]
  • 141.Wu H, Mey AS, Rosta E & Noe F. Statistically optimal analysis of state-discretized trajectory data from multiple thermodynamic states, J. Chem. Phys 141, 214106, (2014). [DOI] [PubMed] [Google Scholar]
  • 142.Wang J, Chmiela S, Muller KR, Noe F & Clementi C. Ensemble learning of coarse-grained molecular dynamics force fields with a kernel approach, The Journal of chemical physics 152, 194106, (2020). [DOI] [PubMed] [Google Scholar]
  • 143.Conrad TOF, Genzel M, Cvetkovic N, Wulkow N, Leichtle A, Vybiral J, Kutyniok G & Schutte C. Sparse Proteomics Analysis - a compressed sensing-based approach for feature selection and classification of high-dimensional proteomics mass spectrometry data, Bmc Bioinformatics 18, (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Rotskoff GM & Vanden-Eijnden E. Dynamical Computation of the Density of States and Bayes Factors Using Nonequilibrium Importance Sampling, Physical review letters 122, 150602, (2019). [DOI] [PubMed] [Google Scholar]
  • 145.Wang J, Olsson S, Wehmeyer C, Perez A, Charron NE, de Fabritiis G, Noe F & Clementi C. Machine Learning of Coarse-Grained Molecular Dynamics Force Fields, ACS Cent Sci 5, 755–767, (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Bittracher A, Klus S, Hamzi B, Koltai P & Schutte C. Dimensionality Reduction of Complex Metastable Systems via Kernel Embeddings of Transition Manifolds, J Nonlinear Sci 31, (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Pan AC & Roux B. Building Markov state models along pathways to determine free energies and rates of transitions, J. Chem. Phys 129, 064107, (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Ung PM, Rahman R & Schlessinger A. Redefining the Protein Kinase Conformational Space with Machine Learning, Cell Chem Biol 25, 916–924 e912, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES