Abstract
The impact of harmonic restraints on protein heavy atoms and ligand atoms on end-point free energy calculations is systematically characterized for 54 protein-ligand complexes. We observe that stronger restraints reduce the equilibration time and statistical inefficiency, suppress conformational sampling, influence correlation with experiment, and monotonically decrease the estimated loss of entropy upon binding, leading to stronger estimated binding free energies in most systems. A statistical estimator that reweights for the biasing potential and includes data prior to the estimated equilibration time has the highest correlation with experiment. A spring constant of 20 cal mol−1 Å−2 maintains a near-native energy landscape and suppresses artifactual energy minima while minimally limiting thermal fluctuations about the crystal structure.
Keywords: Molecular Mechanics, Entropy, Harmonic Restraints, Cumulant Expansion
Graphical Abstract
The effect of varying spring constants of harmonic restraints on the configuration space sampling and end-point binding free energy estimates of 54 protein-ligand complexes is assessed.
INTRODUCTION
We systematically vary the spring constant for harmonic restraints on backbone heavy atoms and ligand atoms, finding that a spring constant of 20 cal mol−1 Å−2 maintains a near-native structure and suppresses artifactual energy minima while minimally limiting thermal fluctuations about the crystal structure and reducing end-point entropy estimates to a lesser extent than stronger restraints.
Noncovalent binding free energy calculations, which have wide applications in chemical biology and structure-based drug design, are an important objective in computational chemistry. From the standpoint of statistical mechanics, the most rigorous approaches to these calculations are alchemical pathway techniques1. Alchemical pathway methods involve simulating a series of thermodynamic states between bound and unbound states. They are referred to as “alchemical” because intermediate states are not necessarily physically meaningful. In the unbound state, the species may either be physically separated or decoupled in a way such that they do not interact with each other. As reviewed by Mobley and Klimovich2, many scientists including Roux et. al.3–18 have contributed to developing both classes (decoupling or physical separation) of calculations. An accumulating number of examples19–26 have demonstrated the ability of alchemical pathway methods to accurately reproduce experimental binding free energies and provide mechanistic insight into binding processes.
Because of the computational expense required to simulate alchemical pathways, methods based on simulating the end points — the bound (and sometimes unbound) states that bookend the pathway — are widely used in the molecular modeling community. We summarized the history of end-point methods in a recent article27. Here, we will focus on variants of the most prominent approach: Molecular Mechanics/Poisson-Boltzmann Surface Area (MM/PBSA)28, 29. In MM/PBSA, the enthalpy of binding is estimated using the average potential energy of the solvated species. MM/PBSA has multiple variants in which the potential energy is based on different force fields30. The most popular of these is the Molecular Mechanics/Generalized Born Surface Area (MM/GBSA) method, which uses a faster continuum dielectric implicit solvent model.
In MM/PSBA and its variants, the average difference between the potential energy of the complex and individual binding partners is evaluated either from three independent trajectories or, more commonly, a single trajectory of the complex. These approaches are based on different theoretical principles. Calculations using a single trajectory are based on a first-order expansion of an exponential average27. Unfortunately, the single-trajectory approach can often be statistically biased. Exponential averages of potential energy differences can suffer from finite-sample bias if the important configuration spaces of the sampled and targeted thermodynamic states do not overlap31. In the case of binding free energies, the bias occurs if the ensemble of conformations accessed by either partner significantly changes upon binding, a common situation in protein-ligand complexes. In contrast to the single-trajectory approach, the use of three independent trajectories is based on estimating absolute partition functions of each species28, 29. Although the three-trajectory approach does not suffer from the same statistical bias, results are noisier and require a greater number of samples to apparently converge. Hence, there have been relatively few studies that employ the three-trajectory approach. For these reasons, the present contribution focuses on the single-trajectory approach.
Notable exceptions to the trend of using single trajectories are from the group of Peter Coveney, who has used sets of multiple short trajectories instead of a single long trajectory to compute binding free energies32–34. Within a set of trajectories, each is initiated from the same starting configuration but with different initial velocities. Compared to using a single long trajectory, using a set of short trajectories improves the precision, reproducibility, and parallelizability of free energy calculations. As a drawback, short trajectories also limit structural sampling to the vicinity of the initial configuration32. In spite of this shortcoming, Coveney et al. have found that when using multiple short trajectories, free energy calculations based on independent sampling of the complex and of the binding partners outperform those based on sampling of the complex alone. For binding of 12 peptides to a major histocompatibility complex molecule, end-point binding free energy calculations based on three sets of trajectories were more strongly correlated with experiment than those based on a single set of trajectories33. Wright et al.34 recognized that when computing the binding free energy of multiple ligands to the same protein receptor, the average energy of the protein is constant. For both a diverse and congeneric set of ligands to the bromodomain-containing protein 4, they found that free energies computed with sets of trajectories for the ligand and the complex were better correlated with experiment than sets of trajectories for the complex alone34.
In MM/PBSA and its variants, the entropy of binding is estimated via quasiharmonic or normal modes analysis. Unfortunately, both quasiharmonic and normal modes analysis are flawed ways to estimate entropy. Quasiharmonic analysis assumes that atomic fluctuations observed in a molecular dynamics simulation are a multivariate Gaussian resulting from a harmonic energy surface. Normal modes analysis also assumes that the energy landscape is harmonic, but spring constants are obtained from a Hessian matrix instead of inferred from observed samples. The assumptions underlying both methods fall short when degrees of freedom are coupled or the system samples multiple energy minima35, 36. Additionally, normal modes analysis tends to be computationally expensive and numerically unstable37–39. For these reasons, MM/PBSA calculations often entirely ignore entropy40, 41.
Recently, alternate approaches to calculating the entropy of binding based on an exponential average42 or cumulant expansion27 of the interaction energy have been introduced. In Menzer et al.27, we showed that the MM/PBSA enthalpy estimate can be derived as the first order in a cumulant expansion and that entropy may be evaluated with higher-order terms. While the cumulant expansion has been applied in other types of free energy calculations43, our paper was the first, to our knowledge, to apply it to absolute binding free energies and relate it with MM/PBSA. In contrast to normal modes analysis, these approaches consume trivial computing time beyond the simulation and post-processing required for standard MM/PBSA, are numerically stable, and improve (or at least do not deteriorate) correspondence with experiment. A limitation of these methods is that a larger number of samples is required to estimate second- or highe-rorder cumulants and exponential averages than to estimate the mean. For 200 ns of simulation, we found that a second-order cumulant expansion provided a good balance between statistical rigor and numerical stability.
In many of the simulations in Menzer et al.27, sampling issues were exacerbated by motion of the ligand relative to the protein. Within 200 ns of unrestrained simulation starting from a crystal structure, the ligand often assumed a non-native pose or completely dissociated from the protein. Although spontaneous dissociation (and re-association) is expected over a long simulation, the relatively fast ligand motion observed was likely due to limitations of the force field. We found that correspondence with experimental binding free energies was improved by excluding snapshots in which the ligand root mean square deviation (RMSD) from the native pose is greater than 3 Å. Excluding non-native poses is likely helpful because they are unlikely to significantly contribute to the binding free energy. Excluding dissociated poses is essential because they are not part of the bound state.
Harmonic restraints have the potential to circumvent force field issues and improve sampling of near-native poses. Such restraints have yielded similar benefits in at least three previous studies. Raval et al.44 assessed the ability of long all-atom molecular dynamics simulations to refine homology models. For 24 proteins selected from a structure prediction challenge, they performed at least 100 microseconds of three types of simulations: unrestrained, initiated from a homology model; unrestrained, initiated from the crystal structure; and restrained to the homology model. Unrestrained simulations initiated from homology models and crystal structures both generally drifted away from the native structure, a fact they attributed to force field (opposed to sampling) error. On the other hand, most of the weakly restrained simulations drew subtly closer to the respective crystallographic configuration. Xue and Skrynnikov45 performed simulations of ubiquitin crystals without restraints and with harmonic restraints applied to the ensemble-average structure. They found that the latter led to better reproduction of crystallographic temperature factors and solid-state NMR spectroscopy parameters. Finally, Wall et al.46 performed unrestrained and restrained simulations of endoglucanase crystals. The restrained simulations were much more successful at recovering crystallographic positions of solvent.
Apropos to the topic of this paper, harmonic restraints are beneficial to and are routinely used in alchemical pathways for computing binding free energies. In pathways where species are physically separated, harmonic restraints are placed on the distance between binding partners6, 47–50. In both physical separation and decoupling pathways, restraints are usually also placed on the orientation and position of the ligand relative to the protein3, 4, 6–8, 13–16, 20, 51–54. The free energy of imposing restraints on external degrees of freedom (translation and rotation of the ligand) can be computed analytically or by a fast numerical quadrature. The effects of other restraints are removed before the end points of the pathways such that they do not affect binding free energy estimates. These restraints are beneficial because they limit the configuration space of intermediate states, simplifying sampling and reducing equilibration and autocorrelation times within the reduced space.
While restraints can maintain near-native configurations and improve sampling within a reduced configuration space, it is not obvious, a priori, how they should be applied to end-point binding free energy calculations. First, it is not clear which atoms they should be applied to. They could be applied to all protein atoms, protein heavy atoms, backbone atoms, or only α carbons. They could be applied to all ligand atoms or simply to the center of mass. Moreover, it is unclear what spring constant should be used. In an alchemical pathway, multiple thermodynamic states with different restraint levels may be used to release a strong restraint. In an end-point calculation, an excessively strong restraint can truncate the accessible configuration space. For example, in Figure 1, adding a restraint with a spring constant of 2 kcal mol−1 Å−2 will likely prevent a simulation from accessing the two free energy minima within a reasonable amount of sampling. On the other extreme, a restraint with a spring constant of 2 cal mol−1 Å−2 will have little influence on the simulation whatsoever. Therefore, spring constants should be selected to balance sampling near a desired configuration with the ability to explore relevant configuration space.
Unfortunately, most of the scientific literature does not provide clear guidance on how to apply harmonic restraints. More often than not, choices of restrained atoms and spring constants are treated as minor technical details that are stated without justification. These choices vary widely across different papers. On the weak end, Raval et al.44 used a spring constant of 70 cal mol−1 Å−2 to restrain α carbon atoms within secondary structure elements to their positions in low-resolution homology models. Wall et al.46 used a stronger spring constant of 0.5 kcal mol−1 Å−2 to restraint protein heavy atoms and all ligand atoms to their crystallographic positions. In a series of simulations used to build a Markov state model of an enzyme-inhibitor binding process, Buch et al.56 applied a harmonic restraint of 1 kcal mol−1 Å−2 to α carbon atoms farther than 9 Å from the binding pocket. In their end-point binding free energy calculations, Duan et al.42 used an even stronger spring constant of 10 kcal mol−1 Å−2 to restrain all atoms.
An important precedent for the principled selection of spring constants was provided by Hamelberg and McCammon52. Hamelberg and McCammon52 performed alchemical binding free energy calculations of water to two protein-ligand complexes. To determine parameters for a harmonic potential restraining water to the binding site, they simulated the complex and observed fluctuations of the water’s oxygen atom. They subsequently selected spring constants that, according to the equipartition theorem, would reproduce the observed fluctuations. The spring constants were 6.6 kcal mol−1 Å−2 for a complex of trypsin and benzyldiamine complex and 16.4 kcal mol−1 Å−2 for a complex of HIV-1 protease and its inhibitor KNI-272.
The equipartition theorem can also inform the selection of spring constants in our more general calculations. The theorem specifies that any degree of freedom with a quadratic energy function has an average energy of . A harmonic restraint has the form , where K is the spring constant and d is the displacement from the restraint center. Hence, if the energy for a displacement from a restraint center is described solely by the restraint, then the root mean square displacement is . When K is 0.2, 2, 20, 200, and 2000 cal mol−1 Å−2, the corresponding are 54.6, 17.3, 5.5, 1.7, and 0.5 Å, respectively. Since the system includes other energy terms, equipartition only provides a reasonable estimate of thermal fluctuations in d for the strongest restraints.
In X-ray crystal structures, the extent of thermal fluctuations may be inferred from Debye-Waller factors (also known as B values). The B value of a particular atom is based on thermal fluctuations, experimental error, model error, and lattice error. For well-solved structures, model error will likely contribute very little to B values. While experimental error is impossible to completely discount, structures with small R Merge values will also have small experimental error. Furthermore, when considering protein backbone atoms, the contribution from lattice error is expected to be very low. Thus, by converting B values to spherical fluctuations, a good approximation to thermal fluctuations can be obtained: (c.f. Bahar et al.57).
As a representative example, consider a structure in the Protein Data Bank: 3mxf55. B values for protein backbone atoms are generally between 5 and 20 Å2, with the mean being 9.2 Å2. Thus, we can estimate that under crystallographic conditions, where the temperature is approximately 90 K, is generally between 0.44 and 0.87 Å. Assuming that fluctuations are described by a quadratic energy, and scaling them according to the equipartition theorem, at 300 K should be between 0.80 and 1.6 Å. Hence, restraints of 2 kcal mol−1 Å−2 are likely to suppress thermal fluctuations of at least some atoms in the crystal structure. They may have an even more significant effect on fluctuations in solution, which are presumably larger.
In this work, we systematically assess the effects of varying the spring constant from 0.2 to 2000 cal mol−1 Å−2. Restraints were applied to protein backbone heavy atoms and ligand atoms, a selection that allows side chains to relax. Simulations were performed on 54 protein-ligand complexes that make up a significant subset of systems from Menzer et al.27. All of the systems have both crystal structures and binding affinity data publicly available. Subsequently, they were analyzed to better understand how the spring constant affects the equilibration time and statistical inefficiency, the conformational sampling, the estimated entropy and free energy of binding, and the accuracy of binding free energy estimates compared to experimental data.
METHODS
System Preparation
We performed simulations with most of the systems from Menzer et al.27: eight complexes with the first bromodomain of human bromodomain-containing protein 4 (bromodomains); seven complexes with human phenylethanolamine N-methyltransferase (hPNMT); nine complexes with lysozyme L99A (lysozyme); 17 complexes with mitogen-activated protein kinase kinase kinase kinase 4 (MAP4K4); and 13 complexes from the interaction entropy (IE) data set42. The farnesoid X receptor (FXR) data set was excluded because of its high computational cost and poor performance in Menzer et al.27.
Due to the inability to prepare two systems - MAP07 from MAP4K4 and 4dew from IE - using the methods described below, they were excluded from the current study. In these systems there were problems with the file containing the parameters and topology. In both cases remedies were not pursued so as to preserve the consistency of system preparation.
Models were prepared for simulation using the same procedure as in Menzer et al.27, except that protein parameters were from the AMBER58 ff14SB (instead of ff14SB.redq) force field. Some key details are that residue protonation and disulfide bond formation was predicted and assigned using PROPKA 3.059, 60 with a pH value of 7.0, as implemented in PDB2PQR 1.9.061. The mbondi2 option was used to set atomic radii. Ligands were parametrized with the generalized AMBER force field 2 (leaprc.gaff2)62, with AM1BCC63, 64 partial charges. Solvent was modeled with OBC265 generalized Born surface area implicit solvent66 with a solvent dielectric of 78.5. No cutoffs were used.
Molecular Dynamics Simulation
Seven molecular dynamics simulations were performed for every complex. In one of the simulations, the system was unrestrained. In five others, harmonic restraints were applied to protein backbone heavy atoms and all ligand atoms. The restraint potential was,
(1) |
where K is the spring constant, xo, yo, and zo are the initial x, y, and z positions of each restrained atom, respectively, and x, y, and z are the positions of the atom in the current frame. The spring constants used were K ∈ {0.2, 2, 20, 200, 2000} cal mol−1 Å−2. Lastly, to compare the convergence when restraints are placed on different atoms, one set of simulations was performed using harmonic restraints with a spring constant of 20 cal mol−1 Å−2 on all heavy atoms (opposed to heavy atoms in the protein backbone).
Simulations were performed using OpenMM 7.067, 68. First, systems were minimized for 500 steps without restraints. Then, Langevin dynamics was performed at 300 K with a time step of 2 fs and constraints on bonds involving hydrogen. Restrained simulations were conducted for 50 ns and unrestrained simulations for 200 ns. Harmonic restraints towards the initial positions were implemented with the CustomExternalForce class in OpenMM. Energies were recorded every 2 ps.
Equilibration Detection
The equilibration time and statistical inefficiency of the MD simulations was calculated using the procedure introduced by Chodera et al.69. This procedure is based on the effective sample size Neff = N/g, where N is the number of correlated samples, g = 1 + 2τac is the statistical inefficiency, and τac is the integrated autocorrelation time. For a simulation of length T, the effective sample size for the last T − to correlated samples is estimated for various proposed equilibration times, to. The equilibration time is the value of to that maximizes the effective sample size, , and the statistical inefficiency is the estimate of g at . We used the implementation of the method in detectEquilibration from the timeseries module of pymbar70 to analyze the time series of the total potential energy. We chose nskip = 10 and Neff was estimated every 1 ns of simulation.
Conformational Clustering
To facilitate conformational clustering, MD trajectories were analyzed with principal components analysis (PCA). Trajectory files from all six simulations of each protein-ligand system were concatenated with catdcd71. Using MDTraj72 1.7.2, snapshots were aligned to minimize the RMSD of all atoms with the first frame. Eigenvalues and eigenvectors of the covariance matrix for all atoms were calculated with ProDy73 1.8.2.
Snapshots from simulations of every protein-ligand system were clustered using the algorithm described by Rodriguez and Laio74. In this algorithm, clusters are defined based on the density of every point. The density of a point is based on how many neighbors — points located within a specified distance cutoff — it has. A cluster center must have a higher density than its neighbors and must be at least a minimum distance away from a point of equal or higher density. Distance matrices were calculated based on the RMSD between all protein α carbons for every pair of frames in combined trajectories. Distance cutoffs were selected such that the average point in the unrestrained simulation was considered to be a neighbor of 1.5% of other points in the unrestrained simulation. The minimum density was set at 100 and minimum distance between cluster centers was set at 1.5 times the distance cutoff. This clustering algorithm has the advantage of grouping together conformations regardless of the cluster shape.
Binding Free Energy Estimation
After simulation, separate trajectories of the ligand and receptor were prepared using VMD 1.9.175. Potential energies for the complex, receptor, and ligand were evaluated with OpenMM 7.067, 68 using the same force field as in the simulation.
Binding free energies were calculated based on trajectories of the complex (single-trajectory mode) using variations of the statistical estimators described in Menzer et al.27. Specifically, we calculated binding free energies: (1) with and without reweighting based on the biasing potential; (2) either discarding or retaining an initial period; (3) with and without a ligand RMSD cutoff of 3 Å; (4) with a first- to fourth-order cumulant expansion or an exponential average. Because they were found to be ineffective in Menzer et al.27, we did not consider an external entropy correction nor normal modes analysis to estimate configurational entropy.
Before describing modifications to the statistical estimators in Menzer et al.27 to account for harmonic restraints, we will introduce some key notation and equations. The noncovalent binding of a receptor, R, and ligand, L, to form a complex, RL, is described by the chemical equation L + R ⟷ RL. The standard binding free energy, , is defined as,
(2) |
in which kB is Boltzmann’s constant and T is the temperature in Kelvin. CX is the equilibrium concentration of the species X ∈ {R, L, RL} and C° = 1 M is the standard concentration. The standard binding free energy can be calculated from the free energy of restricting the ligand into the binding site, ΔGξ, and of coupling the ligand in the binding site to the receptor, . According to Equation (15) of Menzer et al.27, may be computed by an exponential average,
(3) |
where Ψ(rRL) = U(rRL)−U(rR)−U(rL) is the interaction energy between the receptor and ligand. rX for X ∈ {R, L, RL} is the conformation of the species. U(·) is the potential energy of the species, including the solvation free energy, which may be based on a continuum dielectric implicit solvent model. The angular brackets 〈…〉RL denote an ensemble average of the unbiased complex in implicit solvent.
Because exponential averages may be numerically unstable, it can be helpful to estimate via a cumulant expansion, a power series in β43. As previously described in Equation (16) of Menzer et al.27, the third-order cumulant expansion of Equation (3) is,
(4) |
where Ψ ≡ Ψ(rRL) and δΨ ≡ Ψ(rRL) −〈Ψ(rRL)〉 are used within the brackets for convenience of notation. Assuming that the pressure and volume change negligibly upon ligand binding, the mean interaction energy 〈Ψ〉RL, which is the first-order cumulant, can be equated with the enthalpy of binding. Higher-order cumulants are estimates of the entropy.
Computing from biased simulations of the complex in implicit solvent is an example of importance sampling. Specifically, Equation (3) may be written as,
(5) |
where Ub(rRL) is the biasing potential on the configuration and the angular brackets 〈…〉RL, b denote an ensemble average of the biased complex in implicit solvent. Both of the expectation values in Equation (5) may be expressed as cumulant expansions. If they are both expanded to the first order, 〈Ub(rRL)〉RL, b cancels out and the free energy estimate is simply 〈Ψ〉RL, b. Higher-order cumulants lead to expressions distinct from Equation (4).
We also calculated binding free energies based on including or excluding an initial period. The initial period was the smaller of the equilibration time determined by pymbar70, 76 or half of the included frames. (If the ligand RMSD cutoff was used, the included frames were those within the cutoff. Otherwise, they were all the frames.) This alternate initial period ensures that even if the equilibration period is long, there is an adequate number of samples available for the free energy calculation.
In addition to the end-point binding free energy calculations, we also estimated the binding free energy based on the interaction energy in a single minimized structure of the complex. Minimization was performed using simulation.minimizeEnergy() with default parameters in the python API for OpenmM 7.067, 68.
RESULTS
Stronger restraints reduce equilibration time and statistical inefficiency and accelerate convergence
Across all the data sets, the estimated equilibration time is highly dependent on the spring constant (Figure 2). With a spring constant of 20 cal mol−1 Å−2 or less, most of the estimated equilibration times are a large fraction of the simulation length. In general, as the spring constant is increased, the estimated equilibration time consumes a smaller fraction of the trajectory. At a spring constant of 20 cal mol−1 Å−2, the estimated equilibration time still consumes the majority of the simulation time in most systems, but in many cases it is significantly shorter. With a spring constant of 2 kcal mol−1 Å−2, equilibration is almost always very rapid, rarely requiring over 0.1 of the simulation length.
The statistical inefficiency is also reduced by a larger spring constant (Figure 3). Overall, the statistical inefficiency is largest for the unrestrained simulation. Many unrestrained simulations of the bromodomains have particularly high statistical inefficiency. The general trend across data sets is that statistical inefficiency gradually decreases as restraints are strengthened. At a spring constant of 2 kcal mol−1 Å−2, nearly every sample can be considered independent.
Finally, stronger restraints also affect convergence to the final free energy estimate (Figures S1, S2, S3, S4, and S5 of the Supplementary Material.) For spring constants of 20 cal mol−1 Å−2 and stronger, the root mean square error (RMSE) of the first- and second-order cumulant expansions and exponential average with respect to the final value after 20 ns is within 1 kcal/mol for most and 2 kcal/mol for nearly all data sets. For weaker restraints, the RMSE is often but not always higher after a longer simulation time. The first-order cumulant expansion is more numerically stable than the second-order cumulant expansion and the exponential average, where the RMSE fluctuates more significantly as a function of simulation time. Surprisingly, applying harmonic restraints to all heavy atoms, opposed to backbone heavy atoms, does not accelerate simulation convergence (Figure S6 of the Supplementary Material).
Sufficiently strong restraints suppress configuration space sampling
The RMSD of the protein backbone relative to the crystal structure indicates the extent to which configuration space sampling is suppressed by harmonic restraints (Figure 4). As the spring constant is reduced, the harmonic restraint has a less important role in dictating the observed fluctuations. With at spring constant of 2000 cal mol−1 Å−2, all the values are only slightly less than 0.5 Å, the RMSD predicted based on the equipartition theorem if the harmonic restraint is the only potential energy term. At 200 cal mol−1 Å−2 and weaker restraints, the RMSD is significantly less than the equipartition estimate, indicating that other potential energy terms also play an important role in determining the extent of fluctuations. With the restraint of 0.2 cal mol−1 Å−2, the observed range in RMSD is comparable to the unrestrained simulation, but the mode is lower, at around 2.0 instead of 2.5 Å.
Configuration space sampling was also evaluated by the number of clusters (according to the first two principal components) accessed during the simulations (Figure 5). The unrestrained simulations (which are 200 ns long) access the greatest number of clusters. Simulations with restraint of 0.2 cal mol−1 Å−2 (which are 50 ns long) access slightly fewer clusters. The number of clusters is further reduced as the spring constant is increased from 0.2 and 2 cal mol−1 Å−2. At spring constants of 20 cal mol−1 Å−2 and stronger, simulations only access a single cluster. In the majority of cases, the clusters occupied by tightly restrained simulations are a subset of the clusters occupied by the less restrained and unrestrained simulations.
The two metrics — the average RMSD of the protein backbone and the number of occupied clusters — are closely related to one another (Figure 6). For simulations that occupy a single cluster, the average protein backbone RMSD spans a large range from below 0.5 Å to slightly more than 2.5 Å, with a diminishing number of observations at larger RMSD values. Simulations that occupy more than cluster have a minimum protein backbone RMSD of around 1.5 Å. At an average protein backbone RMSD of 2.0 Å, there are more simulations that occupy multiple clusters than a single cluster. Above 2.5 Å, nearly all simulations occupy multiple clusters.
Numerous factors affect correlation with experiment
As in Menzer et al.27, we systematically considered the influence of statistical estimator options on correlation with experiment. The correlation metrics used were the Pearson R, Spearman ρ, and Kendall τ77. The effect of these correlation metrics were ranked for the bromodomains, lysozyme, and MAP4K4 data sets, for which correlation metrics for most estimator options was generally greater than 0.4. Comparing rankings across all estimator options, consideration of the harmonic bias on the weight of each sample (Equation 5) was found to be beneficial (Table S1 of the Supplementary Material), especially with the Spearman ρ and Kendall τ. Using reweighting and comparing rankings against all other estimator options, including the full simulation opposed to discarding an equilibration time was found to provide much better correlation with experiment (Table S2 of the Supplementary Material). Using reweighting and including the full simulation, keeping all snapshots opposed to removing those with the ligand RMSD greater than 3 Å leads to slightly higher correlation with experiment (Table S3 of the Supplementary Material). Finally, the first-order cumulant expansion was found to outrank the second-order cumulant expansion, followed by the exponential average and then higher-order expansions (Table S4 of the Supplementary Material).
Due to the unexpectedly strong performance of the first-order cumulant expansion, we considered the correlation metrics for different data sets as a function of the spring constant (Figures S7, S8, and S9 of the Supplementary Material). This analysis reveals only slight differences in the first- and secon-dorder cumulants and the exponential average. The performance of the third- and fourth-order cumulant expansions strongly fluctuates with the spring constant.
Estimator options for the remaining free energy calculations in this paper were selected based on this systematic comparison. The biasing potential due to harmonic restraints was used to weight averages of the interaction energy (Equation 5). All interaction energies, including those from the estimated equilibration period, were included in the averages. For restrained simulations, no ligand RMSD cutoff was used. For unrestrained simulations, we applied our previous recommendation27 of excluding snapshots with a ligand RMSD greater than 3 Å.
In general, correlation with experiment was maximized at a restraint strength of 20 or 200 cal mol−1 Å−2 (Figure 7). For the bromodomains, correlation peaked at 200 cal mol−1 Å−2 and gradually deteriorated with weaker restraint strengths. The performance of lysozyme is fairly stable across restraint strengths but is weaker at spring constants of 2 or 0.2 cal mol−1 Å−2. It is slightly better for the unrestrained simulation, which is longer. The MAP4K4 data set followed a similar trend as lysozyme. For all systems, the interaction energy of a single minimized structure had similar correlation with experiment as the strongly restrained simulations.
In addition to correlations, we also considered additional metrics: the adjusted root mean square error (aRMSE) and the slope and intercept of a linear regression (Figure S10 in the Supplementary Material). The aRMSE is,
(6) |
where the and are the sample mean of x and y, respectively. The aRMSE accounts for systematic deviation between the series and is useful for assessing whether relative binding free energies are accurate. For the lysozyme data set, these metrics are fairly consistent across all restraint strengths, except for a reduction in all quantities at 2 cal mol−1 Å−2. For the bromodomains data set, estimator performance is steady for spring constants greater than 20 cal mol−1 Å−2; the aRMSE is higher for weaker restraints. In the MAP4K4 data set, the aRMSE is particularly high at 20 cal mol−1 Å−2 but the slope of the linear regression is also closest to zero. The large shift in the statistics at this restraint strength is likely due to a small number of outlier systems.
Stronger restraints reduce entropic contributions to binding
Unsurprisingly, in the vast majority of systems, the estimated entropic contribution to the binding free energy, , is reduced as restraint strength is increased (Figures 8, 9, S4, S5, and S6). Because binding generally leads to a loss in entropy, ΔS < 0, stronger restraints make a smaller positive quantity. Overall, the trend appears to be monotonic. However, in some systems the estimated entropy of binding for the unrestrained system is contrary to the trend. In these calculations, exclusion of snapshots for which the ligand RMSD is greater than 3 Å likely reduces the estimate of variance of interaction energy, 〈δΨ〉RL, such that is smaller.
In contrast to the trend in , the effects of restraints on the estimated enthalpy of binding are inconsistent. In the bromodomains, MAP4K4, and IE data sets, the enthalpy of binding increases in some systems and decreases in others (Figures 8, 9, and S13 of the Supplementary Material). In lysozyme, the binding enthalpy near universally decreases with increased spring constants (Figure S11 of the Supplementary Material). On the other hand, it increases with the spring constant for the entire hPNMT data set (Figure S12 of the Supplementary Material). In all systems, the enthalpy of binding is estimated to be lower for the interaction energy of the minimized structure than for the strongest restraint. The value is lower because it is based on a minimized structure rather than the mean of a configurational distribution.
Combining consistent trends in with inconsistency in trends, stronger restraints, more often than not, lead to stronger estimated ΔG. The entropic contribution dominates in most of the bromodomains, MAP4k4, and IE data sets (Figures 8, 9, and Figure S13 of the Supplementary Material). In lysozyme, stronger restraints decrease both and such that the estimated binding is stronger (Figure S11 of the Supplementary Material). On the other hand, because stronger restraints in the hPNMT data set increase , they lead to weaker binding free energy estimates in most of the data set (Figure S12 of the Supplementary Material). For all systems, the binding free energy estimate based on the interaction energy of a single minimized structure is nearly always lower than estimates that consider entropy losses.
DISCUSSION
Many systems have force field error
Although crystals are subject to packing forces that are not present in dilute solution, crystal and solution structures are generally consistent. In the context of our calculations, support for this statement comes from the fact that including the initial samples in calculations (which are similar to the crystal structure) improves correlation with experiment. If the initial samples were not representative of the experimental energy landscape, they would probably not improve the correlation between calculations and experimental binding free energies.
In light of the consistency between crystal and solution structures, molecular simulations of protein systems in dilute solution should sample an ensemble of structures dominated by near-crystallographic structures. Other conformations may be present but less populated. In order for simulations to sample such an ensemble, the energy landscape of the model should have a global minimum similar to the crystal structure. Unfortunately, there are several lines of evidence that in many of the unrestrained and weakly restrained simulations, this is not the case and there is force field error.
One piece of evidence for force field error is the long estimated equilibration time of unrestrained and weakly systems. If the global minimum of an energy landscape resembles the crystal structure, then simulations initiated from near crystal structures should not take long to equilibrate. In many simulations with a spring constant of 20 cal mol−1 Å−2 or less, however, the estimated equilibration time is a large fraction of the simulation length (Figure 2). In these cases, it is questionable that equilibration has actually occurred. A system can be considered equilibrated when estimated expectation values are independent of initial conditions. In Chodera’s equilibration detection method76, including data prior to the estimated equilibration time leads to long autocorrelation times and a smaller effective sample size; the estimated expectation value is perturbed by these early values of the observable (in our case, the total potential energy). If the estimated equilibration time is near the end of the simulated trajectory, then there is a high likelihood that it will continue to change with additional simulation time. If a simulation initiated from near the crystal structure has not equilibrated after 50 ns or 200 ns of simulation, then its energy landscape probably has a different global minimum.
The multiple conformations accessed in unrestrained and weakly restrained systems (Figure 5) are also likely to reflect force field error. In principle, it is possible for the experimental solution ensemble to include multiple conformations. Indeed, multiple conformers have been considered in some binding free energy calculations78–82. However, the fact that suppressing these non-native conformations with a restraint of 20 cal mol−1 Å−2 improves the correlation between calculations and experimental binding free energies suggests that, for the selected systems, the conformations are force field artifacts.
Finally, shifts in as a function of restraint strength also suggest force field error. If a modeled energy landscape is similar to a crystal structure, then changing spring constants should not strongly affect mean interaction energy. In many of the selected systems, however, increasing the restraint strength leads to apparently monotonic shifts in (Figures 8, 9, S4, S5, and S6).
More likely than not, the observed force field error is not unique to our calculations. To our knowledge, equilibration time estimates have not been reported in any other end-point binding free energy calculations. (Vosmeer et al.83 reported automated detection of transitions between conformations that contribute to end-point binding free energy estimates). Given that Chodera’s method for automated equilibration detection76 is fairly new, this is unsurprising. However, it is probable that most simulations run with the same or a similar force field, including the implicit solvent model, have a modeled energy landscape where the global minimum is distinct from the crystal structure. As most MM/GBSA calculations are based on much shorter simulations, it is likely that many of the simulations have not equilibrated and are still relaxing towards the global minimum of the model. Consequently, estimated ensemble averages are likely to dependent on the starting model. Mobley84 has suggested that this problem is widespread in molecular simulation.
A comparison of the average protein backbone RMSD and the number of clusters accessed (Figure 6) suggests that the former may be used to diagnose the presence of non-crystallographic conformations in simulations performed with or without restraints. If the average protein backbone RMSD is below 1.5 Å, then the simulation is sampling within the crystallographic conformation. If it is above 2.0 Å, then there is a good chance that it is sampling non-crystallographic conformational minima that may be artifactual. If it is above 2.5 Å, then it is almost certainly sampling these minima. As a caveat, while we have selected an algorithm that groups together snapshots regardless of cluster shape, this suggestion is likely sensitive to the algorithm used to identify distinct conformations.
Restraints shift global minima towards crystal structures
Fortunately, our results also suggest that sufficiently strong harmonic restraints on protein backbone atoms and ligand atoms can ameliorate force field error and shift global minima towards the crystal structure. This result may be unsurprising, but it is not completely trivial because side chains are unrestrained. Side chain motions could allow access to energy minima distinct from the crystal structure. However, the short estimated equilibration times for highly restrained systems (Figure 2) indicate that little conformational exploration away from the crystal structure is required for equilibration. Moreover, restraints of 20 cal mol−1 Å−2 or stronger are sufficient to suppress sampling of degrees of freedom that contribute to the largest variance in weakly restrained and unrestrained systems (Figure 5). As discussed above, the additional conformations are likely to be artifactual.
Our present results are consistent with Raval et al.44, who found that restraints on an even smaller subset of atoms (α carbons) to homology models were sufficient to make the global minimum of their models more consistent with crystal structures. They also agree with other previously-mentioned papers45, 46 which have found that restrained simulations are more consistent with experimental observations than their unrestrained counterparts.
Restraints should be removed from ensemble averages
Given that restraints shift the global minimum towards the crystal structure, it is reasonable to think that they should simply be considered part of the force field. Following this logic, a simple average (e.g. Equation 3) would be used instead of an importance sampling estimator (e.g. Equation 5). We found, however, that importance sampling estimators are more strongly correlated with experiment than simple averages. The stronger performance of importance sampling estimators suggests that the restraints should not be considered part of the force field. This conclusion is reasonable because the harmonic restraints were added without a physical rationale.
A spring constant near 20 cal mol−1 Å−2 is just about right
Based on our systematic assessment of the effect of restraint strength on equilibration time (Figure 2), statistical efficiency (Figure 3), the number of accessed clusters (Figure 5), correlation with experiment (Figure 7), and contributions to free energy (Figures 8 and 9, and S11, S12, and S13 of the Supplementary Material), we recommend using harmonic restraints with a spring constant around 20 cal mol−1 Å−2. Spring constants of 2 cal mol−1 Å−2 or less are too weak. Compared to unrestrained simulations, they have little impact on the equilibration time, statistical inefficiency, number of clusters accessed, and binding free energy. In these simulations, the estimated −TΔS are large and probably unrealistic. On the other hand, a spring constant of 20 cal mol−1 Å−2 appears to be just right. It reduces the equilibration time and statistical inefficiency in a significant subset of systems, suppresses sampling of likely artefactual conformations, leads to near-peak correlation with experimental ΔG. A spring constant of 200 cal mol−1 Å−2 may be too strong. Even though it reduces the equilibration time and statistical inefficiency relative to 20 cal mol−1 Å−2, it further reduces . A spring constant of 2 kcal mol−1 Å−2 is definitely too strong. Although it leads to a short equilibration time for nearly all systems and low statistical inefficiency for all systems, is significant suppressed, weakening correlation with experiment. Finally, although using a single minimized structure for binding free energy estimates may not significantly alter correlation with experiment and is comparatively very fast, requiring only seconds to calculate, an entropy loss of zero is not realistic.
Our results are consistent with predicted influences of spring constants on thermal fluctuations based on Debye-Waller factors and the equipartition theorem. The strong suppression of with a spring constant of 2 kcal mol−1 Å−2 is consistent with the small root mean square displacement anticipated. With a spring constant of 200 cal mol−1 Å−2, for a purely harmonic system is predicted to be 1.7 Å, which is just at the limit of expected variance around the crystallographic pose. Combined with the unbiased potential, this harmonic restraint still restricts sampling of the energy well. For a purely harmonic system, fluctuations allowed by a restraints with our recommended spring constant, 20 cal mol−1 Å−2, are much larger than those expected from crystallographic Debye-Waller factors. Although our recommended restraint prevents sampling of additional energy minima, it does not appear to suppress thermal fluctuations about the crystal structure.
CONCLUSIONS AND FUTURE DIRECTIONS
In conclusion, harmonic restraints with a spring constant around 20 cal mol−1 Å−2 on protein backbone heavy atoms and ligand atoms are a helpful way to maintain a near-crystallographic structure of a protein-ligand complex and perform an end-point binding free energy calculation that apparently converges by around 20 ns of simulation. Using an importance sampling estimator and including initial samples in calculations improve correlation with experiment.
Although our simulations were initiated from crystal structures, we selected a restraint procedure that could be relevant to simulations from docked structures. Restraints on protein backbone heavy atoms and ligand atoms allow side chains to relax, which could lead to energetically favorable interactions. A potentially fruitful future direction would be to assess the described procedure for docked structures. It could be helpful to more thoroughly explore other selections of restrained atoms, including all protein heavy atoms, all atoms, or only α carbons.
Other possibly useful directions include: end-point binding free energy calculations based on three trajectories (of the receptor, ligand, and complex) instead of a single trajectory of the complex; end-point calculations of relative binding free energies using restraints, building on Dolenc et al.85; explicit solvent simulations; longer simulations to more fully characterize trends in equilibration time; and comparing results of end-point calculations to alchemical (opposed to experimental) ΔG, which would isolate the effect of force field and estimator error when comparing statistical estimators.
Supplementary Material
ACKNOWLEDGEMENTS
I (DDLM) thank Benoit Roux for being a supportive mentor. In 2009, I started as a Director’s Postdoctoral Fellow (a semi-independent position) at Argonne National Laboratory, sponsored by Lee Makowski. After I reached out to him, Benoit allowed me to participate in his group meetings. When Makowski moved to Northeastern University in 2010, Benoit graciously agreed to sponsor my fellowship for another year. I was one of Benoit’s worst postdocs, as I never worked on any projects with him. Nonetheless, he continued to meet with me to discuss science and career plans on a regular basis.
We thank Andrew Howard (IIT) for helpful discussions about crystallographic B factors. Computing resources were provided through the Extreme Science and Engineering Discovery Environment (XSEDE)86, which is supported by National Science Foundation grant number ACI-1548562. Use of XSEDE Bridges GPU at the Pittsburg Supercomputer Center was provided through allocation TG-MCB150144. Some MD simulations were also performed on a local cluster at IIT. This research was supported by the National Institutes of Health (R01GM127712).
Footnotes
Additional Supplementary Material may be found in the online version of this article. Figures S1 through S6 are time series. They show the Pearson R with respect to experiment and the RMSE with respect to the final value as a function of simulation time. Figures S7 through S9 show the Pearson R, Spearman ρ, and Kendall τ as a function of restraint strength for the different expansion options. Figure S10 shows the aRMSE and slope and intercept of a linear regression as a function of restraint strength for the different data sets. Figures S11 through S13 show free energy components as a function of the restraint strength. Tables S1 to S4 compare statistical estimator options based on the average rank of correlation metrics. Tables S5 and S6 show free energy estimates. Table S7 shows correlation with experiment for free energy estimates based on a single structure.
References
- 1.Gilson MK, Given JA, Bush BL, and McCammon JA, Biophys. J 72, 1047 (1997). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Mobley DL and Klimovich PV, J. Chem. Phys 137, 230901 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Deng Y and Roux B, J. Chem. Theory Comput 2, 1255 (2006). [DOI] [PubMed] [Google Scholar]
- 4.Wang J, Deng Y, and Roux B, Biophys. J 91, 2798 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Deng Y and Roux B, J. Chem. Phys 128, 115103 (2008). [DOI] [PubMed] [Google Scholar]
- 6.Deng Y and Roux B, J. Phys. Chem. B 113, 2234 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Ge X and Roux B, J. Phys. Chem. B 114, 9525 (2010). [DOI] [PubMed] [Google Scholar]
- 8.Ge X and Roux B, J. Mol. Recognit 23, 128 (2010). [DOI] [PubMed] [Google Scholar]
- 9.Jiang W and Roux B, J. Chem. Theory Comput 6, 2559 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Jiang W, Luo Y, Maragliano L, and Roux B, J. Chem. Theory Comput 8, 4672 (2012). [DOI] [PubMed] [Google Scholar]
- 11.Jo S, Jiang W, Lee HS, Roux B, and Im W, J. Chem. Inf. Model 53, 267 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Gumbart JC, Roux B, and Chipot C, J. Chem. Theory Comput 9, 794 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Lin Y-L, Meng Y, Jiang W, and Roux B, Proc. Natl. Acad. Sci. USA 110, 1664 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Lin Y-L and Roux B, J. Am. Chem. Soc 135, 14741 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Gumbart JC, Roux B, and Chipot C, J. Chem. Theory Comput 9, 3789 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Fu H, Cai W, Hénin J, Roux B, and Chipot C, J. Chem. Theory Comput 13, 5173 (2017). [DOI] [PubMed] [Google Scholar]
- 17.Jiang W, Thirman J, Jo S, and Roux B, J. Phys. Chem. B 122, 9435 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Loeffler HH, Bosisio S, Duarte Ramos Matos G, Suh D, Roux B, Mobley DL, and Michel J, J. Chem. Theory Comput 14, 5567 (2018). [DOI] [PubMed] [Google Scholar]
- 19.Michel J and Essex JW, J. Med. Chem 51, 6654 (2008). [DOI] [PubMed] [Google Scholar]
- 20.Boyce SE, Mobley DL, Rocklin GJ, Graves AP, Dill KA, and Shoichet BK, J. Mol. Biol 394, 747 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Wang L, Berne BJ, and Friesner RA, Proc. Natl. Acad. Sci. USA 109, 1937 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Zhu S, Travis SM, and Elcock AH, J. Chem. Theory Comput 9, 3151 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Wang L, Wu Y, Deng Y, Kim B, Pierce L, Krilov G, Lupyan D, Robinson S, Dahlgren MK, Greenwood J, et al. , J. Am. Chem. Soc 137, 2695 (2015). [DOI] [PubMed] [Google Scholar]
- 24.Aldeghi M, Heifetz A, Bodkin MJ, Knapp S, and Biggin PC, Chem. Sci 7, 207 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Aldeghi M, Heifetz A, Bodkin MJ, Knapp S, and Biggin PC, J. Am. Chem. Soc 139, 946 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Wan S, Bhati AP, Zasada SJ, Wall I, Green D, Bamborough P, and Coveney PV, J. Chem. Theory Comput 13, 784 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Menzer W, Li C, Sun W, Xie B, and Minh DDL, J. Chem. Theory Comput 14, 6035 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Wang W and Kollman PA, J. Mol. Biol 303, 567 (2000). [DOI] [PubMed] [Google Scholar]
- 29.Swanson JMJ, Henchman RH, and McCammon JA, Biophys. J 86, 67 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Genheden S and Ryde U, Expert Opin. Drug Discovery 10, 449 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Wood RH, Muhlbauer WCF, Thompson PT, Mühlbauer WC, and Thompson PT, J. Phys. Chem 95, 6670 (1991). [Google Scholar]
- 32.Sadiq SK, Wright DW, Kenway OA, and Coveney PV, J. Chem. Inf. Model 50, 890 (2010). [DOI] [PubMed] [Google Scholar]
- 33.Wan S, Knapp B, Wright DW, Deane CM, and Coveney PV, J. Chem. Theory Comput 11, 3346 (2015). [DOI] [PubMed] [Google Scholar]
- 34.Wright DW, Wan S, Meyer C, van Vlijmen H, Tresadern G, and Coveney PV, Sci. Rep 9, 6017 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Chang C-E, Chen W, and Gilson MK, J. Chem. Theory Comput 1, 1017 (2005). [DOI] [PubMed] [Google Scholar]
- 36.Minh DDL, Bui JM, Chang C.-e. A., Jain T, Swanson JMJ, and McCammon JA, Biophys. J 89, L25 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Weis A, Katebzadeh K, Soderhjelm P, Nilsson I, Ryde U, Söderhjelm P, Nilsson I, Ryde U, Soderhjelm P, Nilsson I, et al. , J. Med. Chem 49, 6596 (2006). [DOI] [PubMed] [Google Scholar]
- 38.Kongsted J and Ryde U, J. Comput.-Aided Mol. Des 23, 63 (2009). [DOI] [PubMed] [Google Scholar]
- 39.Hou T, Wang J, Li Y, and Wang W, J. Chem. Inf. Model 51, 69 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lindström A, Edvinsson L, Johansson A, Andersson CD, Andersson IE, Raubacher F, Linusson A, and Lindstrom A, J. Chem. Inf. Model 51, 267 (2011). [DOI] [PubMed] [Google Scholar]
- 41.Zhang X, Perez-Sanchez H, and Lightstone FC, Curr. Top. Med. Chem 17, 1631 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Duan L, Liu X, and Zhang JZH, J. Am. Chem. Soc 138, 5722 (2016). [DOI] [PubMed] [Google Scholar]
- 43.Zwanzig R, J. Chem. Phys 22, 1420 (1954). [Google Scholar]
- 44.Raval A, Piana S, Eastwood MP, Dror RO, and Shaw DE, Proteins: Struct., Funct., Bioinf 80, 2071 (2012). [DOI] [PubMed] [Google Scholar]
- 45.Xue Y and Skrynnikov NR, Protein Sci. 23, 488 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Wall ME, Calabró G, Bayly CI, Mobley DL, and Warren GL, J. Am. Chem. Soc 141, 4711 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Woo HJ and Roux B, Proc. Natl. Acad. Sci. USA 102, 6825 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Chen P-C and Kuyucak S, Biophys. J 100, 2466 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Velez-Vega C and Gilson MK, J. Comput. Chem 34, 2360 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Perthold JW and Oostenbrink C, J. Chem. Theory Comput 13, 5697 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Boresch S, Tettinger F, Leitgeb M, and Karplus M, J. Phys. Chem. B 107, 9535 (2003). [Google Scholar]
- 52.Hamelberg D and McCammon JA, J. Am. Chem. Soc 126, 7683 (2004). [DOI] [PubMed] [Google Scholar]
- 53.Mobley DL, Chodera JD, and Dill KA, J. Chem. Phys 125, 84902 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Singh N and Warshel A, Proteins: Struct., Funct., Bioinf 78, 1724 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Filippakopoulos P, Qi J, Picaud S, Shen Y, Smith WB, Fedorov O, Morse EM, Keates T, Hickman TT, Felletar I, et al. , Nature 468, 1067 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Buch I, Giorgino T, and De Fabritiis G, Proc. Natl. Acad. Sci. USA 108, 10184 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Bahar I, Atilgan AR, and Erman B, Fold. Des 2, 173 (1997). [DOI] [PubMed] [Google Scholar]
- 58.Salomon-Ferrer R, Case DA, and Walker RC, WIREs Comput. Mol. Sci 3, 198 (2013). [Google Scholar]
- 59.Søndergaard CR, Olsson MHM, Rostkowski M, and Jensen JH, J. Chem. Theory Comput 7, (2011). [DOI] [PubMed] [Google Scholar]
- 60.Olsson MHM, Søndergaard CR, Rostkowski M, and Jensen JH, J. Chem. Theory Comput 7, 525 (2011). [DOI] [PubMed] [Google Scholar]
- 61.Dolinsky TJ, Nielsen JE, McCammon JA, and Baker NA, Nucleic Acids Res. 32, 665 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Wang J, Wolf RM, Caldwell JW, Kollman PA, and Case DA, J. Comput. Chem 25, 1157 (2004). [DOI] [PubMed] [Google Scholar]
- 63.Jakalian A, Bush BL, Jack DB, and Bayly CI, J. Comput. Chem 21, 132 (2000). [DOI] [PubMed] [Google Scholar]
- 64.Jakalian A, Jack DB, and Bayly CI, J. Comput. Chem 23, 1623 (2002). [DOI] [PubMed] [Google Scholar]
- 65.Onufriev A, Bashford D, and Case DA, Proteins: Struct., Funct., Bioinf 55, 383 (2004). [DOI] [PubMed] [Google Scholar]
- 66.Massova I and Kollman PA, Perspect. Drug Discov 18, 113 (2000). [Google Scholar]
- 67.Eastman P and Pande VS, Comput. Sci. Eng 12, 34 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Eastman P, Swails J, Chodera JD, Mcgibbon RT, Zhao Y, Beauchamp KA, Wang L.-p., Simmonett AC, Harrigan MP, Stern CD, et al. , PLoS Comput. Biol 13, e1005659 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Chodera JD, Swope WC, Pitera JW, Seok C, and Dill KA, J. Chem. Theory Comput 3, 26 (2007). [DOI] [PubMed] [Google Scholar]
- 70.Shirts MR and Chodera JD, J. Chem. Phys 129, 124105 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Gullingsrud J, CatDCD - Concatenate DCD files (2009).
- 72.McGibbon RT, Beauchamp KA, Harrigan MP, Klein C, Swails JM, Hernández CX, Schwantes CR, Wang L-P, Lane TJ, and Pande VS, Biophys. J 109, 1528 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Bakan A, Meireles LM, and Bahar I, Bioinformatics 27, 1575 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Rodriguez A and Laio A, Science 344, 1492 (2014). [DOI] [PubMed] [Google Scholar]
- 75.Humphrey W, Dalke A, and Schulten K, J. Mol. Graphics 14, 33 (1996). [DOI] [PubMed] [Google Scholar]
- 76.Chodera JD, Chem J. Theory Comput. 12, 1799 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Kendall M, Biometrika 30, 81 (1938). [Google Scholar]
- 78.Stjernschantz E and Oostenbrink C, Biophys. J 98, 2682 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Wang K, Chodera JD, Yang Y, and Shirts MR, J. Comput.-Aided Mol. Des 27, 989 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Kaus JW, Harder E, Lin T, Abel R, McCammon JA, and Wang L, J. Chem. Theory Comput 11, 2670 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Xie B, Nguyen TH, and Minh DDL, J. Chem. Theory Comput 13, 2930 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Minh DDL, J. Comput. Chem accepted (2019). 10.1002/jcc.26036. [DOI] [Google Scholar]
- 83.Vosmeer CR, Kooi DP, Capoferri L, Terpstra MM, Vermeulen NPE, and Geerke DP, J. Mol. Model 22, 31 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Mobley DL, Comput J. -Aided Mol. Des 26, 93 (2012). [DOI] [PubMed] [Google Scholar]
- 85.Dolenc J, Riniker S, Gaspari R, Daura X, and van Gunsteren WF, J. Comput.-Aided Mol. Des 25, 709 (2011). [DOI] [PubMed] [Google Scholar]
- 86.Towns J, Cockerill T, Dahan M, Foster I, Gaither K, Grimshaw A, Hazlewood V, Lathrop S, Lifka D, Peterson GD, et al. , Comput. Sci. Eng 16, 62 (2014). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.