ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img
RETURN TO ISSUEPREVPharmaceutical Model...Pharmaceutical ModelingNEXT

Identification of Optimal Ligand Growth Vectors Using an Alchemical Free-Energy Method

  • Alexander D. Wade
    Alexander D. Wade
    TCM Group, Cavendish Laboratory, University of Cambridge, 19 J J Thomson Avenue, Cambridge CB3 0HE, United Kingdom
  •  and 
  • David J. Huggins*
    David J. Huggins
    Tri-Institutional Therapeutics Discovery Institute, Belfer Research Building, 413 East 69th Street, 16th Floor, Box 300, New York, United States
    Department of Physiology and Biophysics, Weill Cornell Medical College of Cornell University, New York, New York 10065, United States
    *E-mail: [email protected]
Cite this: J. Chem. Inf. Model. 2020, 60, 11, 5580–5594
Publication Date (Web):August 18, 2020
https://doi.org/10.1021/acs.jcim.0c00610

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

  • Open Access

Article Views

2171

Altmetric

-

Citations

1
LEARN ABOUT THESE METRICS
PDF (4 MB)
Supporting Info (1)»

Abstract

In this work, a novel method to rationally design inhibitors with improved steric contacts and enhanced binding free energies is presented. This new method uses alchemical single step perturbation calculations to rapidly optimize the van der Waals interactions of a small molecule in a protein–ligand complex in order to maximize its binding affinity. The results of the optimizer are used to predict beneficial growth vectors on the ligand, and good agreement is found between the predictions from the optimizer and a more rigorous free energy calculation, with a Spearman’s rank order correlation of 0.59. The advantage of the method presented here is the significant speed up of over 10-fold compared to traditional free energy calculations and sublinear scaling with the number of growth vectors assessed. Where experimental data were available, mutations from hydrogen to a methyl group at sites highlighted by the optimizer were calculated with MBAR, and the mean unsigned error between experimental and calculated values of the binding free energy was 0.83 kcal/mol.

SPECIAL ISSUE

This article is part of the Novel Directions in Free Energy Methods and Applications special issue.

Introduction

ARTICLE SECTIONS
Jump To

Free energy methods have become a valuable tool in rational drug design. (1) With the increased computational power from hardware accelerated computer resources, (2−4) particularly GPUs, (5−7) allowing for longer simulation times and the advances in biological force field accuracy, (8−11) it is now practical to make accurate predictions of binding free energy changes. (1,12−18) However, protein–ligand binding free energy is a complex balance of enthalpic and entropic terms in both the bound and unbound systems. (19) Designing molecules to effectively tip this balance is extremely challenging to achieve by eye, and intuition about what changes to the ligand might improve the binding free energy is often incorrect. (19) The task of finding beneficial mutations can be made more systematic by employing methods such as pharmacophore modeling, which have become more accurate with the inclusion of dynamic trajectories. (20) The great advantage of free energy methods, however, is their ability to evaluate the balance of enthalpic and entropic terms quantitatively and provide reliable information for what changes to a ligand might improve the binding affinity experimentally.
A significant part of a drug design campaign is ideation. This is the process of generating ideas which may lead to improved properties, for example, binding affinity (21) or selectivity. (22) Free energy methods can be a useful tool to predict whether an ideated drug does have an improved binding affinity, and free energy methods have been widely applied to study ideated compounds. (21−25) However, the application of free energy methods to idea generation has seen much less attention. While it is possible to exhaustively test a large set of ligands around a starting hit molecule as part of an ideation campaign these searches are not directed, instead relying on trial and error. This wastes a lot of time testing “bad” ideas, similar to the inefficiencies of experimental campaigns. Since free energy methods are computational, there exists flexibility in how to approach the problem of finding ligands with improved properties. (26) Here, we consider a reframing of the problem as a numerical optimization.
The binding free energy that is calculated for a ligand is a function of the parameters of the system, such as atomic partial charge. Tidor et al. used this idea in previous work (27,28) to perform charge optimizations of ligands, that is, finding a set of atomic partial charges which minimize the binding free energy. These optimizations used Poisson–Boltzmann calculations in the bound and unbound ligand systems to calculate an optimal set of charges which improve binding affinity. (29) More recently, the authors of this paper have extended these methods including explicit water models and single step perturbation (SSP). (30) SSP is a type of free energy perturbation (FEP) and exists among a myriad of other free energy methods. (31−36) FEP methods are based on the Zwanzig equation, as shown in eq 1:
(1)
with k as Boltzmann’s constant, T as temperature, E1 and E2 as the potential energies of the system calculated using the Hamiltonian of states 1 and 2, and ⟨...⟩1 as a state average over system 1. It is typical in free-energy calculations (34,35) to sample from many intermediate states between the two end states. This is done to increase the sampling overlap between adjacent states and improve convergence. However, this adds computational time as molecular dynamics (MD) simulations must be performed for every state sampled. If the perturbation between end states remains small enough, then eq 1 can be applied effectively without any intermediate states. Using sampling from only one end state is referred to as a unidirectional transformation, and this is the defining characteristic of the SSP method. Numerous studies have used SSP (37−40) with applications of this method made to relative free energy calculations (30,41−43) and demonstrations made that it can be significantly faster than standard FEP approaches. (30,43,44) While not explored in this work it is worth mentioning that in the literature there exist methods to improve the performance of SSP methods. These involve sampling of an unphysical central reference state which has improved overlap to all end states. (45−49)
Another important free energy method used in this work is the multistate Bennett acceptance ratio (MBAR) estimator. (35) MBAR is a generalization of the Bennett acceptance ratio BAR (34) method to multiple intermediate alchemical states. Both SSP and MBAR are used in this work to calculate relative binding free energies. SSP will be applied to calculations where the end states are relatively close in parameter space and MBAR applied when these end states are further apart. In general, to reduce the size of the perturbation needed to calculate a ligand–protein binding free energy, it is common to use a relative binding free energy (1,24,25) calculation as opposed to an absolute one, (50) where a relative calculation is the change in binding free energy between two different ligands in the same protein. Tools such as replica exchange (51,52) can be used to improve the convergence properties of these calculations, and accuracy with respect to the experiment can be improved using approaches that, for example, consider buried water molecules (53−55) or using multiscale modeling. (56,57)

Method

ARTICLE SECTIONS
Jump To

The goal of this work is to optimize the steric parameters of a ligand in a protein–ligand complex and use the results of this optimization to predict beneficial growth vectors on the ligand more efficiently than existing FEP methods. To this end, experimental data have been collected to retrospectively validate this method from the following systems: androgen receptor (AR), (58,59) renin, (60) menin, (61) thrombin, (62) and SARS PL protease. (63,64) These systems were chosen as they contain data for free energy differences for changes from hydrogens to methyl groups. Two calculations were performed with menin using two different ligands. These are termed menin A and B. Four calculations were performed for the thrombin systems using four different ligands. These are termed thrombin A, B, C, and D. All ligands used in this work are shown in Table 1.
Table 1. Ligands Used in Androgen Receptor, SARS PL Protease, Renin, Menin, and Thrombin Systems

System Setup

Protein structures are taken from the PDB IDs given in Table 2. The common steps in preparation of these structures using Schrödinger’s Preparation Wizard (65) are as follows: replace selenomethionines with methionines, add missing side chains, add all hydrogens, assign tautomers, and assign the protonation state of all ionizable residues. All buffer solvents and ions were removed. Force field parameters and partial charges were assigned from the AMBER ff14SB force field. (10) Since we are using an AMBER force field, the Lorenz–Berholet combination rules will be used where relevant. Parameters for the inhibitors were generated using Antechamber (66) with AMBER GAFF 2 (67) and AM1-BCC. (68) Water was modeled with the SPC/E water model. A salt concentration of 150 mM and any required counterions were added using sodium and chloride ions. In every case, the edge of the solvation box was set to be 10 Å from any atom of the receptor or ligand. The SI contains detailed information on the preparation specific to each protein.
Table 2. PDB IDs for Each System Used in This Work
system PDB ID
androgen receptor 2NW4 (59)
SARS PL Pro 3E9S (64)
renin 3OOT (60)
menin 4OG6 (61)
thrombin 2ZNK (62)

Molecular Dynamics

All simulations were performed with OpenMM 7.3.0 as follows. First, OpenMM’s default minimizer was used to minimize all structures. Equilibration was then performed in the NPT ensemble at 300 K and 1 atm using a Langevin integrator and Monte Carlo barostat for at least 200 ps. MD simulations were performed in the NPT ensemble using a time step of 2 fs; van der Waals interactions were truncated at 10.0 Å. Electrostatics were modeled using the particle mesh Ewald method with a cutoff of 10.0 Å. Snapshots were collected every 5 ps, and all simulations were performed with constrained hydrogens. Since all hydrogens are constrained, we assume that hydrogen bond oscillations contribute negligibly to any calculated free energy changes. This protocol was used throughout this work; please see the Supporting Information (SI) for MD settings not provided here.

Workflow

In this work, several free energy methods were used in different contexts. For clarity, we will now outline these methods for future reference. After models for the protein ligand system were built, we performed an optimization of the van der Waals parameters of an inhibitor in the protein ligand complex. This is done using a gradient descent (GD) algorithm with a line search. In order to perform this optimization, two calculations are needed: one for the objective and one for the gradient. The objective was calculated with MBAR. We therefore name this the MBAR objective, and the gradient was calculated with SSP. We therefore name this the SSP gradient. The details of how these calculations are performed can be found in the Optimization, SSP Gradient, and MBAR Objective sections below. Once this optimization is converged, the output is analyzed to determine growth vectors which should be beneficial. We then calculate the relative binding free energy of adding methyl groups at the locations specified by the optimizer. This calculation is performed using MBAR with Hamiltonian replica exchange. We name these calculations FEP scans, with more details in the FEP Scans section below. Figure 1 illustrates how each of these calculations are combined in the full workflow.

Figure 1

Figure 1. Diagrammatic workflow for calculations performed in this work. SSP gradient method uses an exponential averaging method to calculate free energies. MBAR objective and FEP scan both use the MBAR estimator.

Optimization

Optimizations were performed using the Ligand-Optimiser code, which is available on Github at https://github.com/adw62/Ligand-Optimiser. The optimizations were applied to the σ parameter of the Lennard-Jones (LJ) potential shown in eq 2.
(2)
Here, r is the distance between two atoms, ε is the depth of the potential, and σ is the distance between atoms at which the potential is zero. Note that this method is agnostic to the exact form of the LJ potential used and the exact meaning of σ in that form. The objective function for the optimization performed in this work is defined in eq 3.
(3)
ΔGbindingi) is the binding free energy difference between the bound and unbound states of the ligand and receptor for a ligand with i atoms using σi parameters. ΔGoriginal is the binding free energy for the ligand using its original parameters. This is a constant for each given system.
To perform this optimization, an implementation of the gradient descent algorithm with a line search was used. The line search used a step size of 0.6 nm and a convergence tolerance of 0.15 nm in σ. The objective function in eq 3 can be considered as the difference in binding free energies between two ligands with binding free energy ΔGbindingi) and ΔGoriginal. This difference is a relative binding free energy ΔΔGbinding. This ΔΔGbinding can be constructed equivalently as the difference of ΔΔGbinding in the bound and unbound states or the difference of ΔGmutation in the bound and unbound states, where ΔGmutation is defined as the free energy difference of mutating one ligand into the other. We used the latter approach as is standard best practice in relative binding free-energy calculations; (14) see Figure S13 for a graphical depiction of this cycle.

MBAR Objective

To evaluate the objective, we used an MBAR method to calculate ΔΔGbinding. We term this evaluation of ΔΔGbinding as ΔΔGopt. The objective was calculated by simulating 24 alchemical windows. The two end states were (a) the system with the Hamiltonian using the sigmas for the current optimization step n and (b) the system with the Hamiltonian using the sigmas from optimization step n + 1. The intermediate states are created by linearly interpolating the sigma parameters only. Charge, bond, angle, and torsion parameters do not change. The sigmas for the optimization step n + 1 were determined with the standard GD algorithm using a maximum step size of 0.6 nm. Sampling from all windows was collected, and the potentials evaluated from this sampling were combined using the MBAR estimator giving ΔΔGmutationopt. Combining ΔΔGmutationopt from the bound and unbound states gave ΔΔGopt across all lambda windows. The lambda windows of the free energy calculation were treated as a line search in sigma space, and therefore the lambda window with the minimum in ΔΔGopt was selected and the sigma parameters corresponding to that lambda window became the accepted parameters for the next step of optimization. If the minimum was found to be at the last lambda value, then the line search was extended by another 0.6 nm without recalculating the gradient. Each evaluation of the objective used 6 ns of sampling in AR, SARS PLPro, and menin A systems. Renin, menin B, and the four thrombin systems used 12 ns, as these optimizations were observed taking steps which correspond to larger values of ΔΔGopt. Figure 2 shows an example for how ΔΔGopt varies with sampling.

Figure 2

Figure 2. Convergence for a calculation of the objective in the androgen receptor test case. ΔΔGopt is reported as mean of three replicates with shaded area showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

SSP Gradient

Methods exist in the literature which detail the calculation for the analytic gradient of the potential (69) or ensemble properties (70,71) of a system with respect to some argument of the potential. In this work, however a numerical finite difference will be used to calculate the gradient. The gradient of the objective with respect to sigma was calculated as shown in eq 4.
(4)
where h is a finite difference of 0.00015 nm. The numerator on the RHS of eq 4 is a ΔΔGbinding and can be calculated using an SSP approach, using sampling only from a central state containing the ligand using σi values. This calculation of the gradient shows the advantage of SSP. As numerous (tens to hundreds) evaluations of the RHS of eq 4 are required, one for each dimension in σ space, they can all be calculated from the sampling of one central state. The number of dimensions in σ space comes from the number of atoms in the molecule. See Figure 3 where each named hydrogen on the molecule is optimized. Figure 3 shows the ligand from the androgen receptor test case where there are 10 named hydrogens and so a 10-dimensional σ space.

Figure 3

Figure 3. 3D structure for the androgen receptor ligand with all the optimized hydrogens explicitly labeled.

The finite difference in eq 4 is between molecules that are extremely similar, differing only by 0.00015 nm in one atom’s σ. There is therefore likely to be a large sampling overlap between these states allowing SSP to be applied. Throughout this work, the gradient was calculated from 5 ns of sampling in the central reference state. Figure 4 shows an example for how this gradient varies with sampling for the androgen receptor ligand.

Figure 4

Figure 4. Convergence for a calculation of the gradient in the androgen receptor test case. ΔΔGgrag/h is reported as the mean of three replicates with the shaded area showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees. Atom names in the legend can be seen in Figure 3.

Optimization Validation

To test the reproducibility of the optimization, we examined how the optimized values for the sigma of each hydrogen varied across repeats. Three repeats of an optimization for the androgen receptor were made, and the averaged set of optimized sigmas are presented in Table 3. We can see from Table 3 that the variance in the optimized σ is small relative to the difference between original and optimized σ values.
Table 3. Original and Optimized Sigmas for Every Atom in the Androgen Receptor Ligand Alongside the Variancea
hydrogen name original σ [nm] average optimized σ [nm] variance optimized σ [nm2]
H5 0.263 0.154 0.006
H7 0.263 0.301 0.000
H15 0.242 0.032 0.001
H17 0.242 0.245 0.000
H191 0.260 0.285 0.001
H192 0.260 0.427 0.002
H221 0.242 0.233 0.000
H222 0.242 0.296 0.000
H3 0.263 0.381 0.002
H2 0.263 0.264 0.001
a

Averages and variances are taken from three repeats. All hydrogen names are given in Figure 3.

To ensure each optimization is converged, the cumulative ΔΔGopt is plotted over optimization steps. These convergence data are presented in Figure 5, where it can be seen that the optimizations in each case are well converged.

Figure 5

Figure 5. Cumulative ΔΔGopt calculated by summing the MBAR objective for each step of the optimization plotted against optimization step for every system.

Additional calculations were made to verify the value of the cumulative ΔΔGopt. This involved calculating the relative binding free energy between the original and optimized parameters of the inhibitor with MBAR. The mean absolute error between the values of ΔΔGopt and the verification calculation was 1.42 kcal/mol. Details and results of this calculation can be seen in the SI (Figures S1–S3 and Table S2).

FEP Scans

The results of the optimization were analyzed, and the best growth vectors on the ligand were determined. The best growth vectors were tested using the following protocol named FEP scans which used the MBAR method. The binding free energy associated with these FEP scans will be termed ΔΔGscan. ΔΔGscan values were calculated using OpenMMTools (72) to collect sampling using Hamiltonian replica exchange. Free energy changes were calculated from this sampling using the MBAR estimator. Each Hamiltonian in the replica exchange was an intermediate state in the alchemical transformation using a hybrid topology. For these intermediate states, the charge, van der Waals, bond, angle, and torsion parameters are all interpolated between the end states. A soft core potential is used to interpolate the LJ potential for these FEP scan calculations. In this work, we focus on transforming hydrogens to methyls. This makes the exploration of the chemical space tractable. Transformations to methyl at all possible sites allow for a thorough validation of the optimizations because a more complete comparison can be made between the ranking of growth vectors by the optimization and the FEP scans. Exhaustive testing of all methyls is done here for the purpose of validation. If this method is applied in a drug discovery effort, only the top growth vectors ranked highly by the optimizer would need to be tested using FEP scans.
To make a hydrogen disappear and a methyl appear, each FEP scan used 33 ns of sampling split across 22 alchemical windows, except renin, which used 55 ns, as this was observed to require more sampling to converge. Hamiltonian swapping was performed every 5 ps. All calculations were performed in triplicate. In Figure 6, we show the convergence for making all possible hydrogen to methyl mutations on the androgen receptor ligand. All remaining convergence graphs for the FEP scan calculations can be seen in Figures S4–S11.

Figure 6

Figure 6. ΔΔGscan for each methylation in the androgen receptor system as the amount of sampling is increased. ΔΔGscan are reported as mean of three replicates with shaded area showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions. All hydrogen names are taken from Figure 3.

When trying to calculate all hydrogen to methyl mutations, we encountered some methyls which introduced numerical instability into the simulation. The causes of these instabilities in the SARS PLPro, menin, and thrombin test cases were in very close contact between an existing part of the ligand and the added methyl. In the androgen receptor test case, the cause was close contact between the protein and the added methyl. We show the androgen receptors case in Figure S12. Any ΔΔGscan that could not be calculated due to numerical instability is given an NA value in the results.

Summary of Methods

ARTICLE SECTIONS
Jump To

To summarize, in this work, an optimization of the sigma parameters of several ligands was performed. The objective of this optimization was to find a set of sigmas to minimize the relative binding free energy of the ligand to a protein. This objective, ΔΔGopt, was evaluated using FEP calculations and the MBAR estimator. To calculate the gradient, ΔΔGgrad, SSP calculations were used. The results of the optimizer are used to predict beneficial growth vectors on the ligand. To validate these predicted growths, hydrogen to methyl mutations were computed, ΔΔGscan, using FEP and the MBAR estimator. For the readers’ reference, Table 4 contains all the abbreviated terms used and a brief description of the calculation.
Table 4. Abbreviated Terms Used to Reference Various Relative Binding Free Energies Calculated in This Work with a Brief Description of the Calculation for the Readers Referencea
free energy value description
ΔΔGopt the value for the objective, calculated between the original and optimized sigmas with MBAR
ΔΔGgrad the value for the gradient, calculated between many highly related ligands using SSP
ΔΔGscan the value for one or many hydrogen to methyl mutations calculated with MBAR
ΔΔGexp the value for one or many hydrogen to methyl mutations determined experimentally
a

Both ΔΔGopt and ΔΔGgrad are calculated with fixed C–H bond lengths due to the difficulty in implementing alchemical changes to OpenMM constraints. However, ΔΔGscan values are calculated with correct bond lengths at both end states.

Results

ARTICLE SECTIONS
Jump To

The results from the optimizations performed in this work will be considered one system at a time. First, looking at the androgen receptor, each optimization produces a set of optimized σ’s, which minimizes the binding free energy. We show this result with figures made such that any optimized hydrogens are sized in proportion to their calculated Δσ, where Δσ is the difference between the atoms optimized and original σ. One reason we create these figures is to allow the continuous values of σ in the optimization to be converted by the eye into the discrete values of σ associated with adding any atom(s). In the following calculations, we denote symmetry related positions with an obelisk symbol (†). For optimization calculations, no atoms are considered symmetric. This is because the optimizer assigns different sigmas to symmetric atoms, and this breaks any symmetry. In the context of these simulations, symmetric methyl hydrogens rapidly interconvert due to rotation, but symmetric hydrogens on aromatic rings do not. Thus, hydrogens on methyls are considered symmetric for the ΔΔGscan calculations, but hydrogens on aromatic rings are not.
Figure 7 shows the relative size of the optimized hydrogens in the androgen receptor test case. The largest hydrogens are H192 and H3, with H3 as the position chosen in previous experimental work to be methylated. The values of Δσ from the optimization are tabulated alongside calculated values for ΔΔGscan, and available experimental free energies are reported as ΔΔGexp.

Figure 7

Figure 7. Androgen receptor ligand with all optimized hydrogens sized in proportion to their calculated Δσ.

The results in Table 5 show that, for the AR test case, the optimization and FEP scan rank the experimentally verified methylation, H3, second and first, respectively. The overall agreement in the ranking between computational methods can be calculated with the Spearman’s Rank-Order Correlation, ρ. Here, we calculated ρ between the ranking from the optimization and FEP Scan (any hydrogens without calculated value for ΔΔGscan are ranked last), and for the AR test case, ρ was calculated as 0.7. The agreement between the experimental ΔΔGexp and computational ΔΔGscan is not good in this case, differing by more than 1 kcal/mol. Looking at the outlying data in Table 5, it can be seen that while the optimization ranks H2 as not beneficial, the FEP Scan calculated it to be a beneficial position for a methyl. We speculate that this is because during the optimization all σ values are changed simultaneously in the same system. This means that if both H2 and H3 grow simultaneously, they will see each other with increased radius in the simulation. It may be possible that some growths which are close in proximity can interfere with each other such that only one position grows to maximize the binding affinity. This effect would not be seen for the FEP Scan calculation as each methylation is a separate calculation and as such the methylation at H2 and H3 do not need to be accommodated simultaneously.
Table 5. Comparison of Δσ,ΔΔGscan for Mutating a Hydrogen to Methyl for the Androgen Receptor Test Case Alongside Experimental Values ΔΔGexpa
hydrogen name Δσ [nm] ΔΔGscan [kcal/mol] ΔΔGexp [kcal/mol]
H192 0.148 –2.32 [−3.14, −1.5]  
H3 0.122 –3.78 [−3.91, −3.64] –0.83
H191 0.041 –1.0 [−1.29, −0.7]  
H222 0.041 –1.8 [−2.15, −1.45]  
H7 0.038 –0.79 [−1.6, 0.02]  
H17 –0.004 1.45 [0.56, 2.34]  
H221 –0.004 –0.95 [−1.09, −0.82]  
H2 –0.015 –2.28 [−2.5, −2.05]  
H5 –0.187 –0.62 [−2.95, 1.71]  
H15 –0.237 NA  
a

ΔΔGscan values are reported as the mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

Figure 8 provides the result from the optimizer for the SARS PLPro system and shows that three growth vectors have been highlighted. Two of these growths, H4 and H9, are adjacent to each other pointing in approximately the same direction; the other, H10, is separately located on the ortho position of a phenyl. It is this ortho position that was the experimentally chosen position in previous work.

Figure 8

Figure 8. SARS PLPro ligand with all optimized hydrogens sized in proportion to their calculated Δσ.

Looking more closely at the SARS PLPro result in Table 6, it can be seen that the experimentally chosen hydrogen, H10, is ranked second by the optimizer and first by the FEP scan. H4, the position most favored by the optimizer, is confirmed to be a beneficial position for methylation by the FEP scan with a ΔΔGscan of −0.45 [−1.09, 0.19]. For the SARS PLPro test case, the correlation in the ranking of growth vectors by the optimizer and FEP scan is good with ρ calculated to be 0.8. The experimental data which exist for the SARS PLPro systems is for ligands with methyls at positions H10/H13, H17/H19, and H21, but no reference data exist for the ligand with a hydrogen at all of these sites. To compare to the experiment, we therefore use the ΔΔGscan values in Table 6 to calculate the relative binding free energy change for permuting the methyl on sites H10/H13, H17/H19, and H21 and present this result in Table 7.
Table 6. Comparison of Δσ, ΔΔGscan for Mutating a Hydrogen to Methyl for the SARS PLPro Test Casea
hydrogen name Δσ [nm] ΔΔGscan [kcal/mol] experimental rank
H4 0.270 –0.45 [−1.09, 0.19]  
H10 0.115 –1.70 [−2.35, −1.05] 1
H9 0.115 –0.73 [−0.94, −0.53]  
H18 0.047 –0.16 [−0.46, 0.13]  
H19 0.028 –0.26 [−0.65, 0.14] 2
H22 0.013 –0.34 [−0.75, 0.07]  
H21 0.002 –0.10 [−0.27, 0.08] 3
H12 –0.004 –0.18 [−0.67, 0.31]  
H20 –0.007 –0.58 [−0.79, −0.36]  
H13 –0.011 –1.14 [−1.17, −1.11]  
H01 –0.020 0.34 [−0.02, 0.7]†  
H013 –0.021 0.34 [−0.02, 0.7]†  
H012 –0.023 0.34 [−0.02, 0.7]†  
H14 –0.080 NA  
H6 –0.103 NA  
H17 –0.106 1.05 [0.89, 1.21]  
a

ΔΔGscan are reported as the mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions. Symmetry related positions are indicated by a † symbol.

Table 7. ΔΔGscan for Mutating a Methyl from One Position to Another on the Liganda
methyl mutation ΔΔGscan [kcal/mol] ΔΔGexp [kcal/mol]
H10 to H19 1.44 [0.68, 2.20] 0.32
H10 to H21 1.60 [0.93, 2.27] 0.72
H19 to H21 0.16 [−0.27, 0.59] 0.40
a

Experimental values ΔΔGexp for methyl mutations in the SARS PLPro test case. ΔΔGscan are reported as mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

Table 7 shows that the ΔΔGscan for moving the methyl from H19 to H21 is in good agreement with the experiment, within 1 kcal/mol. For the H10 to H19 and H10 to H21 transformations, disagreement with the experiment is greater than 1 kcal/mol. However, the ranking in Table 6, is much more accurate with the experimentally preferred methylation ranked first and second by the FEP scan and optimization, respectively.
The data for the renin system are detailed in Figure 9 and Table 8; from these data it can be seen that there is one clear growth vector highlighted by the optimizer H52. This is not in the direction chosen experimentally (H50/H54). There is, however, agreement between the optimizer and the FEP scan, which also ranked H52 first with a ΔΔGscan of −1.56 [−1.85, −1.27]. This is an example where the ranking between the computational and experimental methods is not in agreement, but the rankings between computational methods is in agreement.

Figure 9

Figure 9. Renin ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

Table 8. Comparison for Δσ, ΔΔGscan for Mutating a Hydrogen to Methyl and Any Experimental Values ΔΔGexp for the Renin Test Casea
hydrogen name Δσ [nm] ΔΔGscan [kcal/mol] ΔΔGexp [kcal/mol]
H52 0.256 –1.56 [−1.85, −1.27]  
H38 0.158 –1.05 [−1.28, −0.82]  
H221 0.158 1.69 [1.01, 2.36]  
H39 0.088 –0.79 [−1.46, −0.13]  
H31 0.084 –0.36 [−0.76, 0.04]  
H6 0.080 –0.31 [−0.54, −0.08]  
H54 0.080 –0.28 [−0.49, −0.07] –1.75
H222 0.077 0.67 [−0.56, 1.91]  
H50 0.069 0.68 [0.26, 1.11]  
H37 0.052 –0.94 [−1.14, −0.74]  
H36 0.037 0.46 [0.07, 0.86]  
H99 0.017 0.17 [−0.54, 0.87]  
H40 –0.006 0.34 [−0.92, 1.59]  
H191 –0.060 6.38 [3.89, 8.88]  
H202 –0.065 1.65 [−0.33, 3.63]  
H192 –0.067 5.8 [5.1, 6.51]  
H201 –0.070 9.14 [8.35, 9.93]  
H1 –0.089 1.42 [1.15, 1.7]  
H53 –0.099 0.46 [0.08, 0.85]  
H231 –0.102 1.78 [−3.86, 7.42]  
H232 –0.176 0.7 [−2.3, 3.7]  
a

ΔΔGscan are reported as the mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

The result for H221 in Table 8 is outlying, and there is not good agreement between the optimizer and the FEP scan. We speculate that this disagreement is caused by the disruption of a favorable charged interaction between the protein and the amine group adjacent to H221. These interactions may be disrupted by the methyl group added in the FEP scan but not by the small change in sigma seen during the optimization. Overall, the correlation in the ranking between the computational methods in the renin system is high with a calculated ρ of 0.7. The difference between the calculated and experimental binding affinity is between 1 and 2 kcal/mol.
The result for the menin A system can be seen in Figure 10 and Table 9. These results show that three growth vectors are highlighted: HAL, HAU2, and HAO2. Our method agrees with the experimental result that HBF is not a position which can beneficially accommodate a methyl group; this can be seen because the Δσ for HBF is not a large positive.

Figure 10

Figure 10. Menin A ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

Table 9. Comparison for Δσ, ΔΔGscan for Mutating a Hydrogen to Methyl and Any Experimental Values ΔΔGexp for the Menin A Test Casea
hydrogen name Δσ [nm] ΔΔGscan [kcal/mol] ΔΔGexp [kcal/mol]
HAU2 0.167 0.46 [−0.27, 1.2]  
HAL 0.111 –1.57 [−2.24, −0.9]  
HAO2 0.094 –0.49 [−1.19, 0.2]  
HAV1 0.059 0.76 [−0.4, 1.92]  
HAV2 0.051 0.78 [−0.66, 2.23]  
HAW1 0.051 0.35 [−0.96, 1.65]  
HAP1 0.033 1.62 [0.69, 2.54]  
HAU1 0.021 1.95 [1.07, 2.84]  
HAS1 0.009 –0.05 [−0.61, 0.52]  
HAJ 0.006 0.72 [−0.58, 2.01]  
HAP2 0.003 0.75 [0.3, 1.2]  
HAS2 0.000 0.28 [−0.58, 1.14]  
HBF –0.001 –0.22 [−0.78, 0.34] 0.14
HAG –0.001 –0.09 [−0.22, 0.05]  
HAF –0.003 0.16 [−0.55, 0.86]  
HAH –0.010 1.42 [1.0, 1.84]  
H3 –0.017 0.83 [0.28, 1.38]  
HAW2 –0.020 0.32 [−0.22, 0.85]  
HAT2 –0.062 0.03 [−0.35, 0.41]  
HAT1 –0.083 0.29 [0.23, 0.35]  
HAI –0.112 NA  
HAK –0.130 0.99 [−0.19, 2.17]  
HAO1 –0.158 1.92 [0.72, 3.12]  
HBD –0.196 NA  
HAE –0.207 0.29 [−0.78, 1.36]  
a

ΔΔGscan are reported as mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

It is worth noting that the HBF site is beneficial for groups larger than methyl such as isopropyl or cyclohexyl with a ΔΔGexp of −2.41 or −2.94 kcal/mol, respectively. In its current form the optimization cannot provide information that this is a beneficial spot for larger mutations. The highlighted positions, HAL and HAO2, are in good agreement between the optimization and FEP Scan, ranked second and third by the optimizer and first and second by the FEP Scan. HAU2 is not agreed between the two computational methods. The overall rank correlation σ for the menin A system was 0.3. The experimental value for adding a methyl at HBF is in reasonable agreement with the experiment, with an experimental value of 0.14 kcal/mol compared to the ΔΔGscan value of −0.22 [−0.78, 0.34]
As mentioned, the site at HBF in the test case menin A will beneficially accommodate mutations larger than a methyl and this will now be investigated with system menin B. The menin B ligand is the same as a menin A with a methyl added at HBF. When ligand menin B is optimized, the newly added hydrogens H55 and H54 are now found to be the most beneficial by the optimizer. This can be seen in Figure 11 where H55 and H44 are ranked first and second, see Table S3 for all values for optimized σ. The experimental data for growths of ligand menin B correspond to adding methyls to H54 and H55 simultaneously. Therefore, the ΔΔGscan values calculated are now also for adding two methyls, one at H54 and one at H55. These ΔΔGscan calculations for two methyls give a value of −3.11 [−4.02, −2.2] kcal/mol, which agrees well with the experimental value −2.55 kcal/mol.

Figure 11

Figure 11. Menin B ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

The final test case to consider is thrombin. In this test case, several mutations are strung together and the optimization method applied iteratively to build a more tightly binding inhibitor. We perform this test case as we imagine this method might be applied in a drug discovery setting. This means that not all methylations will be tested by computing ΔΔGscan, in every round of optimization, only those which would contribute new information to the study. The goal here was to use information from the optimizer to move from thrombin A shown in Table 1, which has an experimental binding free energy of −7.1 kcal/mol, to a ligand with improved binding free energy.
Considering the result of the thrombin A optimization in Figure 12 and Table 10, it can be seen that H44 is ranked first by the optimization, and H27/H45/H48 are ranked first by the FEP Scan. The ranking of H17 is not in good agreement between the optimizer and the FEP scan. This may be a similar effect to that seen for the outlier in the renin test case, where perhaps in the FEP scan the added methyl is disrupting the charged interaction between the protein and the amine groups on the ligand near H17. Another possible cause for the disagreement might be explained by looking at H22. In the original thrombin A ligand, H17 is symmetric to H22, and the optimizer and FEP scan are in agreement that H22 is very unfavorable. Therefore, it may also be possible that in the FEP scan simulation H17 and H22 are adequately interconverting to converge at the solution that H17/H22 is an unfavorable position, but this interconversion is not occurring sufficiently for the optimization simulations, which are shorter. Overall, the ρ calculated between the ranking from the optimization and FEP Scan for the thrombin A system was 0.7. On the basis of the results of Figure 12 and Table 10, either H44 or H27/H45/H48 could be chosen for methylation, and here we show the result for methylating H44. The thrombin A ligand (Table 1) with a methyl added at H44 is the ligand thrombin B in Table 1, and so ligand thrombin B was then considered for an additional round of optimization. With the retrospective knowledge that the H44 is a beneficial location for an amine group the next round of optimization is performed for the partial charges of ligand thrombin B. The charge optimization used was identical to the steric optimization with the exception that the parameters of the optimization were the partial charges of the atoms instead of the atoms σ. The methodology to perform optimization has been explored in previous work. (30)

Figure 12

Figure 12. Thrombin A ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

Table 10. Comparison for Δσ, ΔΔGscan for Mutating a Hydrogen to Methyl and Any Experimental Values ΔΔGexp in the Thrombin A Test Casea
hydrogen name Δσ [nm] ΔΔGscan [kcal/mol] ΔΔGexp [kcal/mol]
H44 0.232 –0.57 [−0.89, −0.24]  
H32 0.056 –0.07 [−0.36, 0.21]  
H27 0.046 –1.75 [−2.66, −0.83]†  
H25 0.045 –0.24 [−0.49, 0.02] –0.24
H48 0.040 –1.75 [−2.66, −0.83]†  
H33 0.033 0.34 [−0.65, 1.33]  
H17 0.028 5.11 [3.78, 6.44]  
H45 0.018 –1.75 [−2.66, −0.83]†  
H16 0.012 –0.8 [−1.05, −0.55]  
H24 0.003 NA  
H26 –0.001 –0.76 [−1.21, −0.31] –0.24
H31 –0.010 1.08 [0.79, 1.37]  
H29 –0.032 2.7 [1.79, 3.6]  
H23 –0.035 1.1 [0.86, 1.34]  
H36 –0.075 5.45 [5.03, 5.87]  
H34 –0.086 1.56 [1.47, 1.66]  
H28 –0.136 2.94 [1.82, 4.06]  
H30 –0.144 2.46 [1.49, 3.44]  
H37 –0.153 3.56 [−0.66, 7.78]  
H22 –0.333 6.09 [3.53, 8.65]  
a

ΔΔGscan are reported as mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

Figure 13 details the result of the charge optimization and shows that hydrogens H49 and H50 can be made more positive to improve the binding free energy. The total change in charge for H49, H50, and H51 made by the optimizer was +0.3e, with all changes in partial charge shown in Table S4. Adding a charged amine group at this position is known experimentally to be beneficial, giving a change in binding affinity of −2.06 kcal/mol. We therefore chose to add an amide group to ligand thrombin B (Table 1), which gave ligand thrombin C (Table 1). We then continued to optimize with another round for the thrombin C ligand. We are now once again optimizing the sigma parameters of ligand thrombin C, and the results of this optimization are presented in Figure 14 and Table 11.

Figure 13

Figure 13. Thrombin B ligand with all optimized hydrogens colored in proportion to their calculated Δq. Red is more positive and blue more negative. See all named hydrogens in Table S1.

Figure 14

Figure 14. Thrombin C ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

Table 11. Comparison for Δσ, ΔΔGscan for Mutating a Hydrogen to Methyl and Any Experimental Values ΔΔGexp in the Thrombin C Test Casea
hydrogen name Δσ [nm] ΔΔGscan [kcal/mol] ΔΔGexp [kcal/mol]
H45 0.305 –1.68 [−2.27, −1.09]†  
H16 0.228 –0.54 [−0.87, −0.22]  
H26 0.196 –0.52 [−1.11, 0.07] –0.60
H24 0.138    
H32 0.079    
H33 0.038    
H23 0.028    
H27 0.007 –1.68 [−2.27, −1.09]†  
H48 –0.001 –1.68 [−2.27, −1.09]†  
H31 –0.003    
H29 –0.027    
H25 –0.029 –0.29 [−0.93, 0.34] –0.60
H34 –0.040    
H37 –0.098    
H28 –0.099    
H22 –0.109    
H17 –0.153    
H30 –0.159    
H36 –0.180    
a

ΔΔGscan are reported as mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

In Table 11, only a subset of all possible ΔΔGscan values is calculated; the calculated values of ΔΔGscan are chosen in the areas where growth was made to the ligand experimentally. Looking at Figure 14 and Table 11, three growth vectors are highlighted: H45, H16, and H26, which are all confirmed to be beneficial by the FEP scan. The ρ calculated between the ranking from the optimization and FEP Scan for the thrombin C system was 0.6. The growth vector ranked first by the optimizer and the methylation scan was H45 with a ΔΔGscan of −1.68 [−2.27, −1.09] kcal/mol. Since the optimizer and FEP scan both rank H45, first it was selected to be mutated to a methyl. Adding a methyl at H45 of ligand thrombin C (Table 1) gives ligand thrombin D (Table 1); we therefore performed a final round of optimization on ligand thrombin D.
Again, only a subset of all possible methylations are calculated in Table 12. With the calculated values of ΔΔGscan chosen in the area, growths were made to the ligand experimentally. For the thrombin D test case, Figure 15 and Table 12 show that the optimizer ranks H16 and H26 as the first and second best growth vectors. The FEP scan ranks H26 as the best growth vector. The ρ calculated between the ranking from the optimization and FEP Scan for the thrombin C system was 0.4. On the basis of the information from the optimizer and FEP scan, H26 was selected to be methylated. To compare to the experiment, the ΔΔGscan’s from thrombin C and thrombin D were combined. The ΔΔGscan for H45 from thrombin C was combined with the ΔΔGscan for methylations in thrombin D, which gave the free energy change for adding a methyl at both sites simultaneously; these calculations are presented in Table 13.

Figure 15

Figure 15. Thrombin D ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

Table 12. Comparison for Δσ, ΔΔGscan for Mutating a Hydrogen to Methyl in the Thrombin D Test Casea
hydrogen name Δσ [nm] ΔΔGscan [kcal/mol]
H16 0.177  
H26 0.145 –1.35 [−1.76, −0.95]
H24 0.098 1.57 [0.26, 2.87]
H53 0.089 –1.02 [−1.32, −0.72]†
H52 0.075 –1.02 [−1.32, −0.72]†
H23 0.019  
H48 0.015 –0.64 [−1.13, −0.14]
H31 0.012  
H54 0.008 –1.02 [−1.32, −0.72]†
H27 0.008 –0.67 [−0.74, −0.6]
H33 0.005  
H25 0.000 –0.27 [−1.13, 0.6]
H32 –0.004  
H29 –0.038  
H34 –0.065  
H22 –0.079  
H28 –0.126  
H17 –0.133  
H37 –0.156  
H30 –0.156  
H36 –0.164  
a

ΔΔGscan are reported as mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

Table 13. Comparison for Δσ, ΔΔGscan for Mutating Thrombin C H45 to a Methyl and One Other Named Hydrogen in Thrombin D to a Methyl and Any Experimental Values ΔΔGexp for the Thrombin D Test Casea
hydrogen name ΔΔGscan [kcal/mol] ΔΔGexp [kcal/mol]
H26 –3.03 [−3.74, −2.32]  
H52 –2.7 [−3.36, −2.04]†  
H53 –2.7 [−3.36, −2.04]†  
H54 –2.7 [−3.36, −2.04]†  
H27 –2.35 [−2.94, −1.76] –1.20
H48 –2.32 [−3.09, −1.55] –1.20
H25 –1.95 [−3.0, −0.9]  
H24 –0.11 [−1.54, 1.32]  
a

ΔΔGscan are reported as mean of three replicates with bracketed values showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

H27 and H48 are the positions methylated in the experimental work, and it can be seen in Table 13 that computationally these positions are also beneficial to adding a methyl. However, H27 and H48 are not ranked highly; instead H26 was ranked first. Comparing the ΔΔGscan for H27 or H48 to the experimental values gives reasonable agreement with the experiment, within 1 kcal/mol for both H27 and H48.
Finally, when all the mutations selected in the thrombin optimization iterations are combined, the final ligand is presented in Figure 16. Figure 16 additionally presents an experimentally optimized ligand also originating from the ligand thrombin A in Table 1. It can be seen in Figure 16 that our optimized ligand is very similar to a ligand determined experimentally, differing only in the placement of one methyl group. The binding free energy of the experimental ligand is −10.4 kcal/mol, (62) we can use this value in combination with our calculation for removing a methyl at H27 and adding a methyl at H26 (Table 12) to estimate the binding free energy of our computationally optimized ligand. The binding free energy of our optimized ligand is calculated to be −11.1 [−10.7, −11.5] significantly better than the starting ligand (thrombin A), which had an experimental binding free energy of −7.1 kcal/mol.

Figure 16

Figure 16. Ligand with improved binding free energy found experimentally and final computationally optimized ligand built by choosing perturbations highlighted by our optimization methodology.

Conclusion

ARTICLE SECTIONS
Jump To

In this work, a novel method to optimize the van der Waals interactions of small molecule inhibitors to maximize the binding affinity to a receptor has been developed. This method combines a rapid single step perturbation calculation with MBAR calculations to calculate the gradient and objective, respectively, allowing optimized inhibitors to be found quickly. This new method was applied to nine inhibitors across five diverse test systems: androgen receptor, SARS PL protease, renin, menin and thrombin. Good agreement was found between the beneficial growth vectors identified by the optimization and the results of full FEP calculations to exhaustively calculate methylations, with a Spearman’s rank order correlation of 0.59. The method developed in this work is found to be approximately 10 times faster than testing all possible growth vectors with FEP. Using an Nvidia P100 GPU, an optimization for the androgen receptor test case (containing around 4000 atoms in the ligand-protein simulation) takes approximately 13 h of wall time. This is compared to 15 h per growth vector, totaling 150 h to test all growth vectors with full FEP. Additionally, the scaling of the optimization compute time with the number of growth vectors is sublinear compared to linear scaling in the full FEP scan case. The sublinear scaling is a result of using single step perturbation theory in the optimization calculations that allowed any number of growth vectors to be assessed with one molecular dynamics simulation. Where experimental data were available mutations highlighted by the optimizer were tested with full FEP, and the mean unsigned error between experimental and calculated values of the binding free energy was 0.83 kcal/mol. We suggest that optimization methods such as this will be useful in a drug discovery setting to identify beneficial growth vectors during the hit-to-lead process, reducing the need for costly trial and error in both computational and experimental campaigns.
This work could be extended by investigating the information that could be extracted by optimizing any other parameters of the system. A natural example might be the epsilon parameter in the LJ potential, which may provide more information as to what groups might be beneficially accommodated beyond methyl groups. In theory, any parameter of the system could be optimized, and looking at the bond, angle or torsion parameters using the methods developed here may also be interesting.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jcim.0c00610.

  • Details of protein preparation specific to each system; all ligand structures in full 3D with all optimized hydrogens labeled explicitly (Table S1); calculations for validation of free energies obtained from final optimized sigmas (Figures S1–S3 and Table S2); all convergence graphs not presented in the main text for ΔΔGscan calculations (Figures S4–S11); diagram for steric clashes in androgen receptor (Figure S12); all optimized sigmas for the menin B ligand (Table S3); all optimized charges for the thrombin B ligand (Table S4); Figure S13 depicting the thermodynamic cycle used for relative binding free energy calculations in this work (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • David J. Huggins - Tri-Institutional Therapeutics Discovery Institute, Belfer Research Building, 413 East 69th Street, 16th Floor, Box 300, New York, United StatesDepartment of Physiology and Biophysics, Weill Cornell Medical College of Cornell University, New York, New York 10065, United StatesOrcidhttp://orcid.org/0000-0003-1579-2496 Email: [email protected]
  • Author
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

A.D.W. would like to acknowledge the EPSRC Centre for Doctoral Training in Computational Methods for Materials Science for funding under grant number EP/L015552/1. All calculations were performed using the Darwin Supercomputer of the University of Cambridge High Performance Computing Service (http://www.hpc.cam.ac.uk/) and were funded by the EPSRC under grant EP/P020259/1. The authors gratefully acknowledge the support to the project generously provided by the Tri-Institutional Therapeutics Discovery Institute (TDI), a 501(c) (3) organization. TDI receives financial support from Takeda Pharmaceutical Company, TDI’s parent institutes (Memorial Sloan Kettering Cancer Center, The Rockefeller University and Weill Cornell Medicine) and from a generous contribution from Mr. Lewis Sanders and other philanthropic sources.

References

ARTICLE SECTIONS
Jump To

This article references 72 other publications.

  1. 1
    Wang, L.; Wu, Y.; Deng, Y.; Kim, B.; Pierce, L.; Krilov, G.; Lupyan, D.; Robinson, S.; Dahlgren, M. K.; Greenwood, J.; Romero, D. L.; Masse, C.; Knight, J. L.; Steinbrecher, T.; Beuming, T.; Damm, W.; Harder, E.; Sherman, W.; Brewer, M.; Wester, R.; Murcko, M.; Frye, L.; Farid, R.; Lin, T.; Mobley, D. L.; Jorgensen, W. L.; Berne, B. J.; Friesner, R. A.; Abel, R. Accurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force Field. J. Am. Chem. Soc. 2015, 137, 26952703,  DOI: 10.1021/ja512751q
  2. 2
    Shaw, D. E.; Deneroff, M. M.; Dror, R. O.; Kuskin, J. S.; Larson, R. H.; Salmon, J. K.; Young, C.; Batson, B.; Bowers, K. J.; Chao, J. C.; Eastwood, M. P.; Gagliardo, J.; Grossman, J. P.; Ho, C. R.; Ierardi, D. J.; Kolossváry, I.; Klepeis, J. L.; Layman, T.; McLeavey, C.; Moraes, M. A.; Mueller, R.; Priest, E. C.; Shan, Y.; Spengler, J.; Theobald, M.; Towles, B.; Wang, S. C. Anton, a Special-Purpose Machine for Molecular Dynamics Simulation. Commun. ACM 2008, 51 (7), 9197,  DOI: 10.1145/1364782.1364802
  3. 3
    Shaw, D. E.; Grossman, J.P.; Bank, J. A.; Batson, B.; Butts, J. A.; Chao, J. C.; Deneroff, M. M.; Dror, R. O.; Even, A.; Fenton, C. H.; Forte, A.; Gagliardo, J.; Gill, G.; Greskamp, B.; Ho, C. R.; Ierardi, D. J.; Iserovich, L.; Kuskin, J. S.; Larson, R. H.; Layman, T.; Lee, L.-S.; Lerer, A. K.; Li, C.; Killebrew, D.; Mackenzie, K. M.; Mok, S. Y.-H.; Moraes, M. A.; Mueller, R.; Nociolo, L. J.; Peticolas, J. L.; Quan, T.; Ramot, D.; Salmon, J. K.; Scarpazza, D. P.; Schafer, U. B.; Siddique, N.; Snyder, C. W.; Spengler, J.; Tang, P. T. P.; Theobald, M.; Toma, H.; Towles, B.; Vitale, B.; Wang, S. C.; Young, C. Anton 2: Raising the Bar for Performance and Programmability in a Special-Purpose Molecular Dynamics Supercomputer. In SC ‘14: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis 2014, 4153,  DOI: 10.1109/SC.2014.9
  4. 4
    Lindorff-Larsen, K.; Maragakis, P.; Piana, S.; Shaw, D. E. Picosecond to Millisecond Structural Dynamics in Human Ubiquitin. J. Phys. Chem. B 2016, 120 (33), 83138320,  DOI: 10.1021/acs.jpcb.6b02024
  5. 5
    Pierce, L. C. T.; Salomon-Ferrer, R.; Augusto F de Oliveira, C.; McCammon, J. A.; Walker, R. C. Routine Access to Millisecond Time Scale Events with Accelerated Molecular Dynamics. J. Chem. Theory Comput. 2012, 8 (9), 29973002,  DOI: 10.1021/ct300284c
  6. 6
    Salomon-Ferrer, R.; Götz, A. W.; Poole, D.; Le Grand, S.; Walker, R. C. Routine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 2. Explicit Solvent Particle Mesh Ewald. J. Chem. Theory Comput. 2013, 9 (9), 38783888,  DOI: 10.1021/ct400314y
  7. 7
    Eastman, P.; Swails, J.; Chodera, J. D.; McGibbon, R. T.; Zhao, Y.; Beauchamp, K. A.; Wang, L.-P.; Simmonett, A. C.; Harrigan, M. P.; Stern, C. D.; Wiewiora, R. P.; Brooks, B. R.; Pande, V. S. OpenMM 7: Rapid Development of High Performance Algorithms for Molecular Dynamics. PLoS Comput. Biol. 2017, 13, e1005659,  DOI: 10.1371/journal.pcbi.1005659
  8. 8
    Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J. Y.; Wang, L.; Lupyan, D.; Dahlgren, M. K.; Knight, J. L.; Kaus, J. W.; Cerutti, D. S.; Krilov, G.; Jorgensen, W. L.; Abel, R.; Friesner, R. A. OPLS3: A Force Field Providing Broad Coverage of Drug-like Small Molecules and Proteins. J. Chem. Theory Comput. 2016, 12 (1), 281296,  DOI: 10.1021/acs.jctc.5b00864
  9. 9
    Roos, K.; Wu, C.; Damm, W.; Reboul, M.; Stevenson, J. M.; Lu, C.; Dahlgren, M. K.; Mondal, S.; Chen, W.; Wang, L.; Abel, R.; Friesner, R. A.; Harder, E. D. OPLS3e: Extending Force Field Coverage for Drug-Like Small Molecules. J. Chem. Theory Comput. 2019, 15, 1863,  DOI: 10.1021/acs.jctc.8b01026
  10. 10
    Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.; Simmerling, C. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Theory Comput. 2015, 11 (8), 36963713,  DOI: 10.1021/acs.jctc.5b00255
  11. 11
    Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D. CHARMM General Force Field: A Force Field for Drug-like Molecules Compatible with the CHARMM All-Atom Additive Biological Force Fields. J. Comput. Chem. 2009, 31 (4), 671690,  DOI: 10.1002/jcc.21367
  12. 12
    Abel, R.; Wang, L.; Harder, E. D.; Berne, B. J.; Friesner, R. A. Advancing Drug Discovery through Enhanced Free Energy Calculations. Acc. Chem. Res. 2017, 50 (7), 16251632,  DOI: 10.1021/acs.accounts.7b00083
  13. 13
    Chodera, J. D.; Mobley, D. L.; Shirts, M. R.; Dixon, R. W.; Branson, K.; Pande, V. S. Alchemical Free Energy Methods for Drug Discovery: Progress and Challenges. Curr. Opin. Struct. Biol. 2011, 21 (2), 150160,  DOI: 10.1016/j.sbi.2011.01.011
  14. 14
    Cournia, Z.; Allen, B.; Sherman, W. Relative Binding Free Energy Calculations in Drug Discovery: Recent Advances and Practical Considerations. J. Chem. Inf. Model. 2017, 57 (12), 29112937,  DOI: 10.1021/acs.jcim.7b00564
  15. 15
    van Vlijmen, H.; Desjarlais, R. L.; Mirzadegan, T. Computational Chemistry at Janssen. J. Comput.-Aided Mol. Des. 2017, 31 (3), 267273,  DOI: 10.1007/s10822-016-9998-9
  16. 16
    Cole, D. J.; Janecek, M.; Stokes, J. E.; Rossmann, M.; Faver, J. C.; McKenzie, G. J.; Venkitaraman, A. R.; Hyvönen, M.; Spring, D. R.; Huggins, D. J.; Jorgensen, W. L. Computationally-Guided Optimization of Small-Molecule Inhibitors of the Aurora A kinase–TPX2 Protein–protein Interaction. Chem. Commun. 2017, 53, 93729375,  DOI: 10.1039/C7CC05379G
  17. 17
    Kuhn, M.; Firth-Clark, S.; Tosco, P.; Mey, A. S. J. S.; Mackey, M. D.; Michel, J. Assessment of Binding Affinity via Alchemical Free Energy Calculations. J. Chem. Inf. Model. 2020, 60, 3120,  DOI: 10.1021/acs.jcim.0c00165
  18. 18
    Riniker, S.; Christ, C. D.; Hansen, H. S.; Hünenberger, P. H.; Oostenbrink, C.; Steiner, D.; van Gunsteren, W. F. Calculation of Relative Free Energies for Ligand-Protein Binding, Solvation, and Conformational Transitions Using the GROMOS Software. J. Phys. Chem. B 2011, 115 (46), 1357013577,  DOI: 10.1021/jp204303a
  19. 19
    Bissantz, C.; Kuhn, B.; Stahl, M. A Medicinal Chemist’s Guide to Molecular Interactions. J. Med. Chem. 2010, 53 (14), 50615084,  DOI: 10.1021/jm100112j
  20. 20
    Wieder, M.; Garon, A.; Perricone, U.; Boresch, S.; Seidel, T.; Almerico, A. M.; Langer, T. Common Hits Approach: Combining Pharmacophore Modeling and Molecular Dynamics Simulations. J. Chem. Inf. Model. 2017, 57 (2), 365385,  DOI: 10.1021/acs.jcim.6b00674
  21. 21
    Newton, A. S.; Faver, J. C.; Micevic, G.; Muthusamy, V.; Kudalkar, S. N.; Bertoletti, N.; Anderson, K. S.; Bosenberg, M. W.; Jorgensen, W. L. Structure-Guided Identification of DNMT3B Inhibitors. ACS Med. Chem. Lett. 2020, 11 (5), 971976,  DOI: 10.1021/acsmedchemlett.0c00011
  22. 22
    Liosi, M.-E.; Krimmer, S. G.; Newton, A. S.; Dawson, T. K.; Puleo, D. E.; Cutrona, K. J.; Suzuki, Y.; Schlessinger, J.; Jorgensen, W. L. Selective Janus Kinase 2 (JAK2) Pseudokinase Ligands with a Diaminotriazole Core. J. Med. Chem. 2020, 63 (10), 53245340,  DOI: 10.1021/acs.jmedchem.0c00192
  23. 23
    Schindler, C.; Rippmann, F.; Kuhn, D. Relative Binding Affinity Prediction of Farnesoid X Receptor in the D3R Grand Challenge 2 Using FEP+. J. Comput.-Aided Mol. Des. 2018, 32 (1), 265272,  DOI: 10.1007/s10822-017-0064-z
  24. 24
    Keränen, H.; Pérez-Benito, L.; Ciordia, M.; Delgado, F.; Steinbrecher, T. B.; Oehlrich, D.; van Vlijmen, H. W. T.; Trabanco, A. A.; Tresadern, G. Acylguanidine Beta Secretase 1 Inhibitors: A Combined Experimental and Free Energy Perturbation Study. J. Chem. Theory Comput. 2017, 13 (3), 14391453,  DOI: 10.1021/acs.jctc.6b01141
  25. 25
    Song, L. F.; Lee, T.-S.; Zhu, C.; York, D. M.; Merz, K. M., Jr. Using AMBER18 for Relative Free Energy Calculations. J. Chem. Inf. Model. 2019, 59 (7), 31283135,  DOI: 10.1021/acs.jcim.9b00105
  26. 26
    Irwin, B. W. J.; Huggins, D. J. Estimating Atomic Contributions to Hydration and Binding Using Free Energy Perturbation. J. Chem. Theory Comput. 2018, 14 (6), 32183227,  DOI: 10.1021/acs.jctc.8b00027
  27. 27
    Lee, L.-P.; Tidor, B. Optimization of Electrostatic Binding Free Energy. J. Chem. Phys. 1997, 106 (21), 86818690,  DOI: 10.1063/1.473929
  28. 28
    Kangas, E.; Tidor, B. Electrostatic Complementarity at Ligand Binding Sites: Application to Chorismate Mutase. J. Phys. Chem. B 2001, 105 (4), 880888,  DOI: 10.1021/jp003449n
  29. 29
    Radhakrishnan, M. L. Theor. Chem. Acc. 2012, 131, 1252,  DOI: 10.1007/s00214-012-1252-5
  30. 30
    Wade, A. D.; Huggins, D. J. Optimization of Protein–Ligand Electrostatic Interactions Using an Alchemical Free-Energy Method. J. Chem. Theory Comput. 2019, 15 (11), 65046512,  DOI: 10.1021/acs.jctc.9b00976
  31. 31
    Wan, S.; Bhati, A. P.; Zasada, S. J.; Wall, I.; Green, D.; Bamborough, P.; Coveney, P. V. Rapid and Reliable Binding Affinity Prediction of Bromodomain Inhibitors: A Computational Study. J. Chem. Theory Comput. 2017, 13 (2), 784795,  DOI: 10.1021/acs.jctc.6b00794
  32. 32
    Lawrenz, M.; Baron, R.; McCammon, J. A. Independent-Trajectories Thermodynamic-Integration Free-Energy Changes for Biomolecular Systems: Determinants of H5N1 Avian Influenza Virus Neuraminidase Inhibition by Peramivir. J. Chem. Theory Comput. 2009, 5 (4), 11061116,  DOI: 10.1021/ct800559d
  33. 33
    Vilseck, J. Z.; Armacost, K. A.; Hayes, R. L.; Goh, G. B.; Brooks, C. L., 3rd Predicting Binding Free Energies in a Large Combinatorial Chemical Space Using Multisite λ Dynamics. J. Phys. Chem. Lett. 2018, 9 (12), 33283332,  DOI: 10.1021/acs.jpclett.8b01284
  34. 34
    Bennett, C. H. Efficient Estimation of Free Energy Differences from Monte Carlo Data. J. Comput. Phys. 1976, 22 (2), 245268,  DOI: 10.1016/0021-9991(76)90078-4
  35. 35
    Shirts, M. R.; Chodera, J. D. Statistically Optimal Analysis of Samples from Multiple Equilibrium States. J. Chem. Phys. 2008, 129 (12), 124105,  DOI: 10.1063/1.2978177
  36. 36
    Riniker, S.; Christ, C. D.; Hansen, N.; Mark, A. E.; Nair, P. C.; van Gunsteren, W. F. Comparison of Enveloping Distribution Sampling and Thermodynamic Integration to Calculate Binding Free Energies of Phenylethanolamine N-Methyltransferase Inhibitors. J. Chem. Phys. 2011, 135 (2), 024105,  DOI: 10.1063/1.3604534
  37. 37
    Liu, H.; Mark, A. E.; van Gunsteren, W. F. Estimating the Relative Free Energy of Different Molecular States with Respect to a Single Reference State. J. Phys. Chem. 1996, 100 (22), 94859494,  DOI: 10.1021/jp9605212
  38. 38
    Mordasini, T. Z.; McCammon, J. A. Calculations of Relative Hydration Free Energies: A Comparative Study Using Thermodynamic Integration and an Extrapolation Method Based on a Single Reference State. J. Phys. Chem. B 2000, 104, 360367,  DOI: 10.1021/jp993102o
  39. 39
    Raman, E. P.; Vanommeslaeghe, K.; MacKerell, A. D. Site-Specific Fragment Identification Guided by Single-Step Free Energy Perturbation Calculations. J. Chem. Theory Comput. 2012, 8, 35133525,  DOI: 10.1021/ct300088r
  40. 40
    Stroet, M.; Koziara, K. B.; Malde, A. K.; Mark, A. E. Optimization of Empirical Force Fields by Parameter Space Mapping: A Single-Step Perturbation Approach. J. Chem. Theory Comput. 2017, 13, 62016212,  DOI: 10.1021/acs.jctc.7b00800
  41. 41
    Oostenbrink, C.; Van Gunsteren, W. F. Single-Step Perturbations to Calculate Free Energy Differences from Unphysical Reference States: Limits on Size, Flexibility, and Character. J. Comput. Chem. 2003, 24, 17301739,  DOI: 10.1002/jcc.10304
  42. 42
    Oostenbrink, C.; van Gunsteren, W. F. Free Energies of Ligand Binding for Structurally Diverse Compounds. Proc. Natl. Acad. Sci. U. S. A. 2005, 102 (19), 67506754,  DOI: 10.1073/pnas.0407404102
  43. 43
    Wade, A. D.; Rizzi, A.; Wang, Y.; Huggins, D. J. Computational Fluorine Scanning Using Free-Energy Perturbation. J. Chem. Inf. Model. 2019, 59, 2776,  DOI: 10.1021/acs.jcim.9b00228
  44. 44
    Raman, E. P.; Lakkaraju, S. K.; Denny, R. A.; MacKerell, A. D. Estimation of Relative Free Energies of Binding Using Pre-Computed Ensembles Based on the Single-Step Free Energy Perturbation and the Site-Identification by Ligand Competitive Saturation Approaches. J. Comput. Chem. 2017, 38, 12381251,  DOI: 10.1002/jcc.24522
  45. 45
    Oostenbrink, C. Free Energy Calculations from One-Step Perturbations. Methods Mol. Biol. 2012, 819, 487499,  DOI: 10.1007/978-1-61779-465-0_28
  46. 46
    Baron, R. Computational Drug Discovery and Design; Baron, R., Ed.; Springer: New York, 2012.
  47. 47
    Graf, M. M. H.; Zhixiong, L.; Bren, U.; Haltrich, D.; van Gunsteren, W. F.; Oostenbrink, C. Pyranose Dehydrogenase Ligand Promiscuity: A Generalized Approach to Simulate Monosaccharide Solvation, Binding, and Product Formation. PLoS Comput. Biol. 2014, 10 (12), e1003995  DOI: 10.1371/journal.pcbi.1003995
  48. 48
    Perthold, J. W.; Petrov, D.; Oostenbrink, C. Towards Automated Free Energy Calculation with Accelerated Enveloping Distribution Sampling (A-EDS). J. Chem. Inf. Model. 2020, DOI: 10.1021/acs.jcim.0c00456 .
  49. 49
    Perthold, J. W.; Oostenbrink, C. Accelerated Enveloping Distribution Sampling: Enabling Sampling of Multiple End States While Preserving Local Energy Minima. J. Phys. Chem. B 2018, 122 (19), 50305037,  DOI: 10.1021/acs.jpcb.8b02725
  50. 50
    Aldeghi, M.; Heifetz, A.; Bodkin, M. J.; Knapp, S.; Biggin, P. C. Accurate Calculation of the Absolute Free Energy of Binding for Drug Molecules. Chem. Sci. 2016, 7 (1), 207218,  DOI: 10.1039/C5SC02678D
  51. 51
    Woods, C. J.; Essex, J. W.; King, M. A. The Development of Replica-Exchange-Based Free-Energy Methods. J. Phys. Chem. B 2003, 107 (49), 1370313710,  DOI: 10.1021/jp0356620
  52. 52
    Chodera, J. D.; Shirts, M. R. Replica Exchange and Expanded Ensemble Simulations as Gibbs Sampling: Simple Improvements for Enhanced Mixing. J. Chem. Phys. 2011, 135 (19), 194110,  DOI: 10.1063/1.3660669
  53. 53
    Macdonald, H. B.; Cave-Ayland, C.; Essex, J. Predicting Water Networks and Ligand Binding Free Energies in Proteins Using Grand Canonical Monte Carlo. In Abstracts of Papers of the American Chemical Society; American Chemical Society: Washington, DC, 2018; Vol. 255.
  54. 54
    Ross, G. A.; Bruce Macdonald, H. E.; Cave-Ayland, C.; Cabedo Martinez, A. I.; Essex, J. W. Replica-Exchange and Standard State Binding Free Energies with Grand Canonical Monte Carlo. J. Chem. Theory Comput. 2017, 13 (12), 63736381,  DOI: 10.1021/acs.jctc.7b00738
  55. 55
    Ben-Shalom, I. Y.; Lin, C.; Kurtzman, T.; Walker, R.; Gilson, M. K. Equilibration of Buried Water Molecules to Enhance Protein-Ligand Binding Free Energy Calculations. Biophys. J. 2020, 118 (3), 144a,  DOI: 10.1016/j.bpj.2019.11.908
  56. 56
    Jagger, B. R.; Kochanek, S. E.; Haldar, S.; Amaro, R. E.; Mulholland, A. J. Multiscale Simulation Approaches to Modeling Drug–protein Binding. Curr. Opin. Struct. Biol. 2020, 61, 213221,  DOI: 10.1016/j.sbi.2020.01.014
  57. 57
    Haldar, S.; Comitani, F.; Saladino, G.; Woods, C.; van der Kamp, M. W.; Mulholland, A. J.; Gervasio, F. L. A Multiscale Simulation Approach to Modeling Drug–Protein Binding Kinetics. J. Chem. Theory Comput. 2018, 14 (11), 60936101,  DOI: 10.1021/acs.jctc.8b00687
  58. 58
    Hamann, L. G.; Manfredi, M. C.; Sun, C.; Krystek, S. R.; Huang, Y.; Bi, Y.; Augeri, D. J.; Wang, T.; Zou, Y.; Betebenner, D. A.; Fura, A.; Seethala, R.; Golla, R.; Kuhns, J. E.; Lupisella, J. A.; Darienzo, C. J.; Custer, L. L.; Price, J. L.; Johnson, J. M.; Biller, S. A.; Zahler, R.; Ostrowski, J. Tandem Optimization of Target Activity and Elimination of Mutagenic Potential in a Potent Series of N-Aryl Bicyclic Hydantoin-Based Selective Androgen Receptor Modulators. Bioorg. Med. Chem. Lett. 2007, 17 (7), 18601864,  DOI: 10.1016/j.bmcl.2007.01.076
  59. 59
    Ostrowski, J.; Kuhns, J. E.; Lupisella, J. A.; Manfredi, M. C.; Beehler, B. C.; Krystek, S. R., Jr; Bi, Y.; Sun, C.; Seethala, R.; Golla, R.; Sleph, P. G.; Fura, A.; An, Y.; Kish, K. F.; Sack, J. S.; Mookhtiar, K. A.; Grover, G. J.; Hamann, L. G. Pharmacological and X-Ray Structural Characterization of a Novel Selective Androgen Receptor Modulator: Potent Hyperanabolic Stimulation of Skeletal Muscle with Hypostimulation of Prostate in Rats. Endocrinology 2007, 148 (1), 412,  DOI: 10.1210/en.2006-0843
  60. 60
    Matter, H.; Scheiper, B.; Steinhagen, H.; Böcskei, Z.; Fleury, V.; McCort, G. Structure-Based Design and Optimization of Potent Renin Inhibitors on 5- or 7-Azaindole-Scaffolds. Bioorg. Med. Chem. Lett. 2011, 21 (18), 54875492,  DOI: 10.1016/j.bmcl.2011.06.112
  61. 61
    He, S.; Senter, T. J.; Pollock, J.; Han, C.; Upadhyay, S. K.; Purohit, T.; Gogliotti, R. D.; Lindsley, C. W.; Cierpicki, T.; Stauffer, S. R.; Grembecka, J. High-Affinity Small-Molecule Inhibitors of the Menin-Mixed Lineage Leukemia (MLL) Interaction Closely Mimic a Natural Protein–Protein Interaction. J. Med. Chem. 2014, 57 (4), 15431556,  DOI: 10.1021/jm401868d
  62. 62
    Baum, B.; Muley, L.; Smolinski, M.; Heine, A.; Hangauer, D.; Klebe, G. Non-Additivity of Functional Group Contributions in Protein–Ligand Binding: A Comprehensive Study by Crystallography and Isothermal Titration Calorimetry. J. Mol. Biol. 2010, 397 (4), 10421054,  DOI: 10.1016/j.jmb.2010.02.007
  63. 63
    Ghosh, A. K.; Takayama, J.; Aubin, Y.; Ratia, K.; Chaudhuri, R.; Baez, Y.; Sleeman, K.; Coughlin, M.; Nichols, D. B.; Mulhearn, D. C.; Prabhakar, B. S.; Baker, S. C.; Johnson, M. E.; Mesecar, A. D. Structure-Based Design, Synthesis, and Biological Evaluation of a Series of Novel and Reversible Inhibitors for the Severe Acute Respiratory Syndrome- Coronavirus Papain-Like Protease. J. Med. Chem. 2009, 52 (16), 52285240,  DOI: 10.1021/jm900611t
  64. 64
    Ratia, K.; Pegan, S.; Takayama, J.; Sleeman, K.; Coughlin, M.; Baliji, S.; Chaudhuri, R.; Fu, W.; Prabhakar, B. S.; Johnson, M. E.; Baker, S. C.; Ghosh, A. K.; Mesecar, A. D. A Noncovalent Class of Papain-like Protease/deubiquitinase Inhibitors Blocks SARS Virus Replication. Proc. Natl. Acad. Sci. U. S. A. 2008, 105 (42), 1611916124,  DOI: 10.1073/pnas.0805240105
  65. 65
    Maestro, version 9.1; Schrödinger, LLC: New York, 2010.
  66. 66
    Wang, J.; Wang, W.; Kollman, P. A.; Case, D. A. Antechamber: An Accessory Software Package for Molecular Mechanical Calculations. J. Am. Chem. Soc. 2001, 222, U403
  67. 67
    Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25 (9), 11571174,  DOI: 10.1002/jcc.20035
  68. 68
    Jakalian, A.; Jack, D. B.; Bayly, C. I. Fast, Efficient Generation of High-Quality Atomic Charges. AM1-BCC Model: II. Parameterization and Validation. J. Comput. Chem. 2002, 23 (16), 16231641,  DOI: 10.1002/jcc.10128
  69. 69
    Naden, L. N.; Shirts, M. R. Linear Basis Function Approach to Efficient Alchemical Free Energy Calculations. 2. Inserting and Deleting Particles with Coulombic Interactions. J. Chem. Theory Comput. 2015, 11 (6), 25362549,  DOI: 10.1021/ct501047e
  70. 70
    Pechlaner, M.; Reif, M. M.; Oostenbrink, C. Reparametrisation of United-Atom Amine Solvation in the GROMOS Force Field. Mol. Phys. 2017, 115 (9–12), 11441154,  DOI: 10.1080/00268976.2016.1255797
  71. 71
    Diem, M.; Oostenbrink, C. Hamiltonian Reweighing To Refine Protein Backbone Dihedral Angle Parameters in the GROMOS Force Field. J. Chem. Inf. Model. 2020, 60 (1), 279288,  DOI: 10.1021/acs.jcim.9b01034
  72. 72
    Chodera, J.; Rizzi, A.; Naden, L.; Beauchamp, K.; Grinaway, P.; Fass, J.; Rustenburg, B.; Ross, G. A.; Simmonett, A.; Swenson, D. W. H. Choderalab/openmmtools: 0.14. 0-Exact Treatment of Alchemical PME Electrostatics, Water Cluster Test System, Optimizations. https://zenodo.org/record/1161149#.X0gc6HlKjIU.

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 1 publications.

  1. Alexander D. Wade, Agastya P. Bhati, Shunzhou Wan, Peter V. Coveney. Alchemical Free Energy Estimators and Molecular Dynamics Engines: Accuracy, Precision, and Reproducibility. Journal of Chemical Theory and Computation 2022, 18 (6) , 3972-3987. https://doi.org/10.1021/acs.jctc.2c00114
  • Abstract

    Figure 1

    Figure 1. Diagrammatic workflow for calculations performed in this work. SSP gradient method uses an exponential averaging method to calculate free energies. MBAR objective and FEP scan both use the MBAR estimator.

    Figure 2

    Figure 2. Convergence for a calculation of the objective in the androgen receptor test case. ΔΔGopt is reported as mean of three replicates with shaded area showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions.

    Figure 3

    Figure 3. 3D structure for the androgen receptor ligand with all the optimized hydrogens explicitly labeled.

    Figure 4

    Figure 4. Convergence for a calculation of the gradient in the androgen receptor test case. ΔΔGgrag/h is reported as the mean of three replicates with the shaded area showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees. Atom names in the legend can be seen in Figure 3.

    Figure 5

    Figure 5. Cumulative ΔΔGopt calculated by summing the MBAR objective for each step of the optimization plotted against optimization step for every system.

    Figure 6

    Figure 6. ΔΔGscan for each methylation in the androgen receptor system as the amount of sampling is increased. ΔΔGscan are reported as mean of three replicates with shaded area showing 95% confidence interval computed as mean ± t2·SEM, where t2 is the t-distribution statistic with two degrees of freedom, and SEM is the standard error of the mean computed from the sample standard deviation of the three independent replicate predictions. All hydrogen names are taken from Figure 3.

    Figure 7

    Figure 7. Androgen receptor ligand with all optimized hydrogens sized in proportion to their calculated Δσ.

    Figure 8

    Figure 8. SARS PLPro ligand with all optimized hydrogens sized in proportion to their calculated Δσ.

    Figure 9

    Figure 9. Renin ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

    Figure 10

    Figure 10. Menin A ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

    Figure 11

    Figure 11. Menin B ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

    Figure 12

    Figure 12. Thrombin A ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

    Figure 13

    Figure 13. Thrombin B ligand with all optimized hydrogens colored in proportion to their calculated Δq. Red is more positive and blue more negative. See all named hydrogens in Table S1.

    Figure 14

    Figure 14. Thrombin C ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

    Figure 15

    Figure 15. Thrombin D ligand with all optimized hydrogens sized in proportion to their calculated Δσ. See all named hydrogens in Table S1.

    Figure 16

    Figure 16. Ligand with improved binding free energy found experimentally and final computationally optimized ligand built by choosing perturbations highlighted by our optimization methodology.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 72 other publications.

    1. 1
      Wang, L.; Wu, Y.; Deng, Y.; Kim, B.; Pierce, L.; Krilov, G.; Lupyan, D.; Robinson, S.; Dahlgren, M. K.; Greenwood, J.; Romero, D. L.; Masse, C.; Knight, J. L.; Steinbrecher, T.; Beuming, T.; Damm, W.; Harder, E.; Sherman, W.; Brewer, M.; Wester, R.; Murcko, M.; Frye, L.; Farid, R.; Lin, T.; Mobley, D. L.; Jorgensen, W. L.; Berne, B. J.; Friesner, R. A.; Abel, R. Accurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force Field. J. Am. Chem. Soc. 2015, 137, 26952703,  DOI: 10.1021/ja512751q
    2. 2
      Shaw, D. E.; Deneroff, M. M.; Dror, R. O.; Kuskin, J. S.; Larson, R. H.; Salmon, J. K.; Young, C.; Batson, B.; Bowers, K. J.; Chao, J. C.; Eastwood, M. P.; Gagliardo, J.; Grossman, J. P.; Ho, C. R.; Ierardi, D. J.; Kolossváry, I.; Klepeis, J. L.; Layman, T.; McLeavey, C.; Moraes, M. A.; Mueller, R.; Priest, E. C.; Shan, Y.; Spengler, J.; Theobald, M.; Towles, B.; Wang, S. C. Anton, a Special-Purpose Machine for Molecular Dynamics Simulation. Commun. ACM 2008, 51 (7), 9197,  DOI: 10.1145/1364782.1364802
    3. 3
      Shaw, D. E.; Grossman, J.P.; Bank, J. A.; Batson, B.; Butts, J. A.; Chao, J. C.; Deneroff, M. M.; Dror, R. O.; Even, A.; Fenton, C. H.; Forte, A.; Gagliardo, J.; Gill, G.; Greskamp, B.; Ho, C. R.; Ierardi, D. J.; Iserovich, L.; Kuskin, J. S.; Larson, R. H.; Layman, T.; Lee, L.-S.; Lerer, A. K.; Li, C.; Killebrew, D.; Mackenzie, K. M.; Mok, S. Y.-H.; Moraes, M. A.; Mueller, R.; Nociolo, L. J.; Peticolas, J. L.; Quan, T.; Ramot, D.; Salmon, J. K.; Scarpazza, D. P.; Schafer, U. B.; Siddique, N.; Snyder, C. W.; Spengler, J.; Tang, P. T. P.; Theobald, M.; Toma, H.; Towles, B.; Vitale, B.; Wang, S. C.; Young, C. Anton 2: Raising the Bar for Performance and Programmability in a Special-Purpose Molecular Dynamics Supercomputer. In SC ‘14: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis 2014, 4153,  DOI: 10.1109/SC.2014.9
    4. 4
      Lindorff-Larsen, K.; Maragakis, P.; Piana, S.; Shaw, D. E. Picosecond to Millisecond Structural Dynamics in Human Ubiquitin. J. Phys. Chem. B 2016, 120 (33), 83138320,  DOI: 10.1021/acs.jpcb.6b02024
    5. 5
      Pierce, L. C. T.; Salomon-Ferrer, R.; Augusto F de Oliveira, C.; McCammon, J. A.; Walker, R. C. Routine Access to Millisecond Time Scale Events with Accelerated Molecular Dynamics. J. Chem. Theory Comput. 2012, 8 (9), 29973002,  DOI: 10.1021/ct300284c
    6. 6
      Salomon-Ferrer, R.; Götz, A. W.; Poole, D.; Le Grand, S.; Walker, R. C. Routine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 2. Explicit Solvent Particle Mesh Ewald. J. Chem. Theory Comput. 2013, 9 (9), 38783888,  DOI: 10.1021/ct400314y
    7. 7
      Eastman, P.; Swails, J.; Chodera, J. D.; McGibbon, R. T.; Zhao, Y.; Beauchamp, K. A.; Wang, L.-P.; Simmonett, A. C.; Harrigan, M. P.; Stern, C. D.; Wiewiora, R. P.; Brooks, B. R.; Pande, V. S. OpenMM 7: Rapid Development of High Performance Algorithms for Molecular Dynamics. PLoS Comput. Biol. 2017, 13, e1005659,  DOI: 10.1371/journal.pcbi.1005659
    8. 8
      Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J. Y.; Wang, L.; Lupyan, D.; Dahlgren, M. K.; Knight, J. L.; Kaus, J. W.; Cerutti, D. S.; Krilov, G.; Jorgensen, W. L.; Abel, R.; Friesner, R. A. OPLS3: A Force Field Providing Broad Coverage of Drug-like Small Molecules and Proteins. J. Chem. Theory Comput. 2016, 12 (1), 281296,  DOI: 10.1021/acs.jctc.5b00864
    9. 9
      Roos, K.; Wu, C.; Damm, W.; Reboul, M.; Stevenson, J. M.; Lu, C.; Dahlgren, M. K.; Mondal, S.; Chen, W.; Wang, L.; Abel, R.; Friesner, R. A.; Harder, E. D. OPLS3e: Extending Force Field Coverage for Drug-Like Small Molecules. J. Chem. Theory Comput. 2019, 15, 1863,  DOI: 10.1021/acs.jctc.8b01026
    10. 10
      Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.; Simmerling, C. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Theory Comput. 2015, 11 (8), 36963713,  DOI: 10.1021/acs.jctc.5b00255
    11. 11
      Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D. CHARMM General Force Field: A Force Field for Drug-like Molecules Compatible with the CHARMM All-Atom Additive Biological Force Fields. J. Comput. Chem. 2009, 31 (4), 671690,  DOI: 10.1002/jcc.21367
    12. 12
      Abel, R.; Wang, L.; Harder, E. D.; Berne, B. J.; Friesner, R. A. Advancing Drug Discovery through Enhanced Free Energy Calculations. Acc. Chem. Res. 2017, 50 (7), 16251632,  DOI: 10.1021/acs.accounts.7b00083
    13. 13
      Chodera, J. D.; Mobley, D. L.; Shirts, M. R.; Dixon, R. W.; Branson, K.; Pande, V. S. Alchemical Free Energy Methods for Drug Discovery: Progress and Challenges. Curr. Opin. Struct. Biol. 2011, 21 (2), 150160,  DOI: 10.1016/j.sbi.2011.01.011
    14. 14
      Cournia, Z.; Allen, B.; Sherman, W. Relative Binding Free Energy Calculations in Drug Discovery: Recent Advances and Practical Considerations. J. Chem. Inf. Model. 2017, 57 (12), 29112937,  DOI: 10.1021/acs.jcim.7b00564
    15. 15
      van Vlijmen, H.; Desjarlais, R. L.; Mirzadegan, T. Computational Chemistry at Janssen. J. Comput.-Aided Mol. Des. 2017, 31 (3), 267273,  DOI: 10.1007/s10822-016-9998-9
    16. 16
      Cole, D. J.; Janecek, M.; Stokes, J. E.; Rossmann, M.; Faver, J. C.; McKenzie, G. J.; Venkitaraman, A. R.; Hyvönen, M.; Spring, D. R.; Huggins, D. J.; Jorgensen, W. L. Computationally-Guided Optimization of Small-Molecule Inhibitors of the Aurora A kinase–TPX2 Protein–protein Interaction. Chem. Commun. 2017, 53, 93729375,  DOI: 10.1039/C7CC05379G
    17. 17
      Kuhn, M.; Firth-Clark, S.; Tosco, P.; Mey, A. S. J. S.; Mackey, M. D.; Michel, J. Assessment of Binding Affinity via Alchemical Free Energy Calculations. J. Chem. Inf. Model. 2020, 60, 3120,  DOI: 10.1021/acs.jcim.0c00165
    18. 18
      Riniker, S.; Christ, C. D.; Hansen, H. S.; Hünenberger, P. H.; Oostenbrink, C.; Steiner, D.; van Gunsteren, W. F. Calculation of Relative Free Energies for Ligand-Protein Binding, Solvation, and Conformational Transitions Using the GROMOS Software. J. Phys. Chem. B 2011, 115 (46), 1357013577,  DOI: 10.1021/jp204303a
    19. 19
      Bissantz, C.; Kuhn, B.; Stahl, M. A Medicinal Chemist’s Guide to Molecular Interactions. J. Med. Chem. 2010, 53 (14), 50615084,  DOI: 10.1021/jm100112j
    20. 20
      Wieder, M.; Garon, A.; Perricone, U.; Boresch, S.; Seidel, T.; Almerico, A. M.; Langer, T. Common Hits Approach: Combining Pharmacophore Modeling and Molecular Dynamics Simulations. J. Chem. Inf. Model. 2017, 57 (2), 365385,  DOI: 10.1021/acs.jcim.6b00674
    21. 21
      Newton, A. S.; Faver, J. C.; Micevic, G.; Muthusamy, V.; Kudalkar, S. N.; Bertoletti, N.; Anderson, K. S.; Bosenberg, M. W.; Jorgensen, W. L. Structure-Guided Identification of DNMT3B Inhibitors. ACS Med. Chem. Lett. 2020, 11 (5), 971976,  DOI: 10.1021/acsmedchemlett.0c00011
    22. 22
      Liosi, M.-E.; Krimmer, S. G.; Newton, A. S.; Dawson, T. K.; Puleo, D. E.; Cutrona, K. J.; Suzuki, Y.; Schlessinger, J.; Jorgensen, W. L. Selective Janus Kinase 2 (JAK2) Pseudokinase Ligands with a Diaminotriazole Core. J. Med. Chem. 2020, 63 (10), 53245340,  DOI: 10.1021/acs.jmedchem.0c00192
    23. 23
      Schindler, C.; Rippmann, F.; Kuhn, D. Relative Binding Affinity Prediction of Farnesoid X Receptor in the D3R Grand Challenge 2 Using FEP+. J. Comput.-Aided Mol. Des. 2018, 32 (1), 265272,  DOI: 10.1007/s10822-017-0064-z
    24. 24
      Keränen, H.; Pérez-Benito, L.; Ciordia, M.; Delgado, F.; Steinbrecher, T. B.; Oehlrich, D.; van Vlijmen, H. W. T.; Trabanco, A. A.; Tresadern, G. Acylguanidine Beta Secretase 1 Inhibitors: A Combined Experimental and Free Energy Perturbation Study. J. Chem. Theory Comput. 2017, 13 (3), 14391453,  DOI: 10.1021/acs.jctc.6b01141
    25. 25
      Song, L. F.; Lee, T.-S.; Zhu, C.; York, D. M.; Merz, K. M., Jr. Using AMBER18 for Relative Free Energy Calculations. J. Chem. Inf. Model. 2019, 59 (7), 31283135,  DOI: 10.1021/acs.jcim.9b00105
    26. 26
      Irwin, B. W. J.; Huggins, D. J. Estimating Atomic Contributions to Hydration and Binding Using Free Energy Perturbation. J. Chem. Theory Comput. 2018, 14 (6), 32183227,  DOI: 10.1021/acs.jctc.8b00027
    27. 27
      Lee, L.-P.; Tidor, B. Optimization of Electrostatic Binding Free Energy. J. Chem. Phys. 1997, 106 (21), 86818690,  DOI: 10.1063/1.473929
    28. 28
      Kangas, E.; Tidor, B. Electrostatic Complementarity at Ligand Binding Sites: Application to Chorismate Mutase. J. Phys. Chem. B 2001, 105 (4), 880888,  DOI: 10.1021/jp003449n
    29. 29
      Radhakrishnan, M. L. Theor. Chem. Acc. 2012, 131, 1252,  DOI: 10.1007/s00214-012-1252-5
    30. 30
      Wade, A. D.; Huggins, D. J. Optimization of Protein–Ligand Electrostatic Interactions Using an Alchemical Free-Energy Method. J. Chem. Theory Comput. 2019, 15 (11), 65046512,  DOI: 10.1021/acs.jctc.9b00976
    31. 31
      Wan, S.; Bhati, A. P.; Zasada, S. J.; Wall, I.; Green, D.; Bamborough, P.; Coveney, P. V. Rapid and Reliable Binding Affinity Prediction of Bromodomain Inhibitors: A Computational Study. J. Chem. Theory Comput. 2017, 13 (2), 784795,  DOI: 10.1021/acs.jctc.6b00794
    32. 32
      Lawrenz, M.; Baron, R.; McCammon, J. A. Independent-Trajectories Thermodynamic-Integration Free-Energy Changes for Biomolecular Systems: Determinants of H5N1 Avian Influenza Virus Neuraminidase Inhibition by Peramivir. J. Chem. Theory Comput. 2009, 5 (4), 11061116,  DOI: 10.1021/ct800559d
    33. 33
      Vilseck, J. Z.; Armacost, K. A.; Hayes, R. L.; Goh, G. B.; Brooks, C. L., 3rd Predicting Binding Free Energies in a Large Combinatorial Chemical Space Using Multisite λ Dynamics. J. Phys. Chem. Lett. 2018, 9 (12), 33283332,  DOI: 10.1021/acs.jpclett.8b01284
    34. 34
      Bennett, C. H. Efficient Estimation of Free Energy Differences from Monte Carlo Data. J. Comput. Phys. 1976, 22 (2), 245268,  DOI: 10.1016/0021-9991(76)90078-4
    35. 35
      Shirts, M. R.; Chodera, J. D. Statistically Optimal Analysis of Samples from Multiple Equilibrium States. J. Chem. Phys. 2008, 129 (12), 124105,  DOI: 10.1063/1.2978177
    36. 36
      Riniker, S.; Christ, C. D.; Hansen, N.; Mark, A. E.; Nair, P. C.; van Gunsteren, W. F. Comparison of Enveloping Distribution Sampling and Thermodynamic Integration to Calculate Binding Free Energies of Phenylethanolamine N-Methyltransferase Inhibitors. J. Chem. Phys. 2011, 135 (2), 024105,  DOI: 10.1063/1.3604534
    37. 37
      Liu, H.; Mark, A. E.; van Gunsteren, W. F. Estimating the Relative Free Energy of Different Molecular States with Respect to a Single Reference State. J. Phys. Chem. 1996, 100 (22), 94859494,  DOI: 10.1021/jp9605212
    38. 38
      Mordasini, T. Z.; McCammon, J. A. Calculations of Relative Hydration Free Energies: A Comparative Study Using Thermodynamic Integration and an Extrapolation Method Based on a Single Reference State. J. Phys. Chem. B 2000, 104, 360367,  DOI: 10.1021/jp993102o
    39. 39
      Raman, E. P.; Vanommeslaeghe, K.; MacKerell, A. D. Site-Specific Fragment Identification Guided by Single-Step Free Energy Perturbation Calculations. J. Chem. Theory Comput. 2012, 8, 35133525,  DOI: 10.1021/ct300088r
    40. 40
      Stroet, M.; Koziara, K. B.; Malde, A. K.; Mark, A. E. Optimization of Empirical Force Fields by Parameter Space Mapping: A Single-Step Perturbation Approach. J. Chem. Theory Comput. 2017, 13, 62016212,  DOI: 10.1021/acs.jctc.7b00800
    41. 41
      Oostenbrink, C.; Van Gunsteren, W. F. Single-Step Perturbations to Calculate Free Energy Differences from Unphysical Reference States: Limits on Size, Flexibility, and Character. J. Comput. Chem. 2003, 24, 17301739,  DOI: 10.1002/jcc.10304
    42. 42
      Oostenbrink, C.; van Gunsteren, W. F. Free Energies of Ligand Binding for Structurally Diverse Compounds. Proc. Natl. Acad. Sci. U. S. A. 2005, 102 (19), 67506754,  DOI: 10.1073/pnas.0407404102
    43. 43
      Wade, A. D.; Rizzi, A.; Wang, Y.; Huggins, D. J. Computational Fluorine Scanning Using Free-Energy Perturbation. J. Chem. Inf. Model. 2019, 59, 2776,  DOI: 10.1021/acs.jcim.9b00228
    44. 44
      Raman, E. P.; Lakkaraju, S. K.; Denny, R. A.; MacKerell, A. D. Estimation of Relative Free Energies of Binding Using Pre-Computed Ensembles Based on the Single-Step Free Energy Perturbation and the Site-Identification by Ligand Competitive Saturation Approaches. J. Comput. Chem. 2017, 38, 12381251,  DOI: 10.1002/jcc.24522
    45. 45
      Oostenbrink, C. Free Energy Calculations from One-Step Perturbations. Methods Mol. Biol. 2012, 819, 487499,  DOI: 10.1007/978-1-61779-465-0_28
    46. 46
      Baron, R. Computational Drug Discovery and Design; Baron, R., Ed.; Springer: New York, 2012.
    47. 47
      Graf, M. M. H.; Zhixiong, L.; Bren, U.; Haltrich, D.; van Gunsteren, W. F.; Oostenbrink, C. Pyranose Dehydrogenase Ligand Promiscuity: A Generalized Approach to Simulate Monosaccharide Solvation, Binding, and Product Formation. PLoS Comput. Biol. 2014, 10 (12), e1003995  DOI: 10.1371/journal.pcbi.1003995
    48. 48
      Perthold, J. W.; Petrov, D.; Oostenbrink, C. Towards Automated Free Energy Calculation with Accelerated Enveloping Distribution Sampling (A-EDS). J. Chem. Inf. Model. 2020, DOI: 10.1021/acs.jcim.0c00456 .
    49. 49
      Perthold, J. W.; Oostenbrink, C. Accelerated Enveloping Distribution Sampling: Enabling Sampling of Multiple End States While Preserving Local Energy Minima. J. Phys. Chem. B 2018, 122 (19), 50305037,  DOI: 10.1021/acs.jpcb.8b02725
    50. 50
      Aldeghi, M.; Heifetz, A.; Bodkin, M. J.; Knapp, S.; Biggin, P. C. Accurate Calculation of the Absolute Free Energy of Binding for Drug Molecules. Chem. Sci. 2016, 7 (1), 207218,  DOI: 10.1039/C5SC02678D
    51. 51
      Woods, C. J.; Essex, J. W.; King, M. A. The Development of Replica-Exchange-Based Free-Energy Methods. J. Phys. Chem. B 2003, 107 (49), 1370313710,  DOI: 10.1021/jp0356620
    52. 52
      Chodera, J. D.; Shirts, M. R. Replica Exchange and Expanded Ensemble Simulations as Gibbs Sampling: Simple Improvements for Enhanced Mixing. J. Chem. Phys. 2011, 135 (19), 194110,  DOI: 10.1063/1.3660669
    53. 53
      Macdonald, H. B.; Cave-Ayland, C.; Essex, J. Predicting Water Networks and Ligand Binding Free Energies in Proteins Using Grand Canonical Monte Carlo. In Abstracts of Papers of the American Chemical Society; American Chemical Society: Washington, DC, 2018; Vol. 255.
    54. 54
      Ross, G. A.; Bruce Macdonald, H. E.; Cave-Ayland, C.; Cabedo Martinez, A. I.; Essex, J. W. Replica-Exchange and Standard State Binding Free Energies with Grand Canonical Monte Carlo. J. Chem. Theory Comput. 2017, 13 (12), 63736381,  DOI: 10.1021/acs.jctc.7b00738
    55. 55
      Ben-Shalom, I. Y.; Lin, C.; Kurtzman, T.; Walker, R.; Gilson, M. K. Equilibration of Buried Water Molecules to Enhance Protein-Ligand Binding Free Energy Calculations. Biophys. J. 2020, 118 (3), 144a,  DOI: 10.1016/j.bpj.2019.11.908
    56. 56
      Jagger, B. R.; Kochanek, S. E.; Haldar, S.; Amaro, R. E.; Mulholland, A. J. Multiscale Simulation Approaches to Modeling Drug–protein Binding. Curr. Opin. Struct. Biol. 2020, 61, 213221,  DOI: 10.1016/j.sbi.2020.01.014
    57. 57
      Haldar, S.; Comitani, F.; Saladino, G.; Woods, C.; van der Kamp, M. W.; Mulholland, A. J.; Gervasio, F. L. A Multiscale Simulation Approach to Modeling Drug–Protein Binding Kinetics. J. Chem. Theory Comput. 2018, 14 (11), 60936101,  DOI: 10.1021/acs.jctc.8b00687
    58. 58
      Hamann, L. G.; Manfredi, M. C.; Sun, C.; Krystek, S. R.; Huang, Y.; Bi, Y.; Augeri, D. J.; Wang, T.; Zou, Y.; Betebenner, D. A.; Fura, A.; Seethala, R.; Golla, R.; Kuhns, J. E.; Lupisella, J. A.; Darienzo, C. J.; Custer, L. L.; Price, J. L.; Johnson, J. M.; Biller, S. A.; Zahler, R.; Ostrowski, J. Tandem Optimization of Target Activity and Elimination of Mutagenic Potential in a Potent Series of N-Aryl Bicyclic Hydantoin-Based Selective Androgen Receptor Modulators. Bioorg. Med. Chem. Lett. 2007, 17 (7), 18601864,  DOI: 10.1016/j.bmcl.2007.01.076
    59. 59
      Ostrowski, J.; Kuhns, J. E.; Lupisella, J. A.; Manfredi, M. C.; Beehler, B. C.; Krystek, S. R., Jr; Bi, Y.; Sun, C.; Seethala, R.; Golla, R.; Sleph, P. G.; Fura, A.; An, Y.; Kish, K. F.; Sack, J. S.; Mookhtiar, K. A.; Grover, G. J.; Hamann, L. G. Pharmacological and X-Ray Structural Characterization of a Novel Selective Androgen Receptor Modulator: Potent Hyperanabolic Stimulation of Skeletal Muscle with Hypostimulation of Prostate in Rats. Endocrinology 2007, 148 (1), 412,  DOI: 10.1210/en.2006-0843
    60. 60
      Matter, H.; Scheiper, B.; Steinhagen, H.; Böcskei, Z.; Fleury, V.; McCort, G. Structure-Based Design and Optimization of Potent Renin Inhibitors on 5- or 7-Azaindole-Scaffolds. Bioorg. Med. Chem. Lett. 2011, 21 (18), 54875492,  DOI: 10.1016/j.bmcl.2011.06.112
    61. 61
      He, S.; Senter, T. J.; Pollock, J.; Han, C.; Upadhyay, S. K.; Purohit, T.; Gogliotti, R. D.; Lindsley, C. W.; Cierpicki, T.; Stauffer, S. R.; Grembecka, J. High-Affinity Small-Molecule Inhibitors of the Menin-Mixed Lineage Leukemia (MLL) Interaction Closely Mimic a Natural Protein–Protein Interaction. J. Med. Chem. 2014, 57 (4), 15431556,  DOI: 10.1021/jm401868d
    62. 62
      Baum, B.; Muley, L.; Smolinski, M.; Heine, A.; Hangauer, D.; Klebe, G. Non-Additivity of Functional Group Contributions in Protein–Ligand Binding: A Comprehensive Study by Crystallography and Isothermal Titration Calorimetry. J. Mol. Biol. 2010, 397 (4), 10421054,  DOI: 10.1016/j.jmb.2010.02.007
    63. 63
      Ghosh, A. K.; Takayama, J.; Aubin, Y.; Ratia, K.; Chaudhuri, R.; Baez, Y.; Sleeman, K.; Coughlin, M.; Nichols, D. B.; Mulhearn, D. C.; Prabhakar, B. S.; Baker, S. C.; Johnson, M. E.; Mesecar, A. D. Structure-Based Design, Synthesis, and Biological Evaluation of a Series of Novel and Reversible Inhibitors for the Severe Acute Respiratory Syndrome- Coronavirus Papain-Like Protease. J. Med. Chem. 2009, 52 (16), 52285240,  DOI: 10.1021/jm900611t
    64. 64
      Ratia, K.; Pegan, S.; Takayama, J.; Sleeman, K.; Coughlin, M.; Baliji, S.; Chaudhuri, R.; Fu, W.; Prabhakar, B. S.; Johnson, M. E.; Baker, S. C.; Ghosh, A. K.; Mesecar, A. D. A Noncovalent Class of Papain-like Protease/deubiquitinase Inhibitors Blocks SARS Virus Replication. Proc. Natl. Acad. Sci. U. S. A. 2008, 105 (42), 1611916124,  DOI: 10.1073/pnas.0805240105
    65. 65
      Maestro, version 9.1; Schrödinger, LLC: New York, 2010.
    66. 66
      Wang, J.; Wang, W.; Kollman, P. A.; Case, D. A. Antechamber: An Accessory Software Package for Molecular Mechanical Calculations. J. Am. Chem. Soc. 2001, 222, U403
    67. 67
      Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25 (9), 11571174,  DOI: 10.1002/jcc.20035
    68. 68
      Jakalian, A.; Jack, D. B.; Bayly, C. I. Fast, Efficient Generation of High-Quality Atomic Charges. AM1-BCC Model: II. Parameterization and Validation. J. Comput. Chem. 2002, 23 (16), 16231641,  DOI: 10.1002/jcc.10128
    69. 69
      Naden, L. N.; Shirts, M. R. Linear Basis Function Approach to Efficient Alchemical Free Energy Calculations. 2. Inserting and Deleting Particles with Coulombic Interactions. J. Chem. Theory Comput. 2015, 11 (6), 25362549,  DOI: 10.1021/ct501047e
    70. 70
      Pechlaner, M.; Reif, M. M.; Oostenbrink, C. Reparametrisation of United-Atom Amine Solvation in the GROMOS Force Field. Mol. Phys. 2017, 115 (9–12), 11441154,  DOI: 10.1080/00268976.2016.1255797
    71. 71
      Diem, M.; Oostenbrink, C. Hamiltonian Reweighing To Refine Protein Backbone Dihedral Angle Parameters in the GROMOS Force Field. J. Chem. Inf. Model. 2020, 60 (1), 279288,  DOI: 10.1021/acs.jcim.9b01034
    72. 72
      Chodera, J.; Rizzi, A.; Naden, L.; Beauchamp, K.; Grinaway, P.; Fass, J.; Rustenburg, B.; Ross, G. A.; Simmonett, A.; Swenson, D. W. H. Choderalab/openmmtools: 0.14. 0-Exact Treatment of Alchemical PME Electrostatics, Water Cluster Test System, Optimizations. https://zenodo.org/record/1161149#.X0gc6HlKjIU.
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jcim.0c00610.

    • Details of protein preparation specific to each system; all ligand structures in full 3D with all optimized hydrogens labeled explicitly (Table S1); calculations for validation of free energies obtained from final optimized sigmas (Figures S1–S3 and Table S2); all convergence graphs not presented in the main text for ΔΔGscan calculations (Figures S4–S11); diagram for steric clashes in androgen receptor (Figure S12); all optimized sigmas for the menin B ligand (Table S3); all optimized charges for the thrombin B ligand (Table S4); Figure S13 depicting the thermodynamic cycle used for relative binding free energy calculations in this work (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect