Ensemble-Based Replica Exchange Alchemical Free Energy Methods: The Effect of Protein Mutations on Inhibitor Binding
- Agastya P. BhatiAgastya P. BhatiCentre for Computational Science, Department of Chemistry, University College London, 20 Gordon Street, London, WC1H 0AJ, United KingdomMore by Agastya P. Bhati
- ,
- Shunzhou WanShunzhou WanCentre for Computational Science, Department of Chemistry, University College London, 20 Gordon Street, London, WC1H 0AJ, United KingdomMore by Shunzhou Wan
- , and
- Peter V. Coveney*Peter V. Coveney*E-mail: [email protected]. Phone: +44 (0)20 7679 4802.Centre for Computational Science, Department of Chemistry, University College London, 20 Gordon Street, London, WC1H 0AJ, United KingdomMore by Peter V. Coveney
Abstract

The accurate prediction of the binding affinity changes of drugs caused by protein mutations is a major goal in clinical personalized medicine. We have developed an ensemble-based free energy approach called thermodynamic integration with enhanced sampling (TIES), which yields accurate, precise, and reproducible binding affinities. TIES has been shown to perform well for predictions of free energy differences of congeneric ligands to a wide range of target proteins. We have recently introduced variants of TIES, which incorporate the enhanced sampling technique REST2 (replica exchange with solute tempering) and the free energy estimator MBAR (Bennett acceptance ratio). Here we further extend the TIES methodology to study relative binding affinities caused by protein mutations when bound to a ligand, a variant which we call TIES-PM. We apply TIES-PM to fibroblast growth factor receptor 3 (FGFR3) to investigate binding free energy changes upon protein mutations. The results show that TIES-PM with REST2 successfully captures a large conformational change and generates correct free energy differences caused by a gatekeeper mutation located in the binding pocket. Simulations without REST2 fail to overcome the energy barrier between the conformations, and hence the results are highly sensitive to the initial structures. We also discuss situations where REST2 does not improve the accuracy of predictions.
1. Introduction
Figure 1

Figure 1. Structures of FGFR3 and inhibitors studied in this work: (a) the binding site of tyrosine kinase domain for FGFR3 in complex with ACP, an ATP-analogue (PDB ID: 4K33). FGFR3 is depicted in cartoon and ACP in bond representation. Mutations of three residues, V555, I538, and N540 (ball-and-stick representation), are among the most common genetic variants in FGFR3. The chemical structures of four ATP competitive inhibitors are shown in panels b–e: (b) AZD4547, (c) BGJ-398, (d) TKI258, and (e) JNJ4275649.
ΔΔGcalc | |||||
---|---|---|---|---|---|
mutant | drug | TIES | λR2b | λR2-Mb | ΔΔGexp |
V555M | AZD4547-linear | –3.56(0.31) | –2.76(0.12) | –2.70(0.12) | –1.75(0.33) |
AZD4547-bent | 0.55(0.41) | –2.07(0.11) | –1.98(0.12) | ||
BGJ-398 | –3.02(0.44) | –3.66(0.12) | –3.60(0.12) | –1.19(0.08) | |
TKI258 | 0.26(0.25) | –1.17(0.13) | –1.11(0.13) | 0.97(0.22) | |
JNJ42756493 | –5.19(0.38) | –3.99(0.16) | –3.92(0.15) | –3.08(0.17) | |
MAE | 1.75 | 1.37 | 1.30 | ||
RMSE | 1.84 | 1.59 | 1.54 | ||
I538V | AZD4547-linear | 0.25(0.33) | 0.09(0.11) | 0.05(0.11) | –2.11(0.32) |
BGJ-398 | 0.44(0.35) | 0.46(0.11) | 0.45(0.11) | –0.74(0.21) | |
TKI258 | –0.65(0.38) | 0.47(0.13) | 0.38(0.12) | –1.91(0.13) | |
JNJ42756493 | 0.62(0.34) | 0.30(0.12) | 0.28(0.12) | –2.18(0.10) | |
MAE | 1.90 | 2.06 | 2.02 | ||
RMSE | 2.02 | 2.13 | 2.08 | ||
N540S | AZD4547-linear | –0.43(0.43) | 0.91(0.14) | 0.95(0.14) | –0.76(0.33) |
BGJ-398 | –1.00(0.52) | 1.13(0.14) | 1.16(0.13) | 0.25(0.19) | |
TKI258 | –1.77(0.60) | 1.02(0.14) | 1.11(0.14) | –0.90(0.15) | |
JNJ42756493 | –0.87(0.45) | 1.06(0.14) | 1.11(0.14) | –1.75(0.21) | |
MAE | 0.83 | 1.82 | 1.87 | ||
RMSE | 0.89 | 1.94 | 2.00 |
The mean absolute error (MAE) and root mean square error (RMSE) are also shown for all complexes of each mutant using each free energy scheme. Production runs are 4 ns in all cases. All values are in kcal/mol. The statistical uncertainties associated with each value are shown in brackets.
Highest Teff for λ-REST2 simulations is 800 K for receptor and 1500 K for complexes in case of mutants I538V and N540S. In the case of mutant V555M, it is 1500 K for the AZD4547 complexes and 600 K for all other complexes; 600 K is used for the receptor.
2. Methods
2.1. Hybrid Topology
Figure 2

Figure 2. Different regions in the λ-REST2 simulations. The AZD4547-V555M complex is shown here as an example. The hybrid residue, denoted as the alchemical region, is depicted as a ball-and-stick model. It consists of disappearing (red) and appearing (blue) groups which are slightly separated for reasons of clarity. They can fully or partially overlap in the simulation as there are no interactions between them. The REST2 region, including the alchemical region (red and blue ball-and-stick), part of the ligand (orange bond), and surrounding protein residues (orange stick), is designated as the “hot” region. The selection of the REST2 region is described in the main text (section 2.3 REST2 region).
2.2. Free Energy Schemes

2.3. REST2 Region
2.4. Simulation Setup
2.5. Simulations
2.6. Computational Resources
3. Results
3.1. Local Mutation
Figure 3

Figure 3. Comparison of the predicted ΔΔGcalc values using TIES (black circles), TIES-λ-REST2 (λR2, up/down triangles) with those from experiments for V555M mutant complexes with the highest Teff of the chosen REST2 region at 600 K for receptor and complexes except those with AZD4547 which are at 1500 K (red triangles pointing up), and at 1500 K for receptor and 3000 K for complexes (blue triangles pointing down). Results of AZD4547 from the bent conformation are represented using filled circles and triangles. The dotted lines (x = 0 and y = 0) create four quadrants. Data points in quadrants I (x > 0 and y > 0) and III (x < 0 and y < 0) indicate that the calculated binding free energy differences agree directionally with those from the experimental data. The results from TIES-λ-REST2-M (λR2-M) are very close to those from λR2 (Table 1), and are not shown for reasons of clarity.
ΔΔGcalc | ||||||
---|---|---|---|---|---|---|
mutant | drug | λR2b | λR2-Mb | λR2c | λR2-Mc | ΔΔGexp |
V555M | AZD4547-linear | –2.76(0.12) | –2.70(0.12) | –1.85(0.07) | –1.82(0.06) | –1.75(0.33) |
AZD4547-bent | –2.07(0.11) | –1.98(0.12) | –1.07(0.07) | –1.11(0.06) | ||
BGJ-398 | –3.66(0.12) | –3.60(0.12) | –1.92(0.08) | –1.96(0.06) | –1.19(0.08) | |
TKI258 | –1.17(0.13) | –1.11(0.13) | –1.41(0.08) | –1.42(0.06) | 0.97(0.22) | |
JNJ42756493 | –3.99(0.16) | –3.92(0.15) | –2.88(0.12) | –2.87(0.11) | –3.08(0.17) | |
MAE | 1.37 | 1.30 | 0.82 | 0.82 | ||
RMSE | 1.59 | 1.54 | 1.16 | 1.16 | ||
I538V | AZD4547-linear | 0.09(0.11) | 0.05(0.11) | –0.12(0.08) | –0.04(0.07) | –2.11(0.32) |
BGJ-398 | 0.46(0.11) | 0.45(0.11) | 0.01(0.08) | 0.09(0.07) | –0.74(0.21) | |
TKI258 | 0.47(0.13) | 0.38(0.12) | 0.01(0.08) | 0.12(0.08) | –1.91(0.13) | |
JNJ42756493 | 0.30(0.12) | 0.28(0.12) | –0.01(0.07) | 0.11(0.07) | –2.18(0.10) | |
MAE | 2.06 | 2.02 | 1.71 | 1.80 | ||
RMSE | 2.13 | 2.08 | 1.80 | 1.89 | ||
N540S | AZD4547-linear | 0.91(0.14) | 0.95(0.14) | 0.72(0.11) | 0.74(0.11) | –0.76(0.33) |
BGJ-398 | 1.13(0.14) | 1.16(0.13) | 0.67(0.11) | 0.67(0.11) | 0.25(0.19) | |
TKI258 | 1.02(0.14) | 1.11(0.14) | 0.71(0.12) | 0.72(0.12) | –0.9(0.15) | |
JNJ42756493 | 1.06(0.14) | 1.11(0.14) | 0.72(0.12) | 0.72(0.12) | –1.75(0.21) | |
MAE | 1.82 | 1.87 | 1.50 | 1.50 | ||
RMSE | 1.94 | 2.00 | 1.66 | 1.67 |
The mean absolute error (MAE) and root mean square error (RMSE) for all complexes of each mutant using each free energy scheme are also shown. Production runs are 4 ns in all cases. All values are in kcal/mol. Statistical uncertainties associated with each value are shown in the brackets.
Highest Teff for λ-REST2 simulations is 800 K for receptor and 1500 K for complexes in case of mutants I538V and N540S. In the case of mutant V555M, it is 1500 K for the AZD4547 complexes and 600 K for all other complexes; 600 K is used for the receptor.
Highest Teff for λ-REST2 simulations is 1500 K for receptor and 3000 K for complexes.
3.2. Remote Mutations
Figure 4

Figure 4. Comparison of the predicted ΔΔGcalc values using TIES (black circles), TIES-λ-REST2 (λR2, up/down triangles) with those from experiment for for all inhibitors bound to FGFR3: (a) I538V mutant and (b) N540S mutant, when the highest Teff for the chosen REST2 region is at 800 K for receptor and 1500 K for complexes (red triangles pointing up) and at 1500 K for receptor and 3000 K for complexes (blue triangles pointing down). The dotted lines (x = 0 and y = 0) create four quadrants. Data points in quadrants I (x > 0 and y > 0) and III (x < 0 and y < 0) indicate that the ΔΔGcalc values agree directionally with ΔΔGexp. The results from TIES-λ-REST2-M (λR2-M) are very close to those from λR2 (Table 1), and are not shown for reasons of clarity.
3.3. Effect of Extending Simulation Time
Figure 5

Figure 5. Variation of the cumulative average of ΔΔGcalcTIES with simulation length for all inhibitors bound to the FGFR3 I538V mutant. The corresponding experimental value for each inhibitor is shown by a dashed line of the same color.
Figure 6

Figure 6. Variation of the cumulative average of ΔΔG calculated using schemes λR2 and λR2-M with simulation length for all complexes. The highest Teff for receptor and complex are 1500 and 3000 K, respectively, in the case of I538V and N540S mutants, while the corresponding values in the case of V555M mutant are 600 and 600 K/1500 K. The corresponding experimental values for each inhibitor are shown by a dashed line of the same color.
4. Discussion
4.1. Improved Sampling of AZD4547 on “Heating”
Figure 7

Figure 7. Two distinct conformations of inhibitor AZD4547 found experimentally when bound to the FGFR gatekeeper mutant. The three hydrogen bonds, marked with black dashed lines and labeled as H1, H2, and H3, keep the middle portion of the inhibitor stable. The value of the dihedral angle between the four carbon atoms highlighted in orange can be used as an indicator of the occurrence of the two conformations. The atoms displayed as balls lie in the REST2 region while the ones displayed as lines reside outside it.
Figure 8

Figure 9

Figure 9. Normalized frequency distributions of the dihedral angle between the four carbon atoms highlighted in orange in Figure 7 for different λ states of V555M–AZD4547 complexes in standard TIES (in blue) as well as λ-REST2 simulations showing the relative populations of the two conformations of AZD4547. In the case of λ-REST2 simulations, the distributions from the first (1–4 ns; in black) and the last 4 ns (17–20 ns; in red) are shown separately.
4.2. The Exceptional Case of TKI258: Limitations of λ-REST2
Figure 10

Figure 10. Inhibitor TKI258 bound to V555M mutant. It forms two hydrogen bonds with the hinge region of the protein which are displayed with black dashed lines and labeled as H1 and H2. The atoms shown as balls lie in the “hot” region. The atoms are shown in the standard color code: carbon in green, oxygen in red, nitrogen in blue, hydrogen in white, and fluorine in pink.
Figure 11

Figure 11. Normalized frequency distributions of H1 and H2 from Figure 10 for the two end-points in the case of the standard TIES as well as λ-REST2 simulations of the V555M–TKI258 complex. λ-REST2 simulations sample a larger comformational space than TIES, as evidenced by the lower and wider distributions of the distances, and the second peaks in the λ = 1 end-point (V555M).
Figure 12

Figure 12. Normalized frequency distributions of H1 and H2 from Figure 10 for the two end-points in the case of standard TIES as well as λ-REST2 simulations of the I538V–TKI258 complex. The long tails and additional peaks beyond 4 Å indicate that λ-REST2 simulations sample some conformations irrelevant to stable inhibitor binding.
Figure 13

Figure 13. Normalized frequency distributions of H1 and H2 from Figure 10 for the two end-points in the case of standard TIES as well as λ-REST2 simulations of the N540S–TKI258 complex. The long tails and additional peaks beyond 4 Å indicate that λ-REST2 simulations sample some conformations irrelevant to stable inhibitor binding.
5. Conclusions
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jctc.8b01118.
Tables containing individual ΔGalch values for all the relative free energy calculations, and a table for the comparison of occupancies of hydrogen bonds between the inhibitors and the protein (PDF)
Topology and coordinate files for all ligand–protein complexes (ZIP)
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.
Acknowledgments
We thank our colleague Dr. David W. Wright for useful discussions.
References
This article references 33 other publications.
- 1Wright, D. W.; Wan, S.; Shublaq, N.; Zasada, S. J.; Coveney, P. V. From base pair to bedside: molecular simulation and the translation of genomics to personalized medicine. WIREs Syst. Biol. Med. 2012, 4, 585– 598, DOI: 10.1002/wsbm.1186Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXpslSgsg%253D%253D&md5=b6bf5cb6b9fab33543986b9f8ceed6ecFrom base pair to bedside: molecular simulation and the translation of genomics to personalized medicineWright, David W.; Wan, Shunzhou; Shublaq, Nour; Zasada, Stefan J.; Coveney, Peter V.Wiley Interdisciplinary Reviews: Systems Biology and Medicine (2012), 4 (6), 585-598CODEN: WIRSBW; ISSN:1939-005X. (Wiley-Blackwell)A review. Despite the promises made that genomic sequencing would transform therapy by introducing a new era of personalized medicine, relatively few tangible breakthroughs have been made. This has led to the recognition that complex interactions at multiple spatial, temporal, and organizational levels may often combine to produce disease. Understanding this complexity requires that existing and future models are used and interpreted within a framework that incorporates knowledge derived from investigations at multiple levels of biol. function. It also requires a computational infrastructure capable of dealing with the vast quantities of data generated by genomic approaches. In this review, we discuss the use of mol. modeling to generate quant. and qual. insights at the smallest scales of the systems biol. hierarchy, how it can play an important role in the development of a systems understanding of disease and in the application of such knowledge to help discover new therapies and target existing ones on a personal level.
- 2Wang, L. 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, 2695– 2703, DOI: 10.1021/ja512751qGoogle Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsF2iuro%253D&md5=37a4f4a6c085f47ed531342643b6c33bAccurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force FieldWang, Lingle; Wu, Yujie; Deng, Yuqing; Kim, Byungchan; Pierce, Levi; Krilov, Goran; Lupyan, Dmitry; Robinson, Shaughnessy; Dahlgren, Markus K.; Greenwood, Jeremy; Romero, Donna L.; Masse, Craig; Knight, Jennifer L.; Steinbrecher, Thomas; Beuming, Thijs; Damm, Wolfgang; Harder, Ed; Sherman, Woody; Brewer, Mark; Wester, Ron; Murcko, Mark; Frye, Leah; Farid, Ramy; Lin, Teng; Mobley, David L.; Jorgensen, William L.; Berne, Bruce J.; Friesner, Richard A.; Abel, RobertJournal of the American Chemical Society (2015), 137 (7), 2695-2703CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Designing tight-binding ligands is a primary objective of small-mol. drug discovery. Over the past few decades, free-energy calcns. have benefited from improved force fields and sampling algorithms, as well as the advent of low-cost parallel computing. However, it has proven to be challenging to reliably achieve the level of accuracy that would be needed to guide lead optimization (∼5× in binding affinity) for a wide range of ligands and protein targets. Not surprisingly, widespread com. application of free-energy simulations has been limited due to the lack of large-scale validation coupled with the tech. challenges traditionally assocd. with running these types of calcns. Here, we report an approach that achieves an unprecedented level of accuracy across a broad range of target classes and ligands, with retrospective results encompassing 200 ligands and a wide variety of chem. perturbations, many of which involve significant changes in ligand chem. structures. In addn., we have applied the method in prospective drug discovery projects and found a significant improvement in the quality of the compds. synthesized that have been predicted to be potent. Compds. predicted to be potent by this approach have a substantial redn. in false positives relative to compds. synthesized on the basis of other computational or medicinal chem. approaches. Furthermore, the results are consistent with those obtained from our retrospective studies, demonstrating the robustness and broad range of applicability of this approach, which can be used to drive decisions in lead optimization.
- 3Bernardi, R. C.; Melo, M. C.; Schulten, K. Enhanced sampling techniques in molecular dynamics simulations of biological systems. Biochim. Biophys. Acta, Gen. Subj. 2015, 1850, 872– 877, DOI: 10.1016/j.bbagen.2014.10.019Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVanurbN&md5=44e3ac1aa042c35ee08aa86a6a63e78eEnhanced sampling techniques in molecular dynamics simulations of biological systemsBernardi, Rafael C.; Melo, Marcelo C. R.; Schulten, KlausBiochimica et Biophysica Acta, General Subjects (2015), 1850 (5), 872-877CODEN: BBGSB3; ISSN:0304-4165. (Elsevier B.V.)A review. Mol. dynamics has emerged as an important research methodol. covering systems to the level of millions of atoms. However, insufficient sampling often limits its application. The limitation is due to rough energy landscapes, with many local min. sepd. by high-energy barriers, which govern the biomol. motion. In the past few decades methods have been developed that address the sampling problem, such as replica-exchange mol. dynamics, metadynamics and simulated annealing. Here the authors present an overview over theses sampling methods in an attempt to shed light on which should be selected depending on the type of system property studied. Enhanced sampling methods have been employed for a broad range of biol. systems and the choice of a suitable method is connected to biol. and phys. characteristics of the system, in particular system size. While metadynamics and replica-exchange mol. dynamics are the most adopted sampling methods to study biomol. dynamics, simulated annealing is well suited to characterize very flexible systems. The use of annealing methods for a long time was restricted to simulation of small proteins; however, a variant of the method, generalized simulated annealing, can be employed at a relatively low computational cost to large macromol. complexes. Mol. dynamics trajectories frequently do not reach all relevant conformational substates, for example those connected with biol. function, a problem that can be addressed by employing enhanced sampling algorithms. This article is part of a Special Issue entitled Recent developments of mol. dynamics.
- 4Valsson, O.; Parrinello, M. Variational approach to enhanced sampling and free energy calculations. Phys. Rev. Lett. 2014, 113, 090601, DOI: 10.1103/PhysRevLett.113.090601Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhslShtL7O&md5=4d66ba6f0d445e693e28dbeae09cd936Variational approach to enhanced sampling and free energy calculationsValsson, Omar; Parrinello, MichelePhysical Review Letters (2014), 113 (9), 090601CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)The ability of widely used sampling methods, such as mol. dynamics or Monte Carlo simulations, to explore complex free energy landscapes is severely hampered by the presence of kinetic bottlenecks. A large no. of solns. have been proposed to alleviate this problem. Many are based on the introduction of a bias potential which is a function of a small no. of collective variables. However constructing such a bias is not simple. Here we introduce a functional of the bias potential and an assocd. variational principle. The bias that minimizes the functional relates in a simple way to the free energy surface. This variational principle can be turned into a practical, efficient, and flexible sampling method. A no. of numerical examples are presented which include the detn. of a three-dimensional free energy surface. We argue that, beside being numerically advantageous, our variational approach provides a convenient and novel standpoint for looking at the sampling problem.
- 5Fukunishi, H.; Watanabe, O.; Takada, S. On the Hamiltonian replica exchange method for efficient sampling of biomolecular systems: Application to protein structure prediction. J. Chem. Phys. 2002, 116, 9058– 9067, DOI: 10.1063/1.1472510Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XjsFKmsLo%253D&md5=7ac571a5afdd63b0b4b29cfdec06f53bOn the Hamiltonian replica exchange method for efficient sampling of biomolecular systems: Application to protein structure predictionFukunishi, Hiroaki; Watanabe, Osamu; Takada, ShojiJournal of Chemical Physics (2002), 116 (20), 9058-9067CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Motivated by the protein structure prediction problem, we develop two variants of the Hamiltonian replica exchange methods (REMs) for efficient configuration sampling, (1) the scaled hydrophobicity REM and (2) the phantom chain REM, and compare their performance with the ordinary REM. We first point out that the ordinary REM has a shortage for the application to large systems such as biomols. and that the Hamiltonian REM, an alternative formulation of the REM, can give a remedy for it. We then propose two examples of the Hamiltonian REM that are suitable for a coarse-grained protein model. (1) The scaled hydrophobicity REM preps. replicas that are characterized by various strengths of hydrophobic interaction. The strongest interaction that mimics aq. soln. environment makes proteins folding, while weakened hydrophobicity unfolds proteins as in org. solvent. Exchange between these environments enables proteins to escape from misfolded traps and accelerate conformational search. This resembles the roles of mol. chaperone that assist proteins to fold in vivo. (2) The phantom chain REM uses replicas that allow various degrees of at. overlaps. By allowing at. overlap in some of replicas, the peptide chain can cross over itself, which can accelerate conformation sampling. Using a coarse-gained model we developed, we compute equil. probability distributions for poly-alanine 16-mer and for a small protein by these REMs and compare the accuracy of the results. We see that the scaled hydrophobicity REM is the most efficient method among the three REMs studied.
- 6Wang, L.; Friesner, R. A.; Berne, B. J. Replica exchange with solute scaling: A more efficient version of replica exchange with solute tempering (REST2). J. Phys. Chem. B 2011, 115, 9431– 9438, DOI: 10.1021/jp204407dGoogle Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXosFymurs%253D&md5=8acb0d670bcb70eacfc30e32d5dfb6ddReplica Exchange with Solute Scaling: A More Efficient Version of Replica Exchange with Solute Tempering (REST2)Wang, Lingle; Friesner, Richard A.; Berne, B. J.Journal of Physical Chemistry B (2011), 115 (30), 9431-9438CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)A small change in the Hamiltonian scaling in Replica Exchange with Solute Tempering (REST) is found to improve its sampling efficiency greatly, esp. for the sampling of aq. protein solns. in which there are large-scale solute conformation changes. Like the original REST (REST1), the new version (which the authors call REST2) also bypasses the poor scaling with system size of the std. Temp. Replica Exchange Method (TREM), reducing the no. of replicas (parallel processes) from what must be used in TREM. This redn. is accomplished by deforming the Hamiltonian function for each replica in such a way that the acceptance probability for the exchange of replica configurations does not depend on the no. of explicit water mols. in the system. For proof of concept, REST2 is compared with TREM and with REST1 for the folding of the trpcage and β-hairpin in water. The comparisons confirm that REST2 greatly reduces the no. of CPUs required by regular replica exchange and greatly increases the sampling efficiency over REST1. This method reduces the CPU time required for calcg. thermodn. avs. and for the ab initio folding of proteins in explicit water.
- 7Wang, L.; Berne, B. J.; Friesner, R. A. On achieving high accuracy and reliability in the calculation of relative protein–ligand binding affinities. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 1937– 1942, DOI: 10.1073/pnas.1114017109Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XivVSht74%253D&md5=23db7d18e624d5e1214bbd6702619a87On achieving high accuracy and reliability in the calculation of relative protein-ligand binding affinitiesWang, Lingle; Berne, B. J.; Friesner, Richard A.Proceedings of the National Academy of Sciences of the United States of America (2012), 109 (6), 1937-1942, S1937/1-S1937/7CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)We apply a free energy perturbation simulation method, free energy perturbation/replica exchange with solute tempering, to two modifications of protein-ligand complexes that lead to significant conformational changes, the first in the protein and the second in the ligand. The approach is shown to facilitate sampling in these challenging cases where high free energy barriers sep. the initial and final conformations and leads to superior convergence of the free energy as demonstrated both by consistency of the results (independence from the starting conformation) and agreement with exptl. binding affinity data. The second case, consisting of two neutral thrombin ligands that are taken from a recent medicinal chem. program for this interesting pharmaceutical target, is of particular significance in that it demonstrates that good results can be obtained for large, complex ligands, as opposed to relatively simple model systems. To achieve quant. agreement with expt. in the thrombin case, a next generation force field, Optimized Potentials for Liq. Simulations 2.0, is required, which provides superior charges and torsional parameters as compared to earlier alternatives.
- 8Paliwal, H.; Shirts, M. R. A benchmark test set for alchemical free energy transformations and its use to quantify error in common free energy methods. J. Chem. Theory Comput. 2011, 7, 4115– 4134, DOI: 10.1021/ct2003995Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtlaqtrvM&md5=41383373300216165a7f1724ad535451A Benchmark Test Set for Alchemical Free Energy Transformations and Its Use to Quantify Error in Common Free Energy MethodsPaliwal, Himanshu; Shirts, Michael R.Journal of Chemical Theory and Computation (2011), 7 (12), 4115-4134CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)There is a significant need for improved tools to validate thermophys. quantities computed via mol. simulation. In this paper we present the initial version of a benchmark set of testing methods for calcg. free energies of mol. transformation in soln. This set is based on mol. changes common to many mol. design problems, such as insertion and deletion of at. sites and changing at. partial charges. We use this benchmark set to compare the statistical efficiency, reliability, and quality of uncertainty ests. for a no. of published free energy methods, including thermodn. integration, free energy perturbation, the Bennett acceptance ratio (BAR) and its multistate equiv. MBAR. We identify MBAR as the consistently best performing method, though other methods are frequently comparable in reliability and accuracy in many cases. We demonstrate that assumptions of Gaussian distributed errors in free energies are usually valid for most methods studied. We demonstrate that bootstrap error estn. is a robust and useful technique for estg. statistical variance for all free energy methods studied. This benchmark set is provided in a no. of different file formats with the hope of becoming a useful and general tool for method comparisons.
- 9Shirts, M. R.; Chodera, J. D. Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 2008, 129, 124105, DOI: 10.1063/1.2978177Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXht1WnsL7F&md5=479183e1f45fc58dd7c6e5ef1e73d45dStatistically optimal analysis of samples from multiple equilibrium statesShirts, Michael R.; Chodera, John D.Journal of Chemical Physics (2008), 129 (12), 124105/1-124105/10CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We present a new estimator for computing free energy differences and thermodn. expectations as well as their uncertainties from samples obtained from multiple equil. states via either simulation or expt. The estimator, which we call the multistate Bennett acceptance ratio estimator (MBAR) because it reduces to the Bennett acceptance ratio estimator (BAR) when only two states are considered, has significant advantages over multiple histogram reweighting methods for combining data from multiple states. It does not require the sampled energy range to be discretized to produce histograms, eliminating bias due to energy binning and significantly reducing the time complexity of computing a soln. to the estg. equations in many cases. Addnl., an est. of the statistical uncertainty is provided for all estd. quantities. In the large sample limit, MBAR is unbiased and has the lowest variance of any known estimator for making use of equil. data collected from multiple states. We illustrate this method by producing a highly precise est. of the potential of mean force for a DNA hairpin system, combining data from multiple optical tweezer measurements under const. force bias. (c) 2008 American Institute of Physics.
- 10Bhati, A. P.; Wan, S.; Wright, D. W.; Coveney, P. V. Rapid, accurate, precise, and reliable relative free energy prediction using ensemble based thermodynamic integration. J. Chem. Theory Comput. 2017, 13, 210– 222, DOI: 10.1021/acs.jctc.6b00979Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitVGntrrJ&md5=510b70188112a3578030e291ce19b127Rapid, Accurate, Precise, and Reliable Relative Free Energy Prediction Using Ensemble Based Thermodynamic IntegrationBhati, Agastya P.; Wan, Shunzhou; Wright, David W.; Coveney, Peter V.Journal of Chemical Theory and Computation (2017), 13 (1), 210-222CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The accurate prediction of the binding affinities of ligands to proteins is a major goal in drug discovery and personalized medicine. The time taken to make such predictions is of similar importance to their accuracy, precision and reliability. In the last few years, an ensemble based mol. dynamics approach has been proposed that provides a route to reliable predictions of free energies based on the mol. mechanics Poisson-Boltzmann surface area method which meets the requirements of accuracy, precision and reliability. Here, we describe an equiv. methodol. based on thermodn. integration to substantially improve the accuracy, precision and reliability of calcd. relative binding free energies. We report the performance of the method when applied to a diverse set of protein targets and ligands. The results are in very good agreement with exptl. data (90% of calcns. agree to within 1 kcal/mol) while the method is reproducible by construction. Statistical uncertainties of the order of 0.5 kcal/mol or less are achieved. We present a systematic account of how the uncertainty in the predictions may be estd.
- 11Bhati, A. P.; Wan, S.; Hu, Y.; Sherborne, B.; Coveney, P. V. Uncertainty quantification in alchemical free energy methods. J. Chem. Theory Comput. 2018, 14, 2867– 2880, DOI: 10.1021/acs.jctc.7b01143Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXotFaqtr4%253D&md5=5f80e23df9c8aafb58914af06de24e1fUncertainty Quantification in Alchemical Free Energy MethodsBhati, Agastya P.; Wan, Shunzhou; Hu, Yuan; Sherborne, Brad; Coveney, Peter V.Journal of Chemical Theory and Computation (2018), 14 (6), 2867-2880CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Alchem. free energy methods have gained much importance recently from several reports of improved ligand-protein binding affinity predictions based on their implementation using mol. dynamics simulations. A large no. of variants of such methods implementing different accelerated sampling techniques and free energy estimators are available, each claimed to be better than the others in its own way. However, the key features of reproducibility and quantification of assocd. uncertainties in such methods have barely been discussed. Here, we apply a systematic protocol for uncertainty quantification to a no. of popular alchem. free energy methods, covering both abs. and relative free energy predictions. We show that a reliable measure of error estn. is provided by ensemble simulation-an ensemble of independent MD simulations-which applies irresp. of the free energy method. The need to use ensemble methods is fundamental and holds regardless of the duration of time of the mol. dynamics simulations performed.
- 12Wan, S.; Bhati, A. P.; Skerratt, S.; Omoto, K.; Shanmugasundaram, V.; Bagal, S. K.; Coveney, P. V. Evaluation and characterization of Trk kinase inhibitors for the treatment of pain: reliable binding affinity predictions from theory and computation. J. Chem. Inf. Model. 2017, 57, 897– 909, DOI: 10.1021/acs.jcim.6b00780Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXksVyrurc%253D&md5=2e366f56c1ee8c2f2b2760b9098e2030Evaluation and Characterization of Trk Kinase Inhibitors for the Treatment of Pain: Reliable Binding Affinity Predictions from Theory and ComputationWan, Shunzhou; Bhati, Agastya P.; Skerratt, Sarah; Omoto, Kiyoyuki; Shanmugasundaram, Veerabahu; Bagal, Sharan K.; Coveney, Peter V.Journal of Chemical Information and Modeling (2017), 57 (4), 897-909CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Optimization of ligand binding affinity to the target protein of interest is a primary objective in small-mol. drug discovery. Until now, the prediction of binding affinities by computational methods has not been widely applied in the drug discovery process, mainly due to its lack of accuracy and reproducibility, as well as the long turnaround times required to obtain results. Herein, the authors report on a collaborative study that compares tropomyosin receptor kinase A (TrkA) binding affinity predictions using two recently formulated fast computational approaches - namely ESMACS (Enhanced Sampling of Mol. dynamics with Approxn. of Continuum Solvent) and TIES (Thermodn. Integration with Enhanced Sampling) - to exptl. derived TrkA binding affinities for a set of Pfizer pan-Trk compds. ESMACS gives precise and reproducible results and is applicable to highly diverse sets of compds. It also provides detailed chem. insight into the nature of ligand-protein binding. TIES can predict and thus optimize more subtle changes in binding affinities between compds. of similar structure. Individual binding affinities were calcd. in a few hours, exhibiting good correlations with the exptl. data of 0.79 and 0.88 from ESMACS and TIES approaches resp. The speed, level of accuracy and precision of the calcns. are such that the affinity predictions can be used to rapidly explain the effects of compd. modifications on TrkA binding affinity. The methods could therefore be used as tools to guide lead optimization efforts across multiple prospective structurally-enabled programs in the drug discovery setting for a wide range of compds. and targets.
- 13Wan, 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, 784– 795, DOI: 10.1021/acs.jctc.6b00794Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitFektLnF&md5=713436084662420482684fb50db0832eRapid and Reliable Binding Affinity Prediction of Bromodomain Inhibitors: A Computational StudyWan, Shunzhou; Bhati, Agastya P.; Zasada, Stefan J.; Wall, Ian; Green, Darren; Bamborough, Paul; Coveney, Peter V.Journal of Chemical Theory and Computation (2017), 13 (2), 784-795CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Binding free energies of bromodomain inhibitors are calcd. with recently formulated approaches, namely ESMACS (enhanced sampling of mol. dynamics with approxn. of continuum solvent) and TIES (thermodn. integration with enhanced sampling). A set of compds. is provided by GlaxoSmithKline, which represents a range of chem. functionality and binding affinities. The predicted binding free energies exhibit a good Spearman correlation of 0.78 with the exptl. data from the 3-trajectory ESMACS, and an excellent correlation of 0.92 from the TIES approach where applicable. Given access to suitable high end computing resources and a high degree of automation, the authors can compute individual binding affinities in a few hours with precisions no greater than 0.2 kcal/mol for TIES, and no larger than 0.34 kcal/mol and 1.71 kcal/mol for the 1- and 3-trajectory ESMACS approaches.
- 14Porta, R.; Borea, R.; Coelho, A.; Khan, S.; Araújo, A.; Reclusa, P.; Franchina, T.; Steen, N. V. D.; Dam, P. V.; Ferri, J.; Sirera, R.; Naing, A.; Hong, D.; Rolfo, C. FGFR a promising druggable target in cancer: Molecular biology and new drugs. Crit. Rev. Oncol. Hematol. 2017, 113, 256– 267, DOI: 10.1016/j.critrevonc.2017.02.018Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC1cvps1ajsA%253D%253D&md5=d03086e258264a12c8c5cc7f2c5de8e3FGFR a promising druggable target in cancer: Molecular biology and new drugsPorta Rut; Borea Roberto; Coelho Andreia; Reclusa Pablo; Van Dam Peter; Ferri Jose; Sirera Rafael; Khan Shahanavaj; Araujo Antonio; Franchina Tindara; Van Der Steen Nele; Naing Aung; Hong David; Rolfo ChristianCritical reviews in oncology/hematology (2017), 113 (), 256-267 ISSN:.INTRODUCTION: The Fibroblast Growth Factor Receptor (FGFR) family consists of Tyrosine Kinase Receptors (TKR) involved in several biological functions. Recently, alterations of FGFR have been reported to be important for progression and development of several cancers. In this setting, different studies are trying to evaluate the efficacy of different therapies targeting FGFR. AREAS COVERED: This review summarizes the current status of treatments targeting FGFR, focusing on the trials that are evaluating the FGFR profile as inclusion criteria: Multi-Target, Pan-FGFR Inhibitors and anti-FGF (Fibroblast Growth Factor)/FGFR Monoclonal Antibodies. EXPERT OPINION: Most of the TKR share intracellular signaling pathways; therefore, cancer cells tend to overcome the inhibition of one tyrosine kinase receptor by activating another. The future of TKI (Tyrosine Kinase Inhibitor) therapy will potentially come from multi-targeted TKIs that target different TKR simultaneously. It is crucial to understand the interaction of the FGF-FGFR axis with other known driver TKRs. Based on this, it is possible to develop therapeutic strategies targeting multiple connected TKRs at once. One correct step in this direction is the reassessment of multi target inhibitors considering the FGFR status of the tumor. Another opportunity arises from assessing the use of FGFR TKI on patients harboring FGFR alterations.
- 15Bunney, T. D.; Wan, S.; Thiyagarajan, N.; Sutto, L.; Williams, S. V.; Ashford, P.; Koss, H.; Knowles, M. a.; Gervasio, F. L.; Coveney, P. V.; Katan, M. The effect of mutations on drug sensitivity and kinase activity of fibroblast growth factor receptors: A combined experimental and theoretical study. EBioMedicine 2015, 2, 194– 204, DOI: 10.1016/j.ebiom.2015.02.009Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2srotlKmtA%253D%253D&md5=8414f3530d24bbba0ea8cc98cb784b96The Effect of Mutations on Drug Sensitivity and Kinase Activity of Fibroblast Growth Factor Receptors: A Combined Experimental and Theoretical StudyBunney Tom D; Thiyagarajan Nethaji; Ashford Paul; Katan Matilda; Wan Shunzhou; Coveney Peter V; Sutto Ludovico; Gervasio Francesco L; Williams Sarah V; Knowles Margaret A; Koss HansEBioMedicine (2015), 2 (3), 194-204 ISSN:.Fibroblast growth factor receptors (FGFRs) are recognized therapeutic targets in cancer. We here describe insights underpinning the impact of mutations on FGFR1 and FGFR3 kinase activity and drug efficacy, using a combination of computational calculations and experimental approaches including cellular studies, X-ray crystallography and biophysical and biochemical measurements. Our findings reveal that some of the tested compounds, in particular TKI258, could provide therapeutic opportunity not only for patients with primary alterations in FGFR but also for acquired resistance due to the gatekeeper mutation. The accuracy of the computational methodologies applied here shows a potential for their wider application in studies of drug binding and in assessments of functional and mechanistic impacts of mutations, thus assisting efforts in precision medicine.
- 16Patani, H.; Bunney, T. D.; Thiyagarajan, N.; Norman, R. A.; Ogg, D.; Breed, J.; Ashford, P.; Potterton, A.; Edwards, M.; Katan, M. Landscape of activating cancer mutations in FGFR kinases and their differential responses to inhibitors in clinical use. Oncotarget 2016, 7, 24252– 24268, DOI: 10.18632/oncotarget.8132Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28fhsFShsw%253D%253D&md5=beb65006a52ef6062bab9c9d8086bf1dLandscape of activating cancer mutations in FGFR kinases and their differential responses to inhibitors in clinical usePatani Harshnira; Bunney Tom D; Thiyagarajan Nethaji; Ashford Paul; Potterton Andrew; Edwards Mina; Pang Camilla S M; Orengo Christine; Katan Matilda; Norman Richard A; Ogg Derek; Breed Jason; Phillips Chris; Williams Sarah V; Knowles Margaret A; Thomson Gary S; Breeze Alexander LOncotarget (2016), 7 (17), 24252-68 ISSN:.Frequent genetic alterations discovered in FGFRs and evidence implicating some as drivers in diverse tumors has been accompanied by rapid progress in targeting FGFRs for anticancer treatments. Wider assessment of the impact of genetic changes on the activation state and drug responses is needed to better link the genomic data and treatment options. We here apply a direct comparative and comprehensive analysis of FGFR3 kinase domain variants representing the diversity of point-mutations reported in this domain. We reinforce the importance of N540K and K650E and establish that not all highly activating mutations (for example R669G) occur at high-frequency and conversely, that some "hotspots" may not be linked to activation. Further structural characterization consolidates a mechanistic view of FGFR kinase activation and extends insights into drug binding. Importantly, using several inhibitors of particular clinical interest (AZD4547, BGJ-398, TKI258, JNJ42756493 and AP24534), we find that some activating mutations (including different replacements of the same residue) result in distinct changes in their efficacy. Considering that there is no approved inhibitor for anticancer treatments based on FGFR-targeting, this information will be immediately translatable to ongoing clinical trials.
- 17Azam, M.; Seeliger, M. a.; Gray, N. S.; Kuriyan, J.; Daley, G. Q. Activation of tyrosine kinases by mutation of the gatekeeper threonine. Nat. Struct. Mol. Biol. 2008, 15, 1109– 1118, DOI: 10.1038/nsmb.1486Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXht1Sisb%252FM&md5=4afc4e5a0850a4510ced3fa5fbda97faActivation of tyrosine kinases by mutation of the gatekeeper threonineAzam, Mohammad; Seeliger, Markus A.; Gray, Nathanael S.; Kuriyan, John; Daley, George Q.Nature Structural & Molecular Biology (2008), 15 (10), 1109-1118CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Protein kinases targeted by small-mol. inhibitors develop resistance through mutation of the 'gatekeeper' threonine residue of the active site. Here we show that the gatekeeper mutation in the cellular forms of c-ABL, c-SRC, platelet-derived growth factor receptor-α and -β, and epidermal growth factor receptor activates the kinase and promotes malignant transformation of BaF3 cells. Structural anal. reveals that a network of hydrophobic interactions - the hydrophobic spine - characteristic of the active kinase conformation is stabilized by the gatekeeper substitution. Substitution of glycine for the residues constituting the spine disrupts the hydrophobic connectivity and inactivates the kinase. Furthermore, a small-mol. inhibitor (compd. 14) that maximizes complementarity with the dismantled spine inhibits the gatekeeper mutation of BCR-ABL-T315I. These results demonstrate that mutation of the gatekeeper threonine is a common mechanism of activation for tyrosine kinases and provide structural insights to guide the development of next-generation inhibitors.
- 18van Gunsteren, W. F.; Daura, X.; Hansen, N.; Mark, A. E.; Oostenbrink, C.; Riniker, S.; Smith, L. J. Validation of Molecular Simulation: An Overview of Issues. Angew. Chem., Int. Ed. 2018, 57, 884– 902, DOI: 10.1002/anie.201702945Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhs1ertQ%253D%253D&md5=d38b0f3fd0054e0fab46023a2c597442Validation of Molecular Simulation: An Overview of Issuesvan Gunsteren, Wilfred F.; Daura, Xavier; Hansen, Niels; Mark, Alan E.; Oostenbrink, Chris; Riniker, Sereina; Smith, Lorna J.Angewandte Chemie, International Edition (2018), 57 (4), 884-902CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Computer simulation of mol. systems enables structure-energy-function relationships of mol. processes to be described at the sub-at., at., supra-at., or supra-mol. level. To interpret results of such simulations appropriately, the quality of the calcd. properties must be evaluated. This depends on the way the simulations are performed and on the way they are validated by comparison to values Qexp of exptl. observable quantities Q. One must consider (1) the accuracy of Qexp, (2) the accuracy of the function Q(rN) used to calc. a Q-value based on a mol. configuration rN of N particles, (3) the sensitivity of the function Q(rN) to the configuration rN, (4) the relative time scales of the simulation and expt., (5) the degree to which the calcd. and exptl. properties are equiv., and (6) the degree to which the system simulated matches the exptl. conditions. Exptl. data is limited in scope and generally corresponds to avs. over both time and space. A crit. anal. of the various factors influencing the apparent degree of (dis)agreement between simulations and expt. is presented and illustrated using examples from the literature. What can be done to enhance the validation of mol. simulation is also discussed.
- 19Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kal, L.; Schulten, K. Scalable molecular dynamics with NAMD. J. Comput. Chem. 2005, 26, 1781– 1802, DOI: 10.1002/jcc.20289Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SlsbbJ&md5=189051128443b547f4300a1b8fb0e034Scalable molecular dynamics with NAMDPhillips, James C.; Braun, Rosemary; Wang, Wei; Gumbart, James; Tajkhorshid, Emad; Villa, Elizabeth; Chipot, Christophe; Skeel, Robert D.; Kale, Laxmikant; Schulten, KlausJournal of Computational Chemistry (2005), 26 (16), 1781-1802CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)NAMD is a parallel mol. dynamics code designed for high-performance simulation of large biomol. systems. NAMD scales to hundreds of processors on high-end parallel platforms, as well as tens of processors on low-cost commodity clusters, and also runs on individual desktop and laptop computers. NAMD works with AMBER and CHARMM potential functions, parameters, and file formats. This article, directed to novices as well as experts, first introduces concepts and methods used in the NAMD program, describing the classical mol. dynamics force field, equations of motion, and integration methods along with the efficient electrostatics evaluation algorithms employed and temp. and pressure controls used. Features for steering the simulation across barriers and for calcg. both alchem. and conformational free energy differences are presented. The motivations for and a roadmap to the internal design of NAMD, implemented in C++ and based on Charm++ parallel objects, are outlined. The factors affecting the serial and parallel performance of a simulation are discussed. Finally, typical NAMD use is illustrated with representative applications to a small, a medium, and a large biomol. system, highlighting particular features of NAMD, for example, the Tcl scripting language. The article also provides a list of the key features of NAMD and discusses the benefits of combining NAMD with the mol. graphics/sequence anal. software VMD and the grid computing/collab. software BioCoRE. NAMD is distributed free of charge with source code at www.ks.uiuc.edu.
- 20Coveney, P. V.; Wan, S. On the calculation of equilibrium thermodynamic properties from molecular dynamics. Phys. Chem. Chem. Phys. 2016, 18, 30236– 30240, DOI: 10.1039/C6CP02349EGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XntVyisr8%253D&md5=f7950af8255cf494a117d46bc1b5191eOn the calculation of equilibrium thermodynamic properties from molecular dynamicsCoveney, Peter V.; Wan, ShunzhouPhysical Chemistry Chemical Physics (2016), 18 (44), 30236-30240CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The purpose of statistical mechanics is to provide a route to the calcn. of macroscopic properties of matter from their constituent microscopic components. It is well known that the macrostates emerge as ensemble avs. of microstates. However, this is more often stated than implemented in computer simulation studies. Here we consider foundational aspects of statistical mechanics which are overlooked in most textbooks and research articles that purport to compute macroscopic behavior from microscopic descriptions based on classical mechanics and show how due attention to these issues leads in directions which have not been widely appreciated in the field of mol. dynamics simulation.
- 21Lim, N. M.; Wang, L.; Abel, R.; Mobley, D. L. Sensitivity in binding free energies due to protein reorganization. J. Chem. Theory Comput. 2016, 12, 4620– 4631, DOI: 10.1021/acs.jctc.6b00532Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xht1CltrbO&md5=5f4525fb7a55b6c9ad0510168e13f316Sensitivity in Binding Free Energies Due to Protein ReorganizationLim, Nathan M.; Wang, Lingle; Abel, Robert; Mobley, David L.Journal of Chemical Theory and Computation (2016), 12 (9), 4620-4631CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Tremendous recent improvements in computer hardware, coupled with advances in sampling techniques and force fields, are now allowing protein-ligand binding free energy calcns. to be routinely used to aid pharmaceutical drug discovery projects. However, despite these recent innovations, there are still needs for further improvement in sampling algorithms to more adequately sample protein motion relevant to protein-ligand binding. Here, we report our work identifying and studying such clear and remaining needs in the apolar cavity of T4 lysozyme L99A. In this study, we model recent exptl. results that show the progressive opening of the binding pocket in response to a series of homologous ligands. Even while using enhanced sampling techniques, we demonstrate that the predicted relative binding free energies (RBFE) are sensitive to the initial protein conformational state. Particularly, we highlight the importance of sufficient sampling of protein conformational changes and demonstrate how inclusion of three key protein residues in the "hot" region of the FEP/REST simulation improves the sampling and resolves this sensitivity, given enough simulation time.
- 22Clark, A. J.; Gindin, T.; Zhang, B.; Wang, L.; Abel, R.; Murret, C. S.; Xu, F.; Bao, A.; Lu, N. J.; Zhou, T.; Kwong, P. D.; Shapiro, L.; Honig, B.; Friesner, R. A. Free energy perturbation calculation of relative binding free energy between broadly neutralizing antibodies and the gp120 glycoprotein of HIV-1. J. Mol. Biol. 2017, 429, 930– 947, DOI: 10.1016/j.jmb.2016.11.021Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitVSmsL3I&md5=ff264d642726916a45bae5b121bbdd06Free energy perturbation calculation of relative binding free energy between broadly neutralizing antibodies and the gp120 glycoprotein of HIV-1Clark, Anthony J.; Gindin, Tatyana; Zhang, Baoshan; Wang, Lingle; Abel, Robert; Murret, Colleen S.; Xu, Fang; Bao, Amy; Lu, Nina J.; Zhou, Tongqing; Kwong, Peter D.; Shapiro, Lawrence; Honig, Barry; Friesner, Richard A.Journal of Molecular Biology (2017), 429 (7), 930-947CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)Direct calcn. of relative binding affinities between antibodies and antigens is a long-sought goal. However, despite substantial efforts, no generally applicable computational method has been described. Here, we describe a systematic free energy perturbation (FEP) protocol and calc. the binding affinities between the gp120 envelope glycoprotein of HIV-1 and three broadly neutralizing antibodies (bNAbs) of the VRC01 class. The protocol has been adapted from successful studies of small mols. to address the challenges assocd. with modeling protein-protein interactions. Specifically, we built homol. models of the three antibody-gp120 complexes, extended the sampling times for large bulky residues, incorporated the modeling of glycans on the surface of gp120, and utilized continuum solvent-based loop prediction protocols to improve sampling. We present three exptl. surface plasmon resonance data sets, in which antibody residues in the antibody/gp120 interface were systematically mutated to alanine. The RMS error in the large set (55 total cases) of FEP tests as compared to these expts., 0.68 kcal/mol, is near exptl. accuracy, and it compares favorably with the results obtained from a simpler, empirical methodol. The correlation coeff. for the combined data set including residues with glycan contacts, R2 = 0.49, should be sufficient to guide the choice of residues for antibody optimization projects, assuming that this level of accuracy can be realized in prospective prediction. More generally, these results are encouraging with regard to the possibility of using an FEP approach to calc. the magnitude of protein-protein binding affinities.
- 23Fratev, F.; Sirimulla, S. An improved free energy perturbation FEP+ sampling protocol for flexible ligand-binding domains. ChemRxiv 2018, DOI: 10.26434/chemrxiv.6204167.v1 .Google ScholarThere is no corresponding record for this reference.
- 24Fiser, A.; Sali, A. ModLoop: automated modeling of loops in protein structures. Bioinformatics 2003, 19, 2500– 2501, DOI: 10.1093/bioinformatics/btg362Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXpvVSisLk%253D&md5=4b0fdaca0a412715682496883ab2019cModLoop: automated modeling of loops in protein structuresFiser, Andras; Sali, AndrejBioinformatics (2003), 19 (18), 2500-2501CODEN: BOINFP; ISSN:1367-4803. (Oxford University Press)Summary: ModLoop is a web server for automated modeling of loops in protein structures. The input is the at. coordinates of the protein structure in the Protein Data Bank format, and the specification of the starting and ending residues of one or more segments to be modeled, contg. no more than 20 residues in total. The output is the coordinates of the non-hydrogen atoms in the modeled segments. A user provides the input to the server via a simple web interface, and receives the output by e-mail. The server relies on the loop modeling routine in MODELLER that predicts the loop conformations by satisfaction of spatial restraints, without relying on a database of known protein structures. For a rapid response, ModLoop runs on a cluster of Linux PC computers.
- 25Sohl, C. D.; Ryan, M. R.; Luo, B.; Frey, K. M.; Anderson, K. S. Illuminating the molecular mechanisms of tyrosine kinase inhibitor resistance for the FGFR1 gatekeeper mutation: the Achilles’ heel of targeted therapy. ACS Chem. Biol. 2015, 10, 1319– 1329, DOI: 10.1021/acschembio.5b00014Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXislGitLo%253D&md5=6c13f193e9e4c3f38b1e59f46fbc6342Illuminating the Molecular Mechanisms of Tyrosine Kinase Inhibitor Resistance for the FGFR1 Gatekeeper Mutation: The Achilles' Heel of Targeted TherapySohl, Christal D.; Ryan, Molly R.; Luo, BeiBei; Frey, Kathleen M.; Anderson, Karen S.ACS Chemical Biology (2015), 10 (5), 1319-1329CODEN: ACBCCT; ISSN:1554-8929. (American Chemical Society)Human fibroblast growth factor receptors (FGFRs) 1-4 are a family of receptor tyrosine kinases that can serve as drivers of tumorigenesis. In particular, FGFR1 gene amplification has been implicated in squamous cell lung and breast cancers. Tyrosine kinase inhibitors (TKIs) targeting FGFR1, including AZD4547 and E3810 (Lucitanib), are currently in early phase clin. trials. Unfortunately, drug resistance limits the long-term success of TKIs, with mutations at the "gatekeeper" residue leading to tumor progression. Here we show the first structural and kinetic characterization of the FGFR1 gatekeeper mutation, V561M FGFR1. The V561M mutation confers a 38-fold increase in autophosphorylation achieved at least in part by a network of interacting residues forming a hydrophobic spine to stabilize the active conformation. Moreover, kinetic assays established that the V561M mutation confers significant resistance to E3810, while retaining affinity for AZD4547. Structural analyses of these TKIs with wild type (WT) and gatekeeper mutant forms of FGFR1 offer clues to developing inhibitors that maintain potency against gatekeeper mutations. We show that AZD4547 affinity is preserved by V561M FGFR1 due to a flexible linker that allows multiple inhibitor binding modes. This is the first example of a TKI binding in distinct conformations to WT and gatekeeper mutant forms of FGFR, highlighting adaptable regions in both the inhibitor and binding pocket crucial for drug design. Exploiting inhibitor flexibility to overcome drug resistance has been a successful strategy for combating diseases such as AIDS and may be an important approach for designing inhibitors effective against kinase gatekeeper mutations.
- 26Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03; Gaussian, Inc.: Wallingford, CT, 2004.Google ScholarThere is no corresponding record for this reference.
- 27Case, D. A.; Cheatham, T. E.; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. J. The Amber biomolecular simulation programs. J. Comput. Chem. 2005, 26, 1668– 1688, DOI: 10.1002/jcc.20290Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SlsbbM&md5=93be29ff894bab96c783d24e9886c7d0The amber biomolecular simulation programsCase, David A.; Cheatham, Thomas E., III; Darden, Tom; Gohlke, Holger; Luo, Ray; Merz, Kenneth M., Jr.; Onufriev, Alexey; Simmerling, Carlos; Wang, Bing; Woods, Robert J.Journal of Computational Chemistry (2005), 26 (16), 1668-1688CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The authors describe the development, current features, and some directions for future development of the Amber package of computer programs. This package evolved from a program that was constructed in the late 1970s to do Assisted Model Building with Energy Refinement, and now contains a group of programs embodying a no. of powerful tools of modern computational chem., focused on mol. dynamics and free energy calcns. of proteins, nucleic acids, and carbohydrates.
- 28Wang, 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, 1157– 1174, DOI: 10.1002/jcc.20035Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksFakurc%253D&md5=2992017a8cf51f89290ae2562403b115Development and testing of a general Amber force fieldWang, Junmei; Wolf, Romain M.; Caldwell, James W.; Kollman, Peter A.; Case, David A.Journal of Computational Chemistry (2004), 25 (9), 1157-1174CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)We describe here a general Amber force field (GAFF) for org. mols. GAFF is designed to be compatible with existing Amber force fields for proteins and nucleic acids, and has parameters for most org. and pharmaceutical mols. that are composed of H, C, N, O, S, P, and halogens. It uses a simple functional form and a limited no. of atom types, but incorporates both empirical and heuristic models to est. force consts. and partial at. charges. The performance of GAFF in test cases is encouraging. In test I, 74 crystallog. structures were compared to GAFF minimized structures, with a root-mean-square displacement of 0.26 Å, which is comparable to that of the Tripos 5.2 force field (0.25 Å) and better than those of MMFF 94 and CHARMm (0.47 and 0.44 Å, resp.). In test II, gas phase minimizations were performed on 22 nucleic acid base pairs, and the minimized structures and intermol. energies were compared to MP2/6-31G* results. The RMS of displacements and relative energies were 0.25 Å and 1.2 kcal/mol, resp. These data are comparable to results from Parm99/RESP (0.16 Å and 1.18 kcal/mol, resp.), which were parameterized to these base pairs. Test III looked at the relative energies of 71 conformational pairs that were used in development of the Parm99 force field. The RMS error in relative energies (compared to expt.) is about 0.5 kcal/mol. GAFF can be applied to wide range of mols. in an automatic fashion, making it suitable for rational drug design and database searching.
- 29Maier, 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, 3696– 3713, DOI: 10.1021/acs.jctc.5b00255Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFequ7rN&md5=7b803577b3b6912cc6750cfbd356596eff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SBMaier, James A.; Martinez, Carmenza; Kasavajhala, Koushik; Wickstrom, Lauren; Hauser, Kevin E.; Simmerling, CarlosJournal of Chemical Theory and Computation (2015), 11 (8), 3696-3713CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Mol. mechanics is powerful for its speed in atomistic simulations, but an accurate force field is required. The Amber ff99SB force field improved protein secondary structure balance and dynamics from earlier force fields like ff99, but weaknesses in side chain rotamer and backbone secondary structure preferences have been identified. Here, we performed a complete refit of all amino acid side chain dihedral parameters, which had been carried over from ff94. The training set of conformations included multidimensional dihedral scans designed to improve transferability of the parameters. Improvement in all amino acids was obtained as compared to ff99SB. Parameters were also generated for alternate protonation states of ionizable side chains. Av. errors in relative energies of pairs of conformations were under 1.0 kcal/mol as compared to QM, reduced 35% from ff99SB. We also took the opportunity to make empirical adjustments to the protein backbone dihedral parameters as compared to ff99SB. Multiple small adjustments of φ and ψ parameters were tested against NMR scalar coupling data and secondary structure content for short peptides. The best results were obtained from a phys. motivated adjustment to the φ rotational profile that compensates for lack of ff99SB QM training data in the β-ppII transition region. Together, these backbone and side chain modifications (hereafter called ff14SB) not only better reproduced their benchmarks, but also improved secondary structure content in small peptides and reprodn. of NMR χ1 scalar coupling measurements for proteins in soln. We also discuss the Amber ff12SB parameter set, a preliminary version of ff14SB that includes most of its improvements.
- 30Jo, S.; Jiang, W. A generic implementation of replica exchange with solute tempering (REST2) algorithm in NAMD for complex biophysical simulations. Comput. Phys. Commun. 2015, 197, 304– 311, DOI: 10.1016/j.cpc.2015.08.030Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsFSlu73I&md5=776ea438c536bf84540b1443494e647cA generic implementation of replica exchange with solute tempering (REST2) algorithm in NAMD for complex biophysical simulationsJo, Sunhwan; Jiang, WeiComputer Physics Communications (2015), 197 (), 304-311CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier B.V.)Replica Exchange with Solute Tempering (REST2) is a powerful sampling enhancement algorithm of mol. dynamics (MD) in that it needs significantly smaller no. of replicas but achieves higher sampling efficiency relative to std. temp. exchange algorithm. In this paper, we extend the applicability of REST2 for quant. biophys. simulations through a robust and generic implementation in greatly scalable MD software NAMD. The rescaling procedure of force field parameters controlling REST2 "hot region" is implemented into NAMD at the source code level. A user can conveniently select hot region through VMD and write the selection information into a PDB file. The rescaling keyword/parameter is written in NAMD Tcl script interface that enables an on-the-fly simulation parameter change. Our implementation of REST2 is within communication-enabled Tcl script built on top of Charm++, thus communication overhead of an exchange attempt is vanishingly small. Such a generic implementation facilitates seamless cooperation between REST2 and other modules of NAMD to provide enhanced sampling for complex biomol. simulations. Three challenging applications including native REST2 simulation for peptide folding-unfolding transition, free energy perturbation/REST2 for abs. binding affinity of protein-ligand complex and umbrella sampling/REST2 Hamiltonian exchange for free energy landscape calcn. were carried out on IBM Blue Gene/Q supercomputer to demonstrate efficacy of REST2 based on the present implementation.
- 31Ballester, P. J.; Mitchell, J. B. O. A machine learning approach to predicting protein–ligand binding affinity with applications to molecular docking. Bioinformatics 2010, 26, 1169– 1175, DOI: 10.1093/bioinformatics/btq112Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXlt1Cjs78%253D&md5=6d2d48344dc5ba70f0ff85f432d0ec63A machine learning approach to predicting protein-ligand binding affinity with applications to molecular dockingBallester, Pedro J.; Mitchell, John B. O.Bioinformatics (2010), 26 (9), 1169-1175CODEN: BOINFP; ISSN:1367-4803. (Oxford University Press)Accurately predicting the binding affinities of large sets of diverse protein-ligand complexes is an extremely challenging task. The scoring functions that attempt such computational prediction are essential for analyzing the outputs of mol. docking, which in turn is an important technique for drug discovery, chem. biol. and structural biol. Each scoring function assumes a predetd. theory-inspired functional form for the relationship between the variables that characterize the complex, which also include parameters fitted to exptl. or simulation data and its predicted binding affinity. The inherent problem of this rigid approach is that it leads to poor predictivity for those complexes that do not conform to the modeling assumptions. Moreover, resampling strategies, such as cross-validation or bootstrapping, are still not systematically used to guard against the overfitting of calibration data in parameter estn. for scoring functions. We propose a novel scoring function (RF-Score) that circumvents the need for problematic modeling assumptions via non-parametric machine learning. In particular, Random Forest was used to implicitly capture binding effects that are hard to model explicitly. RF-Score is compared with the state of the art on the demanding PDBbind benchmark. Results show that RF-Score is a very competitive scoring function. Importantly, RF-Score's performance was shown to improve dramatically with training set size and hence the future availability of more high-quality structural and interaction data is expected to lead to improved versions of RF-Score.
- 32Jimenez, J.; Skalic, M.; Martinez-Rosell, G.; De Fabritiis, G. KDEEP: Protein–Ligand Absolute Binding Affinity Prediction via 3D-Convolutional Neural Networks. J. Chem. Inf. Model. 2018, 58, 287– 296, DOI: 10.1021/acs.jcim.7b00650Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXkslSitQ%253D%253D&md5=81943a6732be99e5439e1e3a25fa4414KDEEP: Protein-Ligand Absolute Binding Affinity Prediction via 3D-Convolutional Neural NetworksJimenez, Jose; Skalic, Miha; Martinez-Rosell, Gerard; De Fabritiis, GianniJournal of Chemical Information and Modeling (2018), 58 (2), 287-296CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Accurately predicting protein-ligand binding affinities is an important problem in computational chem. since it can substantially accelerate drug discovery for virtual screening and lead optimization. We propose here a fast machine-learning approach for predicting binding affinities using state-of-the-art 3D-convolutional neural networks and compare this approach to other machine-learning and scoring methods using several diverse data sets. The results for the std. PDBbind (v.2016) core test-set are state-of-the-art with a Pearson's correlation coeff. of 0.82 and a RMSE of 1.27 in pK units between exptl. and predicted affinity, but accuracy is still very sensitive to the specific protein used. KDEEP is made available via PlayMol.org for users to test easily their own protein-ligand complexes, with each prediction taking a fraction of a second. We believe that the speed, performance, and ease of use of KDEEP makes it already an attractive scoring function for modern computational chem. pipelines.
- 33Coveney, P. V.; Dougherty, E. R.; Highfield, R. R. Big data need big theory too. Philos. Trans. R. Soc., A 2016, 374, 20160153, DOI: 10.1098/rsta.2016.0153Google ScholarThere is no corresponding record for this reference.
Cited By
This article is cited by 27 publications.
- Shunzhou Wan, Agastya P. Bhati, Peter V. Coveney. Comparison of Equilibrium and Nonequilibrium Approaches for Relative Binding Free Energy Predictions. Journal of Chemical Theory and Computation 2023, 19
(21)
, 7846-7860. https://doi.org/10.1021/acs.jctc.3c00842
- Christopher Lockhart, Xingyu Luo, Audrey Olson, Bryan M. Delfing, Xavier E. Laracuente, Kenneth W. Foreman, Mikell Paige, Kylene Kehn-Hall, Dmitri K. Klimov. Can Free Energy Perturbation Simulations Coupled with Replica-Exchange Molecular Dynamics Study Ligands with Distributed Binding Sites?. Journal of Chemical Information and Modeling 2023, 63
(15)
, 4791-4802. https://doi.org/10.1021/acs.jcim.3c00631
- Wei Jiang. Enhanced Configurational Sampling Approaches to Alchemical Ligand Binding Free Energy Simulations: Current Status and Challenges. The Journal of Physical Chemistry B 2023, 127
(31)
, 6835-6841. https://doi.org/10.1021/acs.jpcb.3c02020
- Agastya P. Bhati, Art Hoti, Andrew Potterton, Mateusz K. Bieniek, Peter V. Coveney. Long Time Scale Ensemble Methods in Molecular Dynamics: Ligand–Protein Interactions and Allostery in SARS-CoV-2 Targets. Journal of Chemical Theory and Computation 2023, 19
(11)
, 3359-3378. https://doi.org/10.1021/acs.jctc.3c00020
- Tai-Sung Lee, Hsu-Chun Tsai, Abir Ganguly, Darrin M. York. ACES: Optimized Alchemically Enhanced Sampling. Journal of Chemical Theory and Computation 2023, 19
(2)
, 472-487. https://doi.org/10.1021/acs.jctc.2c00697
- 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
- Agastya P. Bhati, Peter V. Coveney. Large Scale Study of Ligand–Protein Relative Binding Free Energy Calculations: Actionable Predictions from Statistically Robust Protocols. Journal of Chemical Theory and Computation 2022, 18
(4)
, 2687-2702. https://doi.org/10.1021/acs.jctc.1c01288
- Mateusz K. Bieniek, Agastya P. Bhati, Shunzhou Wan, Peter V. Coveney. TIES 20: Relative Binding Free Energy with a Flexible Superimposition Algorithm and Partial Ring Morphing. Journal of Chemical Theory and Computation 2021, 17
(2)
, 1250-1265. https://doi.org/10.1021/acs.jctc.0c01179
- Yaozong Li, Kwangho Nam. Repulsive Soft-Core Potentials for Efficient Alchemical Free Energy Calculations. Journal of Chemical Theory and Computation 2020, 16
(8)
, 4776-4789. https://doi.org/10.1021/acs.jctc.0c00163
- Rhys Evans, Ladislav Hovan, Gareth A. Tribello, Benjamin P. Cossins, Carolina Estarellas, Francesco L. Gervasio. Combining Machine Learning and Enhanced Sampling Techniques for Efficient and Accurate Calculation of Absolute Binding Free Energies. Journal of Chemical Theory and Computation 2020, 16
(7)
, 4641-4654. https://doi.org/10.1021/acs.jctc.0c00075
- Kyriakos Georgiou, Athina Konstantinidi, Johanna Hutterer, Kathrin Freudenberger, Felix Kolarov, George Lambrinidis, Ioannis Stylianakis, Margarita Stampelou, Günter Gauglitz, Antonios Kolocouris. Accurate calculation of affinity changes to the close state of influenza A M2 transmembrane domain in response to subtle structural changes of adamantyl amines using free energy perturbation methods in different lipid bilayers. Biochimica et Biophysica Acta (BBA) - Biomembranes 2024, 1866
(2)
, 184258. https://doi.org/10.1016/j.bbamem.2023.184258
- Martiniano Bello, Cindy Bandala. Evaluating the ability of end-point methods to predict the binding affinity tendency of protein kinase inhibitors. RSC Advances 2023, 13
(36)
, 25118-25128. https://doi.org/10.1039/D3RA04916G
- Yang Yu, Zhe Wang, Lingling Wang, Sheng Tian, Tingjun Hou, Huiyong Sun. Predicting the mutation effects of protein–ligand interactions via end-point binding free energy calculations: strategies and analyses. Journal of Cheminformatics 2022, 14
(1)
https://doi.org/10.1186/s13321-022-00639-y
- Alice E. Brankin, Philip W. Fowler. Predicting antibiotic resistance in complex protein targets using alchemical free energy methods. Journal of Computational Chemistry 2022, 43
(26)
, 1771-1782. https://doi.org/10.1002/jcc.26979
- Emilia P. Barros, Benjamin Ries, Lennard Böselt, Candide Champion, Sereina Riniker. Recent developments in multiscale free energy simulations. Current Opinion in Structural Biology 2022, 72 , 55-62. https://doi.org/10.1016/j.sbi.2021.08.003
- Ran Friedman. Computational studies of protein–drug binding affinity changes upon mutations in the drug target. WIREs Computational Molecular Science 2022, 12
(1)
https://doi.org/10.1002/wcms.1563
- Agastya P. Bhati, Shunzhou Wan, Dario Alfè, Austin R. Clyde, Mathis Bode, Li Tan, Mikhail Titov, Andre Merzky, Matteo Turilli, Shantenu Jha, Roger R. Highfield, Walter Rocchia, Nicola Scafuri, Sauro Succi, Dieter Kranzlmüller, Gerald Mathias, David Wifling, Yann Donon, Alberto Di Meglio, Sofia Vallecorsa, Heng Ma, Anda Trifan, Arvind Ramanathan, Tom Brettin, Alexander Partin, Fangfang Xia, Xiaotan Duan, Rick Stevens, Peter V. Coveney. Pandemic drugs at pandemic speed: infrastructure for accelerating COVID-19 drug discovery with hybrid machine learning- and physics-based simulations on high-performance computers. Interface Focus 2021, 11
(6)
https://doi.org/10.1098/rsfs.2021.0018
- Shunzhou Wan, Deepak Kumar, Valentin Ilyin, Ussama Al Homsi, Gulab Sher, Alexander Knuth, Peter V. Coveney. The effect of protein mutations on drug binding suggests ensuing personalised drug selection. Scientific Reports 2021, 11
(1)
https://doi.org/10.1038/s41598-021-92785-w
- Aparna Vilas Dongre, Sudip Das, Asutosh Bellur, Sanjeev Kumar, Anusha Chandrashekarmath, Tarak Karmakar, Padmanabhan Balaram, Sundaram Balasubramanian, Hemalatha Balaram. Structural basis for the hyperthermostability of an archaeal enzyme induced by succinimide formation. Biophysical Journal 2021, 120
(17)
, 3732-3746. https://doi.org/10.1016/j.bpj.2021.07.014
- Edward King, Erick Aitchison, Han Li, Ray Luo. Recent Developments in Free Energy Calculations for Drug Discovery. Frontiers in Molecular Biosciences 2021, 8 https://doi.org/10.3389/fmolb.2021.712085
- Himanshu Goel, Anthony Hazel, Vincent D. Ustach, Sunhwan Jo, Wenbo Yu, Alexander D. MacKerell. Rapid and accurate estimation of protein–ligand relative binding affinities using site-identification by ligand competitive saturation. Chemical Science 2021, 12
(25)
, 8844-8858. https://doi.org/10.1039/D1SC01781K
- Shunzhou Wan, Robert C. Sinclair, Peter V. Coveney. Uncertainty quantification in classical molecular dynamics. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 2021, 379
(2197)
https://doi.org/10.1098/rsta.2020.0082
- Natacha Gillet, Alessio Bartocci, Elise Dumont. Assessing the sequence dependence of pyrimidine–pyrimidone (6–4) photoproduct in a duplex double-stranded DNA: A pitfall for microsecond range simulation. The Journal of Chemical Physics 2021, 154
(13)
https://doi.org/10.1063/5.0041332
- Shunzhou Wan, Andrew Potterton, Fouad S. Husseini, David W. Wright, Alexander Heifetz, Maciej Malawski, Andrea Townsend-Nicholson, Peter V. Coveney. Hit-to-lead and lead optimization binding free energy calculations for G protein-coupled receptors. Interface Focus 2020, 10
(6)
, 20190128. https://doi.org/10.1098/rsfs.2019.0128
- Shunzhou Wan, Agastya P. Bhati, Stefan J. Zasada, Peter V. Coveney. Rapid, accurate, precise and reproducible ligand–protein binding free energy prediction. Interface Focus 2020, 10
(6)
, 20200007. https://doi.org/10.1098/rsfs.2020.0007
- Robert Hall, Tom Dixon, Alex Dickson. On Calculating Free Energy Differences Using Ensembles of Transition Paths. Frontiers in Molecular Biosciences 2020, 7 https://doi.org/10.3389/fmolb.2020.00106
- Shunzhou Wan, Gary Tresadern, Laura Pérez‐Benito, Herman van Vlijmen, Peter V. Coveney. Accuracy and Precision of Alchemical Relative Free Energy Predictions with and without Replica‐Exchange. Advanced Theory and Simulations 2020, 3
(1)
https://doi.org/10.1002/adts.201900195
Abstract
Figure 1
Figure 1. Structures of FGFR3 and inhibitors studied in this work: (a) the binding site of tyrosine kinase domain for FGFR3 in complex with ACP, an ATP-analogue (PDB ID: 4K33). FGFR3 is depicted in cartoon and ACP in bond representation. Mutations of three residues, V555, I538, and N540 (ball-and-stick representation), are among the most common genetic variants in FGFR3. The chemical structures of four ATP competitive inhibitors are shown in panels b–e: (b) AZD4547, (c) BGJ-398, (d) TKI258, and (e) JNJ4275649.
Figure 2
Figure 2. Different regions in the λ-REST2 simulations. The AZD4547-V555M complex is shown here as an example. The hybrid residue, denoted as the alchemical region, is depicted as a ball-and-stick model. It consists of disappearing (red) and appearing (blue) groups which are slightly separated for reasons of clarity. They can fully or partially overlap in the simulation as there are no interactions between them. The REST2 region, including the alchemical region (red and blue ball-and-stick), part of the ligand (orange bond), and surrounding protein residues (orange stick), is designated as the “hot” region. The selection of the REST2 region is described in the main text (section 2.3 REST2 region).
Figure 3
Figure 3. Comparison of the predicted ΔΔGcalc values using TIES (black circles), TIES-λ-REST2 (λR2, up/down triangles) with those from experiments for V555M mutant complexes with the highest Teff of the chosen REST2 region at 600 K for receptor and complexes except those with AZD4547 which are at 1500 K (red triangles pointing up), and at 1500 K for receptor and 3000 K for complexes (blue triangles pointing down). Results of AZD4547 from the bent conformation are represented using filled circles and triangles. The dotted lines (x = 0 and y = 0) create four quadrants. Data points in quadrants I (x > 0 and y > 0) and III (x < 0 and y < 0) indicate that the calculated binding free energy differences agree directionally with those from the experimental data. The results from TIES-λ-REST2-M (λR2-M) are very close to those from λR2 (Table 1), and are not shown for reasons of clarity.
Figure 4
Figure 4. Comparison of the predicted ΔΔGcalc values using TIES (black circles), TIES-λ-REST2 (λR2, up/down triangles) with those from experiment for for all inhibitors bound to FGFR3: (a) I538V mutant and (b) N540S mutant, when the highest Teff for the chosen REST2 region is at 800 K for receptor and 1500 K for complexes (red triangles pointing up) and at 1500 K for receptor and 3000 K for complexes (blue triangles pointing down). The dotted lines (x = 0 and y = 0) create four quadrants. Data points in quadrants I (x > 0 and y > 0) and III (x < 0 and y < 0) indicate that the ΔΔGcalc values agree directionally with ΔΔGexp. The results from TIES-λ-REST2-M (λR2-M) are very close to those from λR2 (Table 1), and are not shown for reasons of clarity.
Figure 5
Figure 5. Variation of the cumulative average of ΔΔGcalcTIES with simulation length for all inhibitors bound to the FGFR3 I538V mutant. The corresponding experimental value for each inhibitor is shown by a dashed line of the same color.
Figure 6
Figure 6. Variation of the cumulative average of ΔΔG calculated using schemes λR2 and λR2-M with simulation length for all complexes. The highest Teff for receptor and complex are 1500 and 3000 K, respectively, in the case of I538V and N540S mutants, while the corresponding values in the case of V555M mutant are 600 and 600 K/1500 K. The corresponding experimental values for each inhibitor are shown by a dashed line of the same color.
Figure 7
Figure 7. Two distinct conformations of inhibitor AZD4547 found experimentally when bound to the FGFR gatekeeper mutant. The three hydrogen bonds, marked with black dashed lines and labeled as H1, H2, and H3, keep the middle portion of the inhibitor stable. The value of the dihedral angle between the four carbon atoms highlighted in orange can be used as an indicator of the occurrence of the two conformations. The atoms displayed as balls lie in the REST2 region while the ones displayed as lines reside outside it.
Figure 8
Figure 9
Figure 9. Normalized frequency distributions of the dihedral angle between the four carbon atoms highlighted in orange in Figure 7 for different λ states of V555M–AZD4547 complexes in standard TIES (in blue) as well as λ-REST2 simulations showing the relative populations of the two conformations of AZD4547. In the case of λ-REST2 simulations, the distributions from the first (1–4 ns; in black) and the last 4 ns (17–20 ns; in red) are shown separately.
Figure 10
Figure 10. Inhibitor TKI258 bound to V555M mutant. It forms two hydrogen bonds with the hinge region of the protein which are displayed with black dashed lines and labeled as H1 and H2. The atoms shown as balls lie in the “hot” region. The atoms are shown in the standard color code: carbon in green, oxygen in red, nitrogen in blue, hydrogen in white, and fluorine in pink.
Figure 11
Figure 11. Normalized frequency distributions of H1 and H2 from Figure 10 for the two end-points in the case of the standard TIES as well as λ-REST2 simulations of the V555M–TKI258 complex. λ-REST2 simulations sample a larger comformational space than TIES, as evidenced by the lower and wider distributions of the distances, and the second peaks in the λ = 1 end-point (V555M).
Figure 12
Figure 12. Normalized frequency distributions of H1 and H2 from Figure 10 for the two end-points in the case of standard TIES as well as λ-REST2 simulations of the I538V–TKI258 complex. The long tails and additional peaks beyond 4 Å indicate that λ-REST2 simulations sample some conformations irrelevant to stable inhibitor binding.
Figure 13
Figure 13. Normalized frequency distributions of H1 and H2 from Figure 10 for the two end-points in the case of standard TIES as well as λ-REST2 simulations of the N540S–TKI258 complex. The long tails and additional peaks beyond 4 Å indicate that λ-REST2 simulations sample some conformations irrelevant to stable inhibitor binding.
References
ARTICLE SECTIONSThis article references 33 other publications.
- 1Wright, D. W.; Wan, S.; Shublaq, N.; Zasada, S. J.; Coveney, P. V. From base pair to bedside: molecular simulation and the translation of genomics to personalized medicine. WIREs Syst. Biol. Med. 2012, 4, 585– 598, DOI: 10.1002/wsbm.1186Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXpslSgsg%253D%253D&md5=b6bf5cb6b9fab33543986b9f8ceed6ecFrom base pair to bedside: molecular simulation and the translation of genomics to personalized medicineWright, David W.; Wan, Shunzhou; Shublaq, Nour; Zasada, Stefan J.; Coveney, Peter V.Wiley Interdisciplinary Reviews: Systems Biology and Medicine (2012), 4 (6), 585-598CODEN: WIRSBW; ISSN:1939-005X. (Wiley-Blackwell)A review. Despite the promises made that genomic sequencing would transform therapy by introducing a new era of personalized medicine, relatively few tangible breakthroughs have been made. This has led to the recognition that complex interactions at multiple spatial, temporal, and organizational levels may often combine to produce disease. Understanding this complexity requires that existing and future models are used and interpreted within a framework that incorporates knowledge derived from investigations at multiple levels of biol. function. It also requires a computational infrastructure capable of dealing with the vast quantities of data generated by genomic approaches. In this review, we discuss the use of mol. modeling to generate quant. and qual. insights at the smallest scales of the systems biol. hierarchy, how it can play an important role in the development of a systems understanding of disease and in the application of such knowledge to help discover new therapies and target existing ones on a personal level.
- 2Wang, L. 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, 2695– 2703, DOI: 10.1021/ja512751qGoogle Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsF2iuro%253D&md5=37a4f4a6c085f47ed531342643b6c33bAccurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force FieldWang, Lingle; Wu, Yujie; Deng, Yuqing; Kim, Byungchan; Pierce, Levi; Krilov, Goran; Lupyan, Dmitry; Robinson, Shaughnessy; Dahlgren, Markus K.; Greenwood, Jeremy; Romero, Donna L.; Masse, Craig; Knight, Jennifer L.; Steinbrecher, Thomas; Beuming, Thijs; Damm, Wolfgang; Harder, Ed; Sherman, Woody; Brewer, Mark; Wester, Ron; Murcko, Mark; Frye, Leah; Farid, Ramy; Lin, Teng; Mobley, David L.; Jorgensen, William L.; Berne, Bruce J.; Friesner, Richard A.; Abel, RobertJournal of the American Chemical Society (2015), 137 (7), 2695-2703CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Designing tight-binding ligands is a primary objective of small-mol. drug discovery. Over the past few decades, free-energy calcns. have benefited from improved force fields and sampling algorithms, as well as the advent of low-cost parallel computing. However, it has proven to be challenging to reliably achieve the level of accuracy that would be needed to guide lead optimization (∼5× in binding affinity) for a wide range of ligands and protein targets. Not surprisingly, widespread com. application of free-energy simulations has been limited due to the lack of large-scale validation coupled with the tech. challenges traditionally assocd. with running these types of calcns. Here, we report an approach that achieves an unprecedented level of accuracy across a broad range of target classes and ligands, with retrospective results encompassing 200 ligands and a wide variety of chem. perturbations, many of which involve significant changes in ligand chem. structures. In addn., we have applied the method in prospective drug discovery projects and found a significant improvement in the quality of the compds. synthesized that have been predicted to be potent. Compds. predicted to be potent by this approach have a substantial redn. in false positives relative to compds. synthesized on the basis of other computational or medicinal chem. approaches. Furthermore, the results are consistent with those obtained from our retrospective studies, demonstrating the robustness and broad range of applicability of this approach, which can be used to drive decisions in lead optimization.
- 3Bernardi, R. C.; Melo, M. C.; Schulten, K. Enhanced sampling techniques in molecular dynamics simulations of biological systems. Biochim. Biophys. Acta, Gen. Subj. 2015, 1850, 872– 877, DOI: 10.1016/j.bbagen.2014.10.019Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVanurbN&md5=44e3ac1aa042c35ee08aa86a6a63e78eEnhanced sampling techniques in molecular dynamics simulations of biological systemsBernardi, Rafael C.; Melo, Marcelo C. R.; Schulten, KlausBiochimica et Biophysica Acta, General Subjects (2015), 1850 (5), 872-877CODEN: BBGSB3; ISSN:0304-4165. (Elsevier B.V.)A review. Mol. dynamics has emerged as an important research methodol. covering systems to the level of millions of atoms. However, insufficient sampling often limits its application. The limitation is due to rough energy landscapes, with many local min. sepd. by high-energy barriers, which govern the biomol. motion. In the past few decades methods have been developed that address the sampling problem, such as replica-exchange mol. dynamics, metadynamics and simulated annealing. Here the authors present an overview over theses sampling methods in an attempt to shed light on which should be selected depending on the type of system property studied. Enhanced sampling methods have been employed for a broad range of biol. systems and the choice of a suitable method is connected to biol. and phys. characteristics of the system, in particular system size. While metadynamics and replica-exchange mol. dynamics are the most adopted sampling methods to study biomol. dynamics, simulated annealing is well suited to characterize very flexible systems. The use of annealing methods for a long time was restricted to simulation of small proteins; however, a variant of the method, generalized simulated annealing, can be employed at a relatively low computational cost to large macromol. complexes. Mol. dynamics trajectories frequently do not reach all relevant conformational substates, for example those connected with biol. function, a problem that can be addressed by employing enhanced sampling algorithms. This article is part of a Special Issue entitled Recent developments of mol. dynamics.
- 4Valsson, O.; Parrinello, M. Variational approach to enhanced sampling and free energy calculations. Phys. Rev. Lett. 2014, 113, 090601, DOI: 10.1103/PhysRevLett.113.090601Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhslShtL7O&md5=4d66ba6f0d445e693e28dbeae09cd936Variational approach to enhanced sampling and free energy calculationsValsson, Omar; Parrinello, MichelePhysical Review Letters (2014), 113 (9), 090601CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)The ability of widely used sampling methods, such as mol. dynamics or Monte Carlo simulations, to explore complex free energy landscapes is severely hampered by the presence of kinetic bottlenecks. A large no. of solns. have been proposed to alleviate this problem. Many are based on the introduction of a bias potential which is a function of a small no. of collective variables. However constructing such a bias is not simple. Here we introduce a functional of the bias potential and an assocd. variational principle. The bias that minimizes the functional relates in a simple way to the free energy surface. This variational principle can be turned into a practical, efficient, and flexible sampling method. A no. of numerical examples are presented which include the detn. of a three-dimensional free energy surface. We argue that, beside being numerically advantageous, our variational approach provides a convenient and novel standpoint for looking at the sampling problem.
- 5Fukunishi, H.; Watanabe, O.; Takada, S. On the Hamiltonian replica exchange method for efficient sampling of biomolecular systems: Application to protein structure prediction. J. Chem. Phys. 2002, 116, 9058– 9067, DOI: 10.1063/1.1472510Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XjsFKmsLo%253D&md5=7ac571a5afdd63b0b4b29cfdec06f53bOn the Hamiltonian replica exchange method for efficient sampling of biomolecular systems: Application to protein structure predictionFukunishi, Hiroaki; Watanabe, Osamu; Takada, ShojiJournal of Chemical Physics (2002), 116 (20), 9058-9067CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Motivated by the protein structure prediction problem, we develop two variants of the Hamiltonian replica exchange methods (REMs) for efficient configuration sampling, (1) the scaled hydrophobicity REM and (2) the phantom chain REM, and compare their performance with the ordinary REM. We first point out that the ordinary REM has a shortage for the application to large systems such as biomols. and that the Hamiltonian REM, an alternative formulation of the REM, can give a remedy for it. We then propose two examples of the Hamiltonian REM that are suitable for a coarse-grained protein model. (1) The scaled hydrophobicity REM preps. replicas that are characterized by various strengths of hydrophobic interaction. The strongest interaction that mimics aq. soln. environment makes proteins folding, while weakened hydrophobicity unfolds proteins as in org. solvent. Exchange between these environments enables proteins to escape from misfolded traps and accelerate conformational search. This resembles the roles of mol. chaperone that assist proteins to fold in vivo. (2) The phantom chain REM uses replicas that allow various degrees of at. overlaps. By allowing at. overlap in some of replicas, the peptide chain can cross over itself, which can accelerate conformation sampling. Using a coarse-gained model we developed, we compute equil. probability distributions for poly-alanine 16-mer and for a small protein by these REMs and compare the accuracy of the results. We see that the scaled hydrophobicity REM is the most efficient method among the three REMs studied.
- 6Wang, L.; Friesner, R. A.; Berne, B. J. Replica exchange with solute scaling: A more efficient version of replica exchange with solute tempering (REST2). J. Phys. Chem. B 2011, 115, 9431– 9438, DOI: 10.1021/jp204407dGoogle Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXosFymurs%253D&md5=8acb0d670bcb70eacfc30e32d5dfb6ddReplica Exchange with Solute Scaling: A More Efficient Version of Replica Exchange with Solute Tempering (REST2)Wang, Lingle; Friesner, Richard A.; Berne, B. J.Journal of Physical Chemistry B (2011), 115 (30), 9431-9438CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)A small change in the Hamiltonian scaling in Replica Exchange with Solute Tempering (REST) is found to improve its sampling efficiency greatly, esp. for the sampling of aq. protein solns. in which there are large-scale solute conformation changes. Like the original REST (REST1), the new version (which the authors call REST2) also bypasses the poor scaling with system size of the std. Temp. Replica Exchange Method (TREM), reducing the no. of replicas (parallel processes) from what must be used in TREM. This redn. is accomplished by deforming the Hamiltonian function for each replica in such a way that the acceptance probability for the exchange of replica configurations does not depend on the no. of explicit water mols. in the system. For proof of concept, REST2 is compared with TREM and with REST1 for the folding of the trpcage and β-hairpin in water. The comparisons confirm that REST2 greatly reduces the no. of CPUs required by regular replica exchange and greatly increases the sampling efficiency over REST1. This method reduces the CPU time required for calcg. thermodn. avs. and for the ab initio folding of proteins in explicit water.
- 7Wang, L.; Berne, B. J.; Friesner, R. A. On achieving high accuracy and reliability in the calculation of relative protein–ligand binding affinities. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 1937– 1942, DOI: 10.1073/pnas.1114017109Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XivVSht74%253D&md5=23db7d18e624d5e1214bbd6702619a87On achieving high accuracy and reliability in the calculation of relative protein-ligand binding affinitiesWang, Lingle; Berne, B. J.; Friesner, Richard A.Proceedings of the National Academy of Sciences of the United States of America (2012), 109 (6), 1937-1942, S1937/1-S1937/7CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)We apply a free energy perturbation simulation method, free energy perturbation/replica exchange with solute tempering, to two modifications of protein-ligand complexes that lead to significant conformational changes, the first in the protein and the second in the ligand. The approach is shown to facilitate sampling in these challenging cases where high free energy barriers sep. the initial and final conformations and leads to superior convergence of the free energy as demonstrated both by consistency of the results (independence from the starting conformation) and agreement with exptl. binding affinity data. The second case, consisting of two neutral thrombin ligands that are taken from a recent medicinal chem. program for this interesting pharmaceutical target, is of particular significance in that it demonstrates that good results can be obtained for large, complex ligands, as opposed to relatively simple model systems. To achieve quant. agreement with expt. in the thrombin case, a next generation force field, Optimized Potentials for Liq. Simulations 2.0, is required, which provides superior charges and torsional parameters as compared to earlier alternatives.
- 8Paliwal, H.; Shirts, M. R. A benchmark test set for alchemical free energy transformations and its use to quantify error in common free energy methods. J. Chem. Theory Comput. 2011, 7, 4115– 4134, DOI: 10.1021/ct2003995Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtlaqtrvM&md5=41383373300216165a7f1724ad535451A Benchmark Test Set for Alchemical Free Energy Transformations and Its Use to Quantify Error in Common Free Energy MethodsPaliwal, Himanshu; Shirts, Michael R.Journal of Chemical Theory and Computation (2011), 7 (12), 4115-4134CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)There is a significant need for improved tools to validate thermophys. quantities computed via mol. simulation. In this paper we present the initial version of a benchmark set of testing methods for calcg. free energies of mol. transformation in soln. This set is based on mol. changes common to many mol. design problems, such as insertion and deletion of at. sites and changing at. partial charges. We use this benchmark set to compare the statistical efficiency, reliability, and quality of uncertainty ests. for a no. of published free energy methods, including thermodn. integration, free energy perturbation, the Bennett acceptance ratio (BAR) and its multistate equiv. MBAR. We identify MBAR as the consistently best performing method, though other methods are frequently comparable in reliability and accuracy in many cases. We demonstrate that assumptions of Gaussian distributed errors in free energies are usually valid for most methods studied. We demonstrate that bootstrap error estn. is a robust and useful technique for estg. statistical variance for all free energy methods studied. This benchmark set is provided in a no. of different file formats with the hope of becoming a useful and general tool for method comparisons.
- 9Shirts, M. R.; Chodera, J. D. Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 2008, 129, 124105, DOI: 10.1063/1.2978177Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXht1WnsL7F&md5=479183e1f45fc58dd7c6e5ef1e73d45dStatistically optimal analysis of samples from multiple equilibrium statesShirts, Michael R.; Chodera, John D.Journal of Chemical Physics (2008), 129 (12), 124105/1-124105/10CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We present a new estimator for computing free energy differences and thermodn. expectations as well as their uncertainties from samples obtained from multiple equil. states via either simulation or expt. The estimator, which we call the multistate Bennett acceptance ratio estimator (MBAR) because it reduces to the Bennett acceptance ratio estimator (BAR) when only two states are considered, has significant advantages over multiple histogram reweighting methods for combining data from multiple states. It does not require the sampled energy range to be discretized to produce histograms, eliminating bias due to energy binning and significantly reducing the time complexity of computing a soln. to the estg. equations in many cases. Addnl., an est. of the statistical uncertainty is provided for all estd. quantities. In the large sample limit, MBAR is unbiased and has the lowest variance of any known estimator for making use of equil. data collected from multiple states. We illustrate this method by producing a highly precise est. of the potential of mean force for a DNA hairpin system, combining data from multiple optical tweezer measurements under const. force bias. (c) 2008 American Institute of Physics.
- 10Bhati, A. P.; Wan, S.; Wright, D. W.; Coveney, P. V. Rapid, accurate, precise, and reliable relative free energy prediction using ensemble based thermodynamic integration. J. Chem. Theory Comput. 2017, 13, 210– 222, DOI: 10.1021/acs.jctc.6b00979Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitVGntrrJ&md5=510b70188112a3578030e291ce19b127Rapid, Accurate, Precise, and Reliable Relative Free Energy Prediction Using Ensemble Based Thermodynamic IntegrationBhati, Agastya P.; Wan, Shunzhou; Wright, David W.; Coveney, Peter V.Journal of Chemical Theory and Computation (2017), 13 (1), 210-222CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The accurate prediction of the binding affinities of ligands to proteins is a major goal in drug discovery and personalized medicine. The time taken to make such predictions is of similar importance to their accuracy, precision and reliability. In the last few years, an ensemble based mol. dynamics approach has been proposed that provides a route to reliable predictions of free energies based on the mol. mechanics Poisson-Boltzmann surface area method which meets the requirements of accuracy, precision and reliability. Here, we describe an equiv. methodol. based on thermodn. integration to substantially improve the accuracy, precision and reliability of calcd. relative binding free energies. We report the performance of the method when applied to a diverse set of protein targets and ligands. The results are in very good agreement with exptl. data (90% of calcns. agree to within 1 kcal/mol) while the method is reproducible by construction. Statistical uncertainties of the order of 0.5 kcal/mol or less are achieved. We present a systematic account of how the uncertainty in the predictions may be estd.
- 11Bhati, A. P.; Wan, S.; Hu, Y.; Sherborne, B.; Coveney, P. V. Uncertainty quantification in alchemical free energy methods. J. Chem. Theory Comput. 2018, 14, 2867– 2880, DOI: 10.1021/acs.jctc.7b01143Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXotFaqtr4%253D&md5=5f80e23df9c8aafb58914af06de24e1fUncertainty Quantification in Alchemical Free Energy MethodsBhati, Agastya P.; Wan, Shunzhou; Hu, Yuan; Sherborne, Brad; Coveney, Peter V.Journal of Chemical Theory and Computation (2018), 14 (6), 2867-2880CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Alchem. free energy methods have gained much importance recently from several reports of improved ligand-protein binding affinity predictions based on their implementation using mol. dynamics simulations. A large no. of variants of such methods implementing different accelerated sampling techniques and free energy estimators are available, each claimed to be better than the others in its own way. However, the key features of reproducibility and quantification of assocd. uncertainties in such methods have barely been discussed. Here, we apply a systematic protocol for uncertainty quantification to a no. of popular alchem. free energy methods, covering both abs. and relative free energy predictions. We show that a reliable measure of error estn. is provided by ensemble simulation-an ensemble of independent MD simulations-which applies irresp. of the free energy method. The need to use ensemble methods is fundamental and holds regardless of the duration of time of the mol. dynamics simulations performed.
- 12Wan, S.; Bhati, A. P.; Skerratt, S.; Omoto, K.; Shanmugasundaram, V.; Bagal, S. K.; Coveney, P. V. Evaluation and characterization of Trk kinase inhibitors for the treatment of pain: reliable binding affinity predictions from theory and computation. J. Chem. Inf. Model. 2017, 57, 897– 909, DOI: 10.1021/acs.jcim.6b00780Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXksVyrurc%253D&md5=2e366f56c1ee8c2f2b2760b9098e2030Evaluation and Characterization of Trk Kinase Inhibitors for the Treatment of Pain: Reliable Binding Affinity Predictions from Theory and ComputationWan, Shunzhou; Bhati, Agastya P.; Skerratt, Sarah; Omoto, Kiyoyuki; Shanmugasundaram, Veerabahu; Bagal, Sharan K.; Coveney, Peter V.Journal of Chemical Information and Modeling (2017), 57 (4), 897-909CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Optimization of ligand binding affinity to the target protein of interest is a primary objective in small-mol. drug discovery. Until now, the prediction of binding affinities by computational methods has not been widely applied in the drug discovery process, mainly due to its lack of accuracy and reproducibility, as well as the long turnaround times required to obtain results. Herein, the authors report on a collaborative study that compares tropomyosin receptor kinase A (TrkA) binding affinity predictions using two recently formulated fast computational approaches - namely ESMACS (Enhanced Sampling of Mol. dynamics with Approxn. of Continuum Solvent) and TIES (Thermodn. Integration with Enhanced Sampling) - to exptl. derived TrkA binding affinities for a set of Pfizer pan-Trk compds. ESMACS gives precise and reproducible results and is applicable to highly diverse sets of compds. It also provides detailed chem. insight into the nature of ligand-protein binding. TIES can predict and thus optimize more subtle changes in binding affinities between compds. of similar structure. Individual binding affinities were calcd. in a few hours, exhibiting good correlations with the exptl. data of 0.79 and 0.88 from ESMACS and TIES approaches resp. The speed, level of accuracy and precision of the calcns. are such that the affinity predictions can be used to rapidly explain the effects of compd. modifications on TrkA binding affinity. The methods could therefore be used as tools to guide lead optimization efforts across multiple prospective structurally-enabled programs in the drug discovery setting for a wide range of compds. and targets.
- 13Wan, 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, 784– 795, DOI: 10.1021/acs.jctc.6b00794Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitFektLnF&md5=713436084662420482684fb50db0832eRapid and Reliable Binding Affinity Prediction of Bromodomain Inhibitors: A Computational StudyWan, Shunzhou; Bhati, Agastya P.; Zasada, Stefan J.; Wall, Ian; Green, Darren; Bamborough, Paul; Coveney, Peter V.Journal of Chemical Theory and Computation (2017), 13 (2), 784-795CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Binding free energies of bromodomain inhibitors are calcd. with recently formulated approaches, namely ESMACS (enhanced sampling of mol. dynamics with approxn. of continuum solvent) and TIES (thermodn. integration with enhanced sampling). A set of compds. is provided by GlaxoSmithKline, which represents a range of chem. functionality and binding affinities. The predicted binding free energies exhibit a good Spearman correlation of 0.78 with the exptl. data from the 3-trajectory ESMACS, and an excellent correlation of 0.92 from the TIES approach where applicable. Given access to suitable high end computing resources and a high degree of automation, the authors can compute individual binding affinities in a few hours with precisions no greater than 0.2 kcal/mol for TIES, and no larger than 0.34 kcal/mol and 1.71 kcal/mol for the 1- and 3-trajectory ESMACS approaches.
- 14Porta, R.; Borea, R.; Coelho, A.; Khan, S.; Araújo, A.; Reclusa, P.; Franchina, T.; Steen, N. V. D.; Dam, P. V.; Ferri, J.; Sirera, R.; Naing, A.; Hong, D.; Rolfo, C. FGFR a promising druggable target in cancer: Molecular biology and new drugs. Crit. Rev. Oncol. Hematol. 2017, 113, 256– 267, DOI: 10.1016/j.critrevonc.2017.02.018Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC1cvps1ajsA%253D%253D&md5=d03086e258264a12c8c5cc7f2c5de8e3FGFR a promising druggable target in cancer: Molecular biology and new drugsPorta Rut; Borea Roberto; Coelho Andreia; Reclusa Pablo; Van Dam Peter; Ferri Jose; Sirera Rafael; Khan Shahanavaj; Araujo Antonio; Franchina Tindara; Van Der Steen Nele; Naing Aung; Hong David; Rolfo ChristianCritical reviews in oncology/hematology (2017), 113 (), 256-267 ISSN:.INTRODUCTION: The Fibroblast Growth Factor Receptor (FGFR) family consists of Tyrosine Kinase Receptors (TKR) involved in several biological functions. Recently, alterations of FGFR have been reported to be important for progression and development of several cancers. In this setting, different studies are trying to evaluate the efficacy of different therapies targeting FGFR. AREAS COVERED: This review summarizes the current status of treatments targeting FGFR, focusing on the trials that are evaluating the FGFR profile as inclusion criteria: Multi-Target, Pan-FGFR Inhibitors and anti-FGF (Fibroblast Growth Factor)/FGFR Monoclonal Antibodies. EXPERT OPINION: Most of the TKR share intracellular signaling pathways; therefore, cancer cells tend to overcome the inhibition of one tyrosine kinase receptor by activating another. The future of TKI (Tyrosine Kinase Inhibitor) therapy will potentially come from multi-targeted TKIs that target different TKR simultaneously. It is crucial to understand the interaction of the FGF-FGFR axis with other known driver TKRs. Based on this, it is possible to develop therapeutic strategies targeting multiple connected TKRs at once. One correct step in this direction is the reassessment of multi target inhibitors considering the FGFR status of the tumor. Another opportunity arises from assessing the use of FGFR TKI on patients harboring FGFR alterations.
- 15Bunney, T. D.; Wan, S.; Thiyagarajan, N.; Sutto, L.; Williams, S. V.; Ashford, P.; Koss, H.; Knowles, M. a.; Gervasio, F. L.; Coveney, P. V.; Katan, M. The effect of mutations on drug sensitivity and kinase activity of fibroblast growth factor receptors: A combined experimental and theoretical study. EBioMedicine 2015, 2, 194– 204, DOI: 10.1016/j.ebiom.2015.02.009Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2srotlKmtA%253D%253D&md5=8414f3530d24bbba0ea8cc98cb784b96The Effect of Mutations on Drug Sensitivity and Kinase Activity of Fibroblast Growth Factor Receptors: A Combined Experimental and Theoretical StudyBunney Tom D; Thiyagarajan Nethaji; Ashford Paul; Katan Matilda; Wan Shunzhou; Coveney Peter V; Sutto Ludovico; Gervasio Francesco L; Williams Sarah V; Knowles Margaret A; Koss HansEBioMedicine (2015), 2 (3), 194-204 ISSN:.Fibroblast growth factor receptors (FGFRs) are recognized therapeutic targets in cancer. We here describe insights underpinning the impact of mutations on FGFR1 and FGFR3 kinase activity and drug efficacy, using a combination of computational calculations and experimental approaches including cellular studies, X-ray crystallography and biophysical and biochemical measurements. Our findings reveal that some of the tested compounds, in particular TKI258, could provide therapeutic opportunity not only for patients with primary alterations in FGFR but also for acquired resistance due to the gatekeeper mutation. The accuracy of the computational methodologies applied here shows a potential for their wider application in studies of drug binding and in assessments of functional and mechanistic impacts of mutations, thus assisting efforts in precision medicine.
- 16Patani, H.; Bunney, T. D.; Thiyagarajan, N.; Norman, R. A.; Ogg, D.; Breed, J.; Ashford, P.; Potterton, A.; Edwards, M.; Katan, M. Landscape of activating cancer mutations in FGFR kinases and their differential responses to inhibitors in clinical use. Oncotarget 2016, 7, 24252– 24268, DOI: 10.18632/oncotarget.8132Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28fhsFShsw%253D%253D&md5=beb65006a52ef6062bab9c9d8086bf1dLandscape of activating cancer mutations in FGFR kinases and their differential responses to inhibitors in clinical usePatani Harshnira; Bunney Tom D; Thiyagarajan Nethaji; Ashford Paul; Potterton Andrew; Edwards Mina; Pang Camilla S M; Orengo Christine; Katan Matilda; Norman Richard A; Ogg Derek; Breed Jason; Phillips Chris; Williams Sarah V; Knowles Margaret A; Thomson Gary S; Breeze Alexander LOncotarget (2016), 7 (17), 24252-68 ISSN:.Frequent genetic alterations discovered in FGFRs and evidence implicating some as drivers in diverse tumors has been accompanied by rapid progress in targeting FGFRs for anticancer treatments. Wider assessment of the impact of genetic changes on the activation state and drug responses is needed to better link the genomic data and treatment options. We here apply a direct comparative and comprehensive analysis of FGFR3 kinase domain variants representing the diversity of point-mutations reported in this domain. We reinforce the importance of N540K and K650E and establish that not all highly activating mutations (for example R669G) occur at high-frequency and conversely, that some "hotspots" may not be linked to activation. Further structural characterization consolidates a mechanistic view of FGFR kinase activation and extends insights into drug binding. Importantly, using several inhibitors of particular clinical interest (AZD4547, BGJ-398, TKI258, JNJ42756493 and AP24534), we find that some activating mutations (including different replacements of the same residue) result in distinct changes in their efficacy. Considering that there is no approved inhibitor for anticancer treatments based on FGFR-targeting, this information will be immediately translatable to ongoing clinical trials.
- 17Azam, M.; Seeliger, M. a.; Gray, N. S.; Kuriyan, J.; Daley, G. Q. Activation of tyrosine kinases by mutation of the gatekeeper threonine. Nat. Struct. Mol. Biol. 2008, 15, 1109– 1118, DOI: 10.1038/nsmb.1486Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXht1Sisb%252FM&md5=4afc4e5a0850a4510ced3fa5fbda97faActivation of tyrosine kinases by mutation of the gatekeeper threonineAzam, Mohammad; Seeliger, Markus A.; Gray, Nathanael S.; Kuriyan, John; Daley, George Q.Nature Structural & Molecular Biology (2008), 15 (10), 1109-1118CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Protein kinases targeted by small-mol. inhibitors develop resistance through mutation of the 'gatekeeper' threonine residue of the active site. Here we show that the gatekeeper mutation in the cellular forms of c-ABL, c-SRC, platelet-derived growth factor receptor-α and -β, and epidermal growth factor receptor activates the kinase and promotes malignant transformation of BaF3 cells. Structural anal. reveals that a network of hydrophobic interactions - the hydrophobic spine - characteristic of the active kinase conformation is stabilized by the gatekeeper substitution. Substitution of glycine for the residues constituting the spine disrupts the hydrophobic connectivity and inactivates the kinase. Furthermore, a small-mol. inhibitor (compd. 14) that maximizes complementarity with the dismantled spine inhibits the gatekeeper mutation of BCR-ABL-T315I. These results demonstrate that mutation of the gatekeeper threonine is a common mechanism of activation for tyrosine kinases and provide structural insights to guide the development of next-generation inhibitors.
- 18van Gunsteren, W. F.; Daura, X.; Hansen, N.; Mark, A. E.; Oostenbrink, C.; Riniker, S.; Smith, L. J. Validation of Molecular Simulation: An Overview of Issues. Angew. Chem., Int. Ed. 2018, 57, 884– 902, DOI: 10.1002/anie.201702945Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhs1ertQ%253D%253D&md5=d38b0f3fd0054e0fab46023a2c597442Validation of Molecular Simulation: An Overview of Issuesvan Gunsteren, Wilfred F.; Daura, Xavier; Hansen, Niels; Mark, Alan E.; Oostenbrink, Chris; Riniker, Sereina; Smith, Lorna J.Angewandte Chemie, International Edition (2018), 57 (4), 884-902CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Computer simulation of mol. systems enables structure-energy-function relationships of mol. processes to be described at the sub-at., at., supra-at., or supra-mol. level. To interpret results of such simulations appropriately, the quality of the calcd. properties must be evaluated. This depends on the way the simulations are performed and on the way they are validated by comparison to values Qexp of exptl. observable quantities Q. One must consider (1) the accuracy of Qexp, (2) the accuracy of the function Q(rN) used to calc. a Q-value based on a mol. configuration rN of N particles, (3) the sensitivity of the function Q(rN) to the configuration rN, (4) the relative time scales of the simulation and expt., (5) the degree to which the calcd. and exptl. properties are equiv., and (6) the degree to which the system simulated matches the exptl. conditions. Exptl. data is limited in scope and generally corresponds to avs. over both time and space. A crit. anal. of the various factors influencing the apparent degree of (dis)agreement between simulations and expt. is presented and illustrated using examples from the literature. What can be done to enhance the validation of mol. simulation is also discussed.
- 19Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kal, L.; Schulten, K. Scalable molecular dynamics with NAMD. J. Comput. Chem. 2005, 26, 1781– 1802, DOI: 10.1002/jcc.20289Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SlsbbJ&md5=189051128443b547f4300a1b8fb0e034Scalable molecular dynamics with NAMDPhillips, James C.; Braun, Rosemary; Wang, Wei; Gumbart, James; Tajkhorshid, Emad; Villa, Elizabeth; Chipot, Christophe; Skeel, Robert D.; Kale, Laxmikant; Schulten, KlausJournal of Computational Chemistry (2005), 26 (16), 1781-1802CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)NAMD is a parallel mol. dynamics code designed for high-performance simulation of large biomol. systems. NAMD scales to hundreds of processors on high-end parallel platforms, as well as tens of processors on low-cost commodity clusters, and also runs on individual desktop and laptop computers. NAMD works with AMBER and CHARMM potential functions, parameters, and file formats. This article, directed to novices as well as experts, first introduces concepts and methods used in the NAMD program, describing the classical mol. dynamics force field, equations of motion, and integration methods along with the efficient electrostatics evaluation algorithms employed and temp. and pressure controls used. Features for steering the simulation across barriers and for calcg. both alchem. and conformational free energy differences are presented. The motivations for and a roadmap to the internal design of NAMD, implemented in C++ and based on Charm++ parallel objects, are outlined. The factors affecting the serial and parallel performance of a simulation are discussed. Finally, typical NAMD use is illustrated with representative applications to a small, a medium, and a large biomol. system, highlighting particular features of NAMD, for example, the Tcl scripting language. The article also provides a list of the key features of NAMD and discusses the benefits of combining NAMD with the mol. graphics/sequence anal. software VMD and the grid computing/collab. software BioCoRE. NAMD is distributed free of charge with source code at www.ks.uiuc.edu.
- 20Coveney, P. V.; Wan, S. On the calculation of equilibrium thermodynamic properties from molecular dynamics. Phys. Chem. Chem. Phys. 2016, 18, 30236– 30240, DOI: 10.1039/C6CP02349EGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XntVyisr8%253D&md5=f7950af8255cf494a117d46bc1b5191eOn the calculation of equilibrium thermodynamic properties from molecular dynamicsCoveney, Peter V.; Wan, ShunzhouPhysical Chemistry Chemical Physics (2016), 18 (44), 30236-30240CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The purpose of statistical mechanics is to provide a route to the calcn. of macroscopic properties of matter from their constituent microscopic components. It is well known that the macrostates emerge as ensemble avs. of microstates. However, this is more often stated than implemented in computer simulation studies. Here we consider foundational aspects of statistical mechanics which are overlooked in most textbooks and research articles that purport to compute macroscopic behavior from microscopic descriptions based on classical mechanics and show how due attention to these issues leads in directions which have not been widely appreciated in the field of mol. dynamics simulation.
- 21Lim, N. M.; Wang, L.; Abel, R.; Mobley, D. L. Sensitivity in binding free energies due to protein reorganization. J. Chem. Theory Comput. 2016, 12, 4620– 4631, DOI: 10.1021/acs.jctc.6b00532Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xht1CltrbO&md5=5f4525fb7a55b6c9ad0510168e13f316Sensitivity in Binding Free Energies Due to Protein ReorganizationLim, Nathan M.; Wang, Lingle; Abel, Robert; Mobley, David L.Journal of Chemical Theory and Computation (2016), 12 (9), 4620-4631CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Tremendous recent improvements in computer hardware, coupled with advances in sampling techniques and force fields, are now allowing protein-ligand binding free energy calcns. to be routinely used to aid pharmaceutical drug discovery projects. However, despite these recent innovations, there are still needs for further improvement in sampling algorithms to more adequately sample protein motion relevant to protein-ligand binding. Here, we report our work identifying and studying such clear and remaining needs in the apolar cavity of T4 lysozyme L99A. In this study, we model recent exptl. results that show the progressive opening of the binding pocket in response to a series of homologous ligands. Even while using enhanced sampling techniques, we demonstrate that the predicted relative binding free energies (RBFE) are sensitive to the initial protein conformational state. Particularly, we highlight the importance of sufficient sampling of protein conformational changes and demonstrate how inclusion of three key protein residues in the "hot" region of the FEP/REST simulation improves the sampling and resolves this sensitivity, given enough simulation time.
- 22Clark, A. J.; Gindin, T.; Zhang, B.; Wang, L.; Abel, R.; Murret, C. S.; Xu, F.; Bao, A.; Lu, N. J.; Zhou, T.; Kwong, P. D.; Shapiro, L.; Honig, B.; Friesner, R. A. Free energy perturbation calculation of relative binding free energy between broadly neutralizing antibodies and the gp120 glycoprotein of HIV-1. J. Mol. Biol. 2017, 429, 930– 947, DOI: 10.1016/j.jmb.2016.11.021Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitVSmsL3I&md5=ff264d642726916a45bae5b121bbdd06Free energy perturbation calculation of relative binding free energy between broadly neutralizing antibodies and the gp120 glycoprotein of HIV-1Clark, Anthony J.; Gindin, Tatyana; Zhang, Baoshan; Wang, Lingle; Abel, Robert; Murret, Colleen S.; Xu, Fang; Bao, Amy; Lu, Nina J.; Zhou, Tongqing; Kwong, Peter D.; Shapiro, Lawrence; Honig, Barry; Friesner, Richard A.Journal of Molecular Biology (2017), 429 (7), 930-947CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)Direct calcn. of relative binding affinities between antibodies and antigens is a long-sought goal. However, despite substantial efforts, no generally applicable computational method has been described. Here, we describe a systematic free energy perturbation (FEP) protocol and calc. the binding affinities between the gp120 envelope glycoprotein of HIV-1 and three broadly neutralizing antibodies (bNAbs) of the VRC01 class. The protocol has been adapted from successful studies of small mols. to address the challenges assocd. with modeling protein-protein interactions. Specifically, we built homol. models of the three antibody-gp120 complexes, extended the sampling times for large bulky residues, incorporated the modeling of glycans on the surface of gp120, and utilized continuum solvent-based loop prediction protocols to improve sampling. We present three exptl. surface plasmon resonance data sets, in which antibody residues in the antibody/gp120 interface were systematically mutated to alanine. The RMS error in the large set (55 total cases) of FEP tests as compared to these expts., 0.68 kcal/mol, is near exptl. accuracy, and it compares favorably with the results obtained from a simpler, empirical methodol. The correlation coeff. for the combined data set including residues with glycan contacts, R2 = 0.49, should be sufficient to guide the choice of residues for antibody optimization projects, assuming that this level of accuracy can be realized in prospective prediction. More generally, these results are encouraging with regard to the possibility of using an FEP approach to calc. the magnitude of protein-protein binding affinities.
- 23Fratev, F.; Sirimulla, S. An improved free energy perturbation FEP+ sampling protocol for flexible ligand-binding domains. ChemRxiv 2018, DOI: 10.26434/chemrxiv.6204167.v1 .Google ScholarThere is no corresponding record for this reference.
- 24Fiser, A.; Sali, A. ModLoop: automated modeling of loops in protein structures. Bioinformatics 2003, 19, 2500– 2501, DOI: 10.1093/bioinformatics/btg362Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXpvVSisLk%253D&md5=4b0fdaca0a412715682496883ab2019cModLoop: automated modeling of loops in protein structuresFiser, Andras; Sali, AndrejBioinformatics (2003), 19 (18), 2500-2501CODEN: BOINFP; ISSN:1367-4803. (Oxford University Press)Summary: ModLoop is a web server for automated modeling of loops in protein structures. The input is the at. coordinates of the protein structure in the Protein Data Bank format, and the specification of the starting and ending residues of one or more segments to be modeled, contg. no more than 20 residues in total. The output is the coordinates of the non-hydrogen atoms in the modeled segments. A user provides the input to the server via a simple web interface, and receives the output by e-mail. The server relies on the loop modeling routine in MODELLER that predicts the loop conformations by satisfaction of spatial restraints, without relying on a database of known protein structures. For a rapid response, ModLoop runs on a cluster of Linux PC computers.
- 25Sohl, C. D.; Ryan, M. R.; Luo, B.; Frey, K. M.; Anderson, K. S. Illuminating the molecular mechanisms of tyrosine kinase inhibitor resistance for the FGFR1 gatekeeper mutation: the Achilles’ heel of targeted therapy. ACS Chem. Biol. 2015, 10, 1319– 1329, DOI: 10.1021/acschembio.5b00014Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXislGitLo%253D&md5=6c13f193e9e4c3f38b1e59f46fbc6342Illuminating the Molecular Mechanisms of Tyrosine Kinase Inhibitor Resistance for the FGFR1 Gatekeeper Mutation: The Achilles' Heel of Targeted TherapySohl, Christal D.; Ryan, Molly R.; Luo, BeiBei; Frey, Kathleen M.; Anderson, Karen S.ACS Chemical Biology (2015), 10 (5), 1319-1329CODEN: ACBCCT; ISSN:1554-8929. (American Chemical Society)Human fibroblast growth factor receptors (FGFRs) 1-4 are a family of receptor tyrosine kinases that can serve as drivers of tumorigenesis. In particular, FGFR1 gene amplification has been implicated in squamous cell lung and breast cancers. Tyrosine kinase inhibitors (TKIs) targeting FGFR1, including AZD4547 and E3810 (Lucitanib), are currently in early phase clin. trials. Unfortunately, drug resistance limits the long-term success of TKIs, with mutations at the "gatekeeper" residue leading to tumor progression. Here we show the first structural and kinetic characterization of the FGFR1 gatekeeper mutation, V561M FGFR1. The V561M mutation confers a 38-fold increase in autophosphorylation achieved at least in part by a network of interacting residues forming a hydrophobic spine to stabilize the active conformation. Moreover, kinetic assays established that the V561M mutation confers significant resistance to E3810, while retaining affinity for AZD4547. Structural analyses of these TKIs with wild type (WT) and gatekeeper mutant forms of FGFR1 offer clues to developing inhibitors that maintain potency against gatekeeper mutations. We show that AZD4547 affinity is preserved by V561M FGFR1 due to a flexible linker that allows multiple inhibitor binding modes. This is the first example of a TKI binding in distinct conformations to WT and gatekeeper mutant forms of FGFR, highlighting adaptable regions in both the inhibitor and binding pocket crucial for drug design. Exploiting inhibitor flexibility to overcome drug resistance has been a successful strategy for combating diseases such as AIDS and may be an important approach for designing inhibitors effective against kinase gatekeeper mutations.
- 26Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03; Gaussian, Inc.: Wallingford, CT, 2004.Google ScholarThere is no corresponding record for this reference.
- 27Case, D. A.; Cheatham, T. E.; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. J. The Amber biomolecular simulation programs. J. Comput. Chem. 2005, 26, 1668– 1688, DOI: 10.1002/jcc.20290Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SlsbbM&md5=93be29ff894bab96c783d24e9886c7d0The amber biomolecular simulation programsCase, David A.; Cheatham, Thomas E., III; Darden, Tom; Gohlke, Holger; Luo, Ray; Merz, Kenneth M., Jr.; Onufriev, Alexey; Simmerling, Carlos; Wang, Bing; Woods, Robert J.Journal of Computational Chemistry (2005), 26 (16), 1668-1688CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The authors describe the development, current features, and some directions for future development of the Amber package of computer programs. This package evolved from a program that was constructed in the late 1970s to do Assisted Model Building with Energy Refinement, and now contains a group of programs embodying a no. of powerful tools of modern computational chem., focused on mol. dynamics and free energy calcns. of proteins, nucleic acids, and carbohydrates.
- 28Wang, 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, 1157– 1174, DOI: 10.1002/jcc.20035Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksFakurc%253D&md5=2992017a8cf51f89290ae2562403b115Development and testing of a general Amber force fieldWang, Junmei; Wolf, Romain M.; Caldwell, James W.; Kollman, Peter A.; Case, David A.Journal of Computational Chemistry (2004), 25 (9), 1157-1174CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)We describe here a general Amber force field (GAFF) for org. mols. GAFF is designed to be compatible with existing Amber force fields for proteins and nucleic acids, and has parameters for most org. and pharmaceutical mols. that are composed of H, C, N, O, S, P, and halogens. It uses a simple functional form and a limited no. of atom types, but incorporates both empirical and heuristic models to est. force consts. and partial at. charges. The performance of GAFF in test cases is encouraging. In test I, 74 crystallog. structures were compared to GAFF minimized structures, with a root-mean-square displacement of 0.26 Å, which is comparable to that of the Tripos 5.2 force field (0.25 Å) and better than those of MMFF 94 and CHARMm (0.47 and 0.44 Å, resp.). In test II, gas phase minimizations were performed on 22 nucleic acid base pairs, and the minimized structures and intermol. energies were compared to MP2/6-31G* results. The RMS of displacements and relative energies were 0.25 Å and 1.2 kcal/mol, resp. These data are comparable to results from Parm99/RESP (0.16 Å and 1.18 kcal/mol, resp.), which were parameterized to these base pairs. Test III looked at the relative energies of 71 conformational pairs that were used in development of the Parm99 force field. The RMS error in relative energies (compared to expt.) is about 0.5 kcal/mol. GAFF can be applied to wide range of mols. in an automatic fashion, making it suitable for rational drug design and database searching.
- 29Maier, 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, 3696– 3713, DOI: 10.1021/acs.jctc.5b00255Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFequ7rN&md5=7b803577b3b6912cc6750cfbd356596eff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SBMaier, James A.; Martinez, Carmenza; Kasavajhala, Koushik; Wickstrom, Lauren; Hauser, Kevin E.; Simmerling, CarlosJournal of Chemical Theory and Computation (2015), 11 (8), 3696-3713CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Mol. mechanics is powerful for its speed in atomistic simulations, but an accurate force field is required. The Amber ff99SB force field improved protein secondary structure balance and dynamics from earlier force fields like ff99, but weaknesses in side chain rotamer and backbone secondary structure preferences have been identified. Here, we performed a complete refit of all amino acid side chain dihedral parameters, which had been carried over from ff94. The training set of conformations included multidimensional dihedral scans designed to improve transferability of the parameters. Improvement in all amino acids was obtained as compared to ff99SB. Parameters were also generated for alternate protonation states of ionizable side chains. Av. errors in relative energies of pairs of conformations were under 1.0 kcal/mol as compared to QM, reduced 35% from ff99SB. We also took the opportunity to make empirical adjustments to the protein backbone dihedral parameters as compared to ff99SB. Multiple small adjustments of φ and ψ parameters were tested against NMR scalar coupling data and secondary structure content for short peptides. The best results were obtained from a phys. motivated adjustment to the φ rotational profile that compensates for lack of ff99SB QM training data in the β-ppII transition region. Together, these backbone and side chain modifications (hereafter called ff14SB) not only better reproduced their benchmarks, but also improved secondary structure content in small peptides and reprodn. of NMR χ1 scalar coupling measurements for proteins in soln. We also discuss the Amber ff12SB parameter set, a preliminary version of ff14SB that includes most of its improvements.
- 30Jo, S.; Jiang, W. A generic implementation of replica exchange with solute tempering (REST2) algorithm in NAMD for complex biophysical simulations. Comput. Phys. Commun. 2015, 197, 304– 311, DOI: 10.1016/j.cpc.2015.08.030Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsFSlu73I&md5=776ea438c536bf84540b1443494e647cA generic implementation of replica exchange with solute tempering (REST2) algorithm in NAMD for complex biophysical simulationsJo, Sunhwan; Jiang, WeiComputer Physics Communications (2015), 197 (), 304-311CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier B.V.)Replica Exchange with Solute Tempering (REST2) is a powerful sampling enhancement algorithm of mol. dynamics (MD) in that it needs significantly smaller no. of replicas but achieves higher sampling efficiency relative to std. temp. exchange algorithm. In this paper, we extend the applicability of REST2 for quant. biophys. simulations through a robust and generic implementation in greatly scalable MD software NAMD. The rescaling procedure of force field parameters controlling REST2 "hot region" is implemented into NAMD at the source code level. A user can conveniently select hot region through VMD and write the selection information into a PDB file. The rescaling keyword/parameter is written in NAMD Tcl script interface that enables an on-the-fly simulation parameter change. Our implementation of REST2 is within communication-enabled Tcl script built on top of Charm++, thus communication overhead of an exchange attempt is vanishingly small. Such a generic implementation facilitates seamless cooperation between REST2 and other modules of NAMD to provide enhanced sampling for complex biomol. simulations. Three challenging applications including native REST2 simulation for peptide folding-unfolding transition, free energy perturbation/REST2 for abs. binding affinity of protein-ligand complex and umbrella sampling/REST2 Hamiltonian exchange for free energy landscape calcn. were carried out on IBM Blue Gene/Q supercomputer to demonstrate efficacy of REST2 based on the present implementation.
- 31Ballester, P. J.; Mitchell, J. B. O. A machine learning approach to predicting protein–ligand binding affinity with applications to molecular docking. Bioinformatics 2010, 26, 1169– 1175, DOI: 10.1093/bioinformatics/btq112Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXlt1Cjs78%253D&md5=6d2d48344dc5ba70f0ff85f432d0ec63A machine learning approach to predicting protein-ligand binding affinity with applications to molecular dockingBallester, Pedro J.; Mitchell, John B. O.Bioinformatics (2010), 26 (9), 1169-1175CODEN: BOINFP; ISSN:1367-4803. (Oxford University Press)Accurately predicting the binding affinities of large sets of diverse protein-ligand complexes is an extremely challenging task. The scoring functions that attempt such computational prediction are essential for analyzing the outputs of mol. docking, which in turn is an important technique for drug discovery, chem. biol. and structural biol. Each scoring function assumes a predetd. theory-inspired functional form for the relationship between the variables that characterize the complex, which also include parameters fitted to exptl. or simulation data and its predicted binding affinity. The inherent problem of this rigid approach is that it leads to poor predictivity for those complexes that do not conform to the modeling assumptions. Moreover, resampling strategies, such as cross-validation or bootstrapping, are still not systematically used to guard against the overfitting of calibration data in parameter estn. for scoring functions. We propose a novel scoring function (RF-Score) that circumvents the need for problematic modeling assumptions via non-parametric machine learning. In particular, Random Forest was used to implicitly capture binding effects that are hard to model explicitly. RF-Score is compared with the state of the art on the demanding PDBbind benchmark. Results show that RF-Score is a very competitive scoring function. Importantly, RF-Score's performance was shown to improve dramatically with training set size and hence the future availability of more high-quality structural and interaction data is expected to lead to improved versions of RF-Score.
- 32Jimenez, J.; Skalic, M.; Martinez-Rosell, G.; De Fabritiis, G. KDEEP: Protein–Ligand Absolute Binding Affinity Prediction via 3D-Convolutional Neural Networks. J. Chem. Inf. Model. 2018, 58, 287– 296, DOI: 10.1021/acs.jcim.7b00650Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXkslSitQ%253D%253D&md5=81943a6732be99e5439e1e3a25fa4414KDEEP: Protein-Ligand Absolute Binding Affinity Prediction via 3D-Convolutional Neural NetworksJimenez, Jose; Skalic, Miha; Martinez-Rosell, Gerard; De Fabritiis, GianniJournal of Chemical Information and Modeling (2018), 58 (2), 287-296CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Accurately predicting protein-ligand binding affinities is an important problem in computational chem. since it can substantially accelerate drug discovery for virtual screening and lead optimization. We propose here a fast machine-learning approach for predicting binding affinities using state-of-the-art 3D-convolutional neural networks and compare this approach to other machine-learning and scoring methods using several diverse data sets. The results for the std. PDBbind (v.2016) core test-set are state-of-the-art with a Pearson's correlation coeff. of 0.82 and a RMSE of 1.27 in pK units between exptl. and predicted affinity, but accuracy is still very sensitive to the specific protein used. KDEEP is made available via PlayMol.org for users to test easily their own protein-ligand complexes, with each prediction taking a fraction of a second. We believe that the speed, performance, and ease of use of KDEEP makes it already an attractive scoring function for modern computational chem. pipelines.
- 33Coveney, P. V.; Dougherty, E. R.; Highfield, R. R. Big data need big theory too. Philos. Trans. R. Soc., A 2016, 374, 20160153, DOI: 10.1098/rsta.2016.0153Google ScholarThere is no corresponding record for this reference.
Supporting Information
Supporting Information
ARTICLE SECTIONSThe Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jctc.8b01118.
Tables containing individual ΔGalch values for all the relative free energy calculations, and a table for the comparison of occupancies of hydrogen bonds between the inhibitors and the protein (PDF)
Topology and coordinate files for all ligand–protein complexes (ZIP)
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.