Skip to main content
Interface Focus logoLink to Interface Focus
. 2020 Oct 16;10(6):20190128. doi: 10.1098/rsfs.2019.0128

Hit-to-lead and lead optimization binding free energy calculations for G protein-coupled receptors

Shunzhou Wan 1,, Andrew Potterton 2,, Fouad S Husseini 1,, David W Wright 1, Alexander Heifetz 2,3, Maciej Malawski 4, Andrea Townsend-Nicholson 2, Peter V Coveney 1,5,
PMCID: PMC7653344  PMID: 33178414

Abstract

We apply the hit-to-lead ESMACS (enhanced sampling of molecular dynamics with approximation of continuum solvent) and lead-optimization TIES (thermodynamic integration with enhanced sampling) methods to compute the binding free energies of a series of ligands at the A1 and A2A adenosine receptors, members of a subclass of the GPCR (G protein-coupled receptor) superfamily. Our predicted binding free energies, calculated using ESMACS, show a good correlation with previously reported experimental values of the ligands studied. Relative binding free energies, calculated using TIES, accurately predict experimentally determined values within a mean absolute error of approximately 1 kcal mol−1. Our methodology may be applied widely within the GPCR superfamily and to other small molecule–receptor protein systems.

Keywords: molecular dynamics, free energy, binding affinity prediction, G protein-coupled receptors, adenosine receptors, ensemble simulations

1. Introduction

There is an urgent need for approaches and tools that permit the prediction of rapid, accurate and reliable properties of systems across science as a whole. We have a longstanding interest in the development of in silico methodologies able to predict values computationally that agree with and therefore may replace experimental measurements [14]. Here, we focus our efforts on a subject of global importance in computational biomedicine: the accurate prediction of protein–small molecule binding affinities. The calculation of accurate binding affinities will provide substantial insight into ligand–receptor interactions for scientists, significantly impact the drug discovery process in industry and expedite the implementation of personalized medicine, making it more commercially attractive and facilitating the development of bespoke, individualized, pharmaceuticals to significantly improve patient prospects and bring about economic savings for healthcare programmes. The prediction of binding free energies is a computationally tractable task that, with iteration, can provide a self-reinforcing loop between experimental data and theoretical calculations.

Binding free energy can be calculated using pathway or endpoint methods. A pathway can be either a physical binding path or an alchemical path. The former is usually defined by a suitable collective variable with which simulation is driven and free energy change is derived. Large conformational space needs to be sampled along the binding path, which usually requires enhanced sampling approaches. Recently a metadynamics protocol has been applied for ligand binding free energies to G protein-coupled receptors (GPCRs) [5,6], generating encouraging results. The approach is valuable to explore the pathways of ligand binding, to find the binding site(s), to predict the binding poses and to estimate binding free energies. However, such simulations require a timescale of microseconds and need to run for days if not weeks on high performance supercomputers. The rapid development of computational power may make it possible for these techniques to deliver actionable predictions within ten years. The metadynamics protocol, however, is unable to satisfy the requirement for pharmaceutical drug development today. Alchemical and endpoint approaches, on the other hand, are being increasingly promoted by pharmaceutical companies collectively [7] as they can be implemented in a rapid, accurate and reliable manner [8].

To accurately and reliably compute these values, it is necessary to appreciate that macromolecular biological systems are capable of adopting various conformations depending on how they are studied by simulation and that results obtained from single trajectory (one-off) simulations—particularly long ones—lack the accuracy and reproducibility needed for convergence with experimentally determined values [9]. This is only now becoming fully understood by many practitioners of molecular simulation. Despite this, single trajectory approaches that use methods such as MMPBSA (molecular mechanics Poisson Boltzmann surface area) [10], WaterMap (WM) [11], and the semi-empirical, linear interaction method (LIE) [12,13] have been used extensively. Although the benefit of using ensembles comprised of multiple simulations, or replicas, has been demonstrated [14], and applied to the calculation of free energies [15] it is only recently that new methods have been introduced which improve sampling and accessibility of the conformational space in molecular dynamics (MD)-produced trajectories [3,1618]. One of these, ESMACS [2,14,16,17,19] (enhanced sampling of molecular dynamics with approximation of continuum solvent), uses ensembles of multiple and typically relatively short duration simulations to calculate absolute binding free energies with high precision. There is a wealth of evidence in the literature and in unpublished work that, under equilibrium conditions, multiple short MD simulations sample better than a single long MD simulation and provide a meaningful uncertainty of the results [8,9,2022]. Under general non-equilibrium conditions, ensembles are essential since there is then no meaning to time averaging. Here, we use an ensemble approach for precise sampling of restricted regions of conformational space which are important for the calculation of the properties of interest. Many other properties, such as kinetics and transition rates, require sampling of a much larger conformational space. Long time scale simulations will be needed, usually with accelerated methods such as metadynamics. It should be noted that single long time simulations are inaccurate, as we have explained before (e.g. [22]). Ensembles are required for all MD simulations [20,22], as precise predictions, along with their uncertainties, can be obtained only when the most relevant conformations have been extensively sampled. A particular benefit of ESMACS is the freedom to choose multiple trajectory versions to enhance predictions and provide qualitative information about and insight into the associated binding mechanisms. Most publications of MMPBSA studies use 1-trajectory approach in which conformations of the complexes, proteins and ligands are all extracted from single simulations of the complexes. The multiple trajectory versions of the approach require separate simulations for complexes, proteins and ligands, and take into account the flexibility and conformational changes of the proteins and ligands upon binding. Such multiple trajectory versions of ESMACS can significantly improve the predictions compared with those from the 1-trajectory approach when ‘induced fit’ of a ligand is a key feature of the recognition mechanism [18,19]. Relative binding free energies in the alchemical free energy domain have also attracted significant interest, particularly for drug design and drug discovery programmes. Methods for the calculation of these values include, but are not limited to, free energy perturbation (FEP)-based approaches [23,24], which have shown some potential for predicting binding affinities at the accelerated time frames needed for drug discovery, although their accuracy remains inconclusive. We have introduced a method, TIES [25] (thermodynamic integration with enhanced sampling), that makes use of ensemble techniques to ensure reproducibility, accuracy and precision in the calculation of relative binding free energies and to control the errors associated with alchemical predictions; it compares favourably with commercially offered solutions based on FEP [8]. ESMACS can be applied to highly diverse sets of ligands [2,14,16,17,19] whereas TIES is applicable to pairs of ligands of similar chemical structure. Hence ESMACS is suitable for hit-to-lead structure identification in drug discovery, while TIES has a key role in lead optimization [2,8,25]. We emphasize that both TIES and ESMACS have the advantage of being more reproducible and reliable because of the ensemble-based approaches that both protocols use [9,20]. While this certainly increases computational cost, running ensemble simulations in parallel on powerful computers reduces the wall clock time to one relevant to timescales for drug discovery and personalized medicine [18,25,26].

Both ESMACS and TIES have been employed on a variety of globular proteins, including kinase domains of different proteins, HIV proteases, peptide-MHC, bromodomains and so on [3,8,18,19,21,2528]. To test the accuracy of computational predictions of binding affinities on membrane proteins, we have elected to use GPCRs because of their importance to the academic community and to the pharmaceutical industry. GPCRs comprise the single biggest drug target [29] with many unexploited receptors remaining to be used for drug discovery, making the calculation of accurate binding affinities an important means by which to improve the number of drugs that successfully progress from the development pipeline to the clinic. Interestingly, and despite the wealth of published experimental data that exist for this receptor superfamily, there have been relatively few studies that report the computational prediction of the binding affinity of a ligand to its target GPCR [3034], providing us with the opportunity to compare computational calculations of binding affinities using ESMACS and TIES with published experimental results.

Proteins, generally, and GPCRs, in particular, are dynamic and function in a complex energy landscape, possessing different conformational states and interconverting between these in response to the available free energy of the system [35,36]. They can be considered to exist ‘mainly as a group of structures not too different from one another in free energy, but frequently differing considerably in energy and entropy’ [37]. The conformational changes that GPCRs undergo are elicited in response to interactions between the receptor and the ligands that bind specifically to it and to interactions between the receptor and additional proteins involved in the signalling process, including G proteins. Binding affinities and receptor conformations are inextricably intertwined. The advent of high-resolution X-ray crystal structures for GPCRs in active and inactive physiological states [38] has provided an unprecedented opportunity to examine the structural coverage of binding sites and receptor–ligand interactions [39] and affords a means by which to explore the conformational states and sub-states of these receptors, a number of which can be correlated with receptor activity. It is somewhat ironic that the experimental confirmation of the multiple active states predicted from the quantitative mathematical models of GPCRs has been provided by X-ray crystallography, a technique that emphasizes a single static macromolecule. However, the available active state GPCR X-ray crystal structures vary between crystal structures of the same GPCR and between those of different GPCRs in a manner that may be attributed either to the formation of an intermediate state which precedes the existence of a fully active state, or to the detection of one of several active structures of the GPCR [40].

We have explicitly chosen to interrogate the A1 and A2A adenosine receptors for this work as high-resolution X-ray crystal structures of both the active and inactive forms [38] are available, substantial amounts of kinetic binding data exist and these are GPCRs with which we are familiar experimentally [4143]. Our findings are broadly applicable to other GPCRs and to other, different cell surface receptors. The automation of the ESMACS and TIES protocols within our binding affinity calculator (BAC) [44] allows the rapid generation of binding affinities for GPCRs and other receptor protein families of interest.

2. Methods

In this section, we first describe the set-up of the simulations before explaining the two methodologies used to predict binding affinity values.

2.1. Creation of receptor models

The computation of accurate binding affinities depends upon having both an accurate model of the target protein and accurately predicted poses for the ligands. GPCR structures are highly plastic, frequently adopting different conformations depending on the type of ligand to which they are bound. Three different states have been identified experimentally for the A2A receptor depending on its binding partners, inactive (antagonist bound), active (also referred to as partially active, agonist bound) and fully active (in complex with a G-protein) [45,46]. The closely related A1 receptor is believed to explore similar states but as yet structures only exist for the inactive and fully active states [47]. The G-protein complexed (fully active) form of the receptors is extremely large and we excluded it from investigation on the grounds of computational cost. In the absence of a suitable active state structure on which to base our models, we chose to simulate A1 receptor agonists in the inactive state model (figure 1).

Figure 1.

Figure 1.

Structures of the (a) inactive A1 (PDB accession number: 5UEN) and (b) inactive (beige) and active (blue) A2A receptors (PDB accession numbers: 5IU4 and 4UHR, respectively) in cartoon representation.

All available structures of the A1 and A2A adenosine receptors are incomplete; all structures contain unresolved loop regions and incorporate mutations designed to facilitate crystallization. In order to obtain complete and wild-type structures for simulation, we employed the homology modelling functionality of the Molecular Operating Environment (MOE) package. The wild-type sequences of both receptors were taken from the GPCR database (gpcrdb.org) [48]. The following PDB structures were used as templates for the modelling of the different receptor states: 4UHR (CGS21680 bound) [45], 5IU4 (ZM-241,385 bound) [49], for the active and inactive forms of the A2A, and 5UEN [50] (co-crystallized with a ligand with no kinetic binding data) for the inactive form of the A1. The complete models of the two receptors, used in the study, are shown in figure 1.

GPCRs are membrane proteins. To ensure physiologically relevant simulations the models we have generated must be inserted into appropriate ligand membranes. Coordinate models of the membrane bound protein were generated within CHARMM [51] using a temporary CGENFF [52] parametrization for the ligands. A 100% DPPC lipid bilayer was generated around each receptor using the replacement method based on scripts adapted from the CHARMM-GUI membrane builder [53]. Each protein–membrane model was solvated in a tetragonal box containing TIP3P water molecules [54]. The ParmEd tool from AmberTools 16 [55] was then used to convert the systems to use the Amber FF14SB [56] forcefield for the protein and Lipid 14 [57] for the membrane (water remained parametrized using TIP3P). Histidines were assigned standard AMBER protonation states for a pH 7 environment. Final box dimensions for the A1 receptor were 86 × 86 × 138 Å, with 100 and 97 lipid molecules in the top and bottom layers of the membrane, respectively. Box dimensions for the inactive and active A2A receptor models were 76 × 76 × 132 Å and 77 × 77 × 132 Å, respectively, with 76 lipid molecules in the top layer and 77 in the bottom layer of the membrane. All systems were examined to identify any water molecules trapped in the centre of the bilayer which were then removed. Counter ions were added to neutralize the simulation boxes, with 12 Cl and 9 Cl for the A1 and A2A receptors, respectively.

2.2. Molecular dynamics simulations

MD simulations were then performed using the NAMD 2.10 package [58]. Periodic boundary conditions were applied, the Particle Mesh Ewald method [59] was applied for long range electrostatics and a Lennard-Jones cut-off of 12 Å employed. The Langevin thermostat [60] was used with a low damping coefficient of 1 ps−1 to keep the fluctuations between the current temperature and the target temperature to a minimum. Simulations were run at 310.15 K, the human physiological temperature, to mimic the behaviour of these receptors in vivo. Langevin piston control [61] was used with a damping period set to 50 fs and a time-decay period of 20 fs to maintain the pressure at 1 atm. Snapshots were saved every 1 ps for all simulations. Furthermore, all covalent hydrogen bonds were constrained using the SHAKE algorithm [62].

30 000 steps of energy minimization were performed for each system using the default conjugate gradient-coupled line search algorithm. An equilibration protocol (table 1) was followed for a total of 12.5 ns during which the first 3 ns were performed in the NVT (constant number of atoms, volume and temperature) ensemble to let the lipid molecules adjust to the volume space, and the remaining 9.5 ns were performed in the NPT (constant number of atoms, pressure and temperature) ensemble to mimic the biological experimental setting. Velocity rescaling was performed every 500 steps while applying constraints to the backbone, side-chains, and the heavy atoms of membrane lipid heads and tails. Constraints were slowly released towards the end of equilibration as described in table 1. The last 4 ns of equilibration for each system was performed in an NPT ensemble to allow sufficient time for the complexes to relax, adjust and adopt initial stable configurations in the absence of restraints. The final frame of the simulation was used in all docking and subsequent simulations.

Table 1.

Description of the equilibration protocol and the harmonic constraints applied per step in the simulation set-up.

step time step (fs) ensemble equilibration time (ns) harmonic constraints (kcal mol−1 Å−2)
backbone sidechains lipid heads lipid tails ions
1 1 NVT 1 10 5 2.5 2.5 5
2 1 NVT 1 5 2.5 2.5 2.5 0
3 1 NVT 1 2.5 1 1 1 0
4 2 NPT 2 1 0.5 0.5 0.5 0
5 2 NPT 2 0.5 0.1 0.1 0.1 0
6 2 NPT 1.5 0.1 0 0 0 0
7 2 NPT 4 0 0 0 0 0

2.3. Ligand dataset

The ligands in the dataset (table 2 and figure 2) were chosen as they all had binding affinities determined using kinetic radioligand binding assays. The dataset is highly diverse comprising both agonists, antagonists and inverse agonists. In addition, eight of the ligands have experimental binding affinity values determined for both the A1 and A2A receptor, enabling experimentally determined receptor-relative selectivity binding free energies to be calculated.

Table 2.

Table of ligands used in this study including associated experimental binding affinity data. The PDB column contains the PDB accession number of A2A receptor structures from which we extract three-dimensional ligand binding poses.

ligand name abbreviation ligand type PDB experimental binding free energies (kcal mol−1) [6269]
A2A A1
CGS15943 CGS antagonist −12.70 ± 0.06 −12.49 ± 0.10
LUF5834 LUF34 agonist −9.77 ± 0.25 −11.53 ± 0.10
LUF5963 LUF3 antagonist −8.70 ± 0.15 −10.96 ± 0.05
LUF5964 LUF4 antagonist −9.28 ± 0.42 −12.59 ± 0.09
LUF5967 LUF7 antagonist −8.54 ± 0.33 −12.03 ± 0.10
NECA NECA agonist 2YDV −9.52 ± 0.13 −8.69 ± 0.13
theophylline Theo antagonist 5MZJ −7.16 ± 0.09 −7.68 ± 0.11
XAC XAC antagonist 3REY −10.11 ± 0.15 −10.86 ± 0.06
CGS21680 NGI agonist 4UHR −8.14 ± 0.09
LUF5448 LUF8 agonist −8.49 ± 0.15
LUF5549 LUF9 agonist −9.90 ± 0.14
LUF5550 LUF0 agonist −8.84 ± 0.15
LUF5631 LUF1 agonist −9.17 ± 0.20
LUF5833 LUF33 agonist −9.83 ± 0.27
LUF5835 LUF35 agonist −9.85 ± 0.26
UK-432,097 UK agonist 3QAK −10.31 ± 0.07
ZM-241,385 ZMA inverse agonist 5IU4 −11.71 ± 0.09
LUF6057 7 agonist −11.19 ± 0.15
CCPA CCPA agonist −9.59 ± 0.10
CHEMBL3613119 119 agonist −11.64 ± 0.10
CHEMBL3613120 120 agonist −11.17 ± 0.22
DPCPX DPX inverse agonist −12.11 ± 0.07
FSCPX FPX antagonist −11.91 ± 0.14

Figure 2.

Figure 2.

The structures of the ligands used in our study with the shortened names corresponding to table 2.

In order to accurately compare the computational predictions of binding values to the experimental results, the experimentally determined equilibrium dissociation constants, KD, were converted into Gibbs free energies using the equation

ΔG=RTlnKD. 2.1

The binding free energies calculated from the experimentally determined KD for all ligands are shown in table 2. Where the same ligand had its KD value measured in multiple publications, the average value is shown.

2.4. Ligand parametrization and docking

Crystal structures of 6 ligands in our datasets are available in complex with the A2A receptor (table 2). In these cases, we use those conformations in our modelling. For the remaining ligands, the structures were manually produced and optimized in IQMol [71].

All ligands were parametrized using the Antechamber component of AmberTools 16 [72,73]. For ESMACS ligands charges were obtained using the AM1BCC approach and parametrized using the general AMBER force field (GAFF) [73]. Individual ligand topologies employed in the creation of hybrid topologies used in TIES were also parametrized with GAFF but partial charges were derived by invoking the RESP algorithm following geometry optimization in Gaussian09 [74] (employing a Hartree–Fock wavefunction with a 6 − 31 + G* basis set).

For the six ligands bound to A2A for which crystal binding poses were available (table 2), the experimental crystal docking poses were retained by aligning the experimental structure with the appropriate target structure. All other ligands were docked into the binding pocket of their respective receptors using the AutoDock Vina [75] plugin in the UCSF Chimera [76] package. Single binding poses were chosen on the basis of the best docking score obtained, which showed that all of the ligands (including the antagonists) bind to the orthosteric binding site. This agrees with the experimental observation from a high number of co-crystallized structures and site-directed mutagenesis binding data. We would like to point out that in cases where a ligand may target multiple binding sites, or target an allosteric pocket, the metadynamics protocol [5,6] may be useful. In the remainder of this work, we refer to a complete parametrized model containing protein, equilibrated membrane and docked ligand as a ‘starting structure’.

2.5. Binding free energy protocols

Here, we use two computational techniques to gain information about ligand binding strengths: ESMACS, which ranks absolute binding free energies (ΔG) directly, and TIES, which computes differences in Gibbs free energies between two related systems (ΔΔG). The set-up of the simulations for each protocol is substantially different and is described in detail below. Simulation execution and analysis for both protocols were automated via our BAC [44] workflow tool.

2.6. ESMACS

ESMACS protocols are designed to provide converged binding free energies calculated using the MMPBSA methodology from ensembles of relatively short duration simulations for diverse ligand datasets. They include a range of methodologies to compute the entropic contribution to binding usually neglected in standard MMPBSA approaches and may also account for ligand and receptor flexibility using multiple trajectories, including not only that of the complex but those of the unbound ligand and apo protein.

The binding free energy associated with the binding of a ligand to its target protein is calculated as follows:

ΔG=GComplexGreceptorGligand, 2.2

where GComplex, Greceptor and Gligand are the average values of the free energy contribution from complex, receptor (protein) and ligand respectively. Traditionally, in MMPBSA calculations, sampling is conducted using simulations of the complex alone.

The estimate of the component free energy provided by ESMACS can be decomposed as follows:

GESMACSi=GMMPBSAiTSconfi=EMMi+GsolventiTSconf=Einti+EvdWi+Eelei+GPBi+GSAiTSconf,} 2.3

where EMMi is the molecular mechanical energy contribution of the species i, in a complex, free receptor or unbound ligand in the gas phase. Sconf is the configurational entropy. This comprises internal bonded energies (Einti), van der Waals (EvdWi), and electrostatic interactions (Eelei). Gsolventi is the solvent free energy term estimated from the sum of the Poisson–Boltzmann (GPBi) and the non-polar solvation free energy terms (GSAi). GSAi is calculated from the solvent accessibility surface area (SASA) using

GSAi=γ×SASA+b, 2.4

where γ is the surface tension, and b the offset (we use the default values of 0.00542 kcal mol−1 Å−2 and 0.92 kcal mol−1 respectively [77]). The entropic term is introduced as the product of the temperature (T) and configurational entropy. The most common method of computing the configurational entropy is through normal mode analysis [78,79]. However, converging these calculations is computationally demanding for large systems [18] (potentially using as many computational resources as the original molecular dynamics calculations). This motivated the creation of an alternative solution: the weighted solvent accessible surface area (WSAS) model [19,80]. This method was parametrized to reproduce normal mode analysis results from computationally cheap atomistic surface area calculations. In this approach the solvent accessible surface area (SAS) and buried surface area (BSAS) are weighted according to atom type and the sum of the contributions of each atom is used to estimate Sconf as per the following relationship:

SconfWSAS=i=1Nwi(SASikBSASi), 2.5

where wi is the atom-type specific weighting of the atom i and k is a scaling parameter of BSAS. BSAS for atom i is computed using:

BSASi=4π(ri+rprob)2SASi, 2.6

where ri is the radius of the atom i, and rprob the probe radius of a water molecule. Here, we compute the surface areas using the Lee and Richards algorithm [81] as implemented in freesasa [80].

The starting structure generated for each protein–ligand was used to initiate ESMACS runs using a protocol modified from that used for previous work on globular proteins [3,16,18]. In each run 25 replica simulations were executed varying only by initial velocities, which were randomly drawn from the Maxwell–Boltzmann distribution. Each run was initialized with weak harmonic constraints (of up to 3 kcal mol−1 Å−2) applied to the receptors' backbone and ligand's heavy atoms, which were slowly released during 0.5 ns of equilibration. Following this, production simulations were instigated. The same NAMD settings were used in production simulations as for the NPT steps of the membrane equilibration protocol.

Between 30 and 45 ns were required for convergence for models containing docked ligands, with slightly more rapid convergence for the ligands using poses copied from crystal structures (up to 28 ns). This is around 10 times longer than the protocol used in previous ESMACS studies and is due to the complex nature of GPCRs. The binding free energies predicted from ESMACS were based on 50 uniformly distributed frames across the last 10 ns of each of the 25 replicas and then averaged.

2.7. TIES

TIES is based on thermodynamic integration (TI), a well-established example of so-called alchemical binding free energy methods [8285]. Alchemical free energy calculations employ unphysical (alchemical) intermediates to calculate changes in free energies between two physically real systems. It is common in these methods to refer to a variable, λ, which describes the path taken to transform one ligand into another. The parameter varies between 0 and 1, with 0 representing the initial ligand, L1, and 1 the final ligand, L2. The potential between these endpoints is given by

V(λ,x)=(1λ)V1(λ,x)+λV2(λ,x), 2.7

where V1 and V2 are the potential energies of L1 and L2, respectively, and x represents the coordinates of the system. The derivative of the hybrid potential energy with respect to λ, V(λ,x)/λ, is used to compute the free energy difference using

ΔGTI=01V(λ,x)λλdλ, 2.8

where λ denotes the ensemble average at the chosen λ. In practice, the integral is calculated numerically, with MD sampling used in the computation of the ensemble averages at a set of discrete points (the so-called λ-windows). In TIES, multiple replica MD simulations are performed at each λ-window.

We employ a thermodynamic cycle approach to calculate relative free energy difference (ΔΔGTIES) between two ligands:

ΔΔGTIES=ΔG1ΔG2=ΔGTIaqueousΔGTIbound, 2.9

where ΔG1 and ΔG2 are the binding energies for L1 and L2, respectively. ΔGTIaqueous and ΔGTIbound are the free-energy components resulting from the alchemical transformation of L1 to L2 in the unbound and bound states.

As described in previous work [25], treating the integrals in equation (2.9) through the lens of stochastic calculus provides a robust method to estimate uncertainties. The ensemble average of the potential derivative is calculated as the average of its values from all replica simulations in our ensemble simulation, where the individual value for each replica is taken to be the average potential derivative over the whole simulation length. The error associated with each λ-window is computed as a bootstrapped standard error of the mean of the λ derivatives from all sampled replicas using

σx=aqueous,bound2=λσλ2(Δλ)2, 2.10

where σλ2 is the variance associated with the relevant λ-window in the aqueous or bound calculation, as appropriate. This error is the convolution of the individual errors for each λ window. The overall error, σ, then is computed using

σ2=σaqueous2+σbound2. 2.11

The domain of validity of TIES targets resides in determining differences in binding free energies between closely chemically related (for example congeneric) ligands between which there are no charge differences. A list of ligand pairs which meet these criteria in our experimental dataset is provided in table 3.

Table 3.

The ligand pairs (L1 and L2) for which TIES calculations were performed in this study and their associated experimentally determined relative binding affinities (ΔΔG).

transformation
ΔΔGExperiment (kcal mol−1)
L1 L2 A2A A1
LUF3 LUF4 −0.58 ± 0.45 −1.63 ± 0.10
LUF3 LUF7 0.16 ± 0.36 −1.07 ± 0.11
LUF4 LUF7 0.74 ± 0.53 0.56 ± 0.13
Theo XAC −2.95 ± 0.17 −3.18 ± 0.13
LUF8 LUF1 −0.68 ± 0.25
LUF34 LUF35 −0.08 ± 0.36
LUF33 LUF34 0.06 ± 0.37
LUF33 LUF35 −0.02 ± 0.37
NECA CCPA −0.90 ± 0.16
119 120 0.47 ± 0.24
XAC DPX −1.25 ± 0.09

A hybrid ligand topology must be created based on the chemically common region, a disappearing domain comprising the atoms only present in L1 and an appearing domain containing atoms unique to L2. We generated initial common region for each pair of ligands using FESetup [86] and used these to align the two ligands. Atoms were removed from the common region if their charge differed by more than 0.1e. The partial atomic charges for the hybrid ligand were obtained from the RESP derived partial atomic charges on the individual ligands such that the common atoms had identical charges, taken to be the average of their charges in the individual ligands. The charges on disappearing and appearing parts were then adapted by reparametrizing the ligands after constraining the charges on the common atoms to their new values.

We deployed the same basic TIES protocol set out by Bhati et al. [25], using 5 replica simulations in each of 13 non-interacting λ-windows, placed at: 0, 0.05, 0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9, 0.95 and 1.0. We employed a soft-core potential [87,88] for the van der Waals interactions to prevent divergent potential energies, which may arise when there are sudden appearances or disappearances of atoms close to the endpoints of the alchemical transformations. The electrostatic interactions of the disappearing atoms were linearly decoupled from the simulations between λ values of 0 and 0.55 and then turned off, while those of the appearing atoms were linearly coupled from λ values of 0.45 to 1 and then fully activated.

The hybrid ligands were docked into the target structures using the same approaches employed for the single ligands within the ESMACS calculations. Each replica simulation was instigated from the starting structure and varied only by the initially randomized velocities [25]. To ensure accurate and precise values of ΔΔGTIES, a three-step equilibration process lasting 2 ns was performed at each λ window. The integration time steps over the three equilibration periods were 0.5, 1 and 2 fs, respectively. The electrostatic interaction energies were computed at every step to ensure integrator stability. As the TIES protocol only computes the difference in binding free energy between two similar ligands, 4 ns production runs were performed, as done for other systems within the TIES protocol [2,25]. All molecular dynamics simulations were performed using NAMD with the same thermostat and barostat settings as applied in ESMACS and the membrane equilibration protocol.

3. Results

In this section, we present binding affinity predictions from both ESMACS and TIES.

3.1. ESMACS

Predictions of binding free energy (ΔG), using the ESMACS protocol, were carried out on 14 and 17 ligands of the A1 and A2A­ receptors, respectively (structures and experimentally determined data shown in figure 2 and table 2). The results of the predicted ΔG compared to experimentally determined data are shown in figure 3. As there is an approximately 1 kcal mol−1 range of experimentally determined ΔG values available from different publications, it is likely that the associated errors are underreported. For the 14 A1 receptor ligands, the correlation obtained is reasonable (RP = 0.53). The overall correlation for the A2A receptor ligands is slightly weaker (RP = 0.43). As ‘active’ and ‘inactive’ starting structures were used to calculate ΔG values for agonists and antagonists (and inverse agonists), respectively, use of two separate correlation lines is more appropriate for the A2A receptor. This results in stronger correlations of RP = 0.73 for the antagonists (and the inverse agonist) and RP = 0.55 for the agonists. If one includes the configurational entropy in the ESMACS ΔG binding free energies predictions, the A1 and A2A correlations become weaker, with the exception of the A2A agonists which retains the same strength in correlation (RP = 0.55). Overall ESMACS performs similarly well for structurally diverse A1 ligands, A2A agonists and A2A antagonists and better without inclusion of the configurational entropy.

Figure 3.

Figure 3.

Summary of results obtained using ESMACS for ligands of the (a) A1 and (b) A2A receptors. The grey dashed line in (a) is the linear correlation line for all the ligands. For the A2A receptor (b), agonists and antagonists are coloured blue and red, respectively. The two dashed lines in (b) show the linear correlation for agonists and antagonists, also coloured blue and red, respectively.

3.2. TIES

Relative binding free energies (ΔΔG) were calculated using TIES for 7 and 8 pairs of ligands for the A1 and A2A receptor, respectively (table 3). These 15 ligand pairs were the only ligands structurally similar enough in the data (table 2) that we were able to calculate hybrid topologies. To our knowledge, this is the first published use of an alchemical binding free energy prediction method on ligands of the A1 receptor. The results of these relative binding free energy calculations against experimentally determined data are plotted in figure 4. As TIES aims to calculate relative binding free energies, the correlation line was set as y = x. The overall mean absolute error was calculated as 1.2 and 0.98 kcal mol−1 for all ligands of the A1 and A2A receptors, respectively. This is similar to values reported previously using TIES on much simpler, non-membrane proteins [25]. All but one of the predicted ΔΔG values are directionally correct.

Figure 4.

Figure 4.

Plot of computed relative binding free energies using TIES (TIES ΔΔG) against experimentally determined relative binding free energies for ligand pairs of the (a) A1 and (b) A2A receptor. All reported ligand transformations are listed in table 3. Values were classified as outliers when their Cook's distance [89] was larger than 4/n, where n represents the total number of points used in the regression. These outliers are labelled and plotted in orange. The dotted line plots y = x, representing correct prediction values.

Furthermore, only one transformation pair's predicted ΔΔG values, from each receptor subtype, was classified as an outlier from its true value when its Cook's distance [89] was greater than 4/n (n being the number of points used in the regression line). These pairs are highlighted in orange in figure 4. One of the outliers, the Theo -> XAC ligand pair in the A1 receptor, is the largest alchemical transformation performed in our TIES calculations, as these ligands are the most structurally divergent among all the ligand pairs and therefore the least reliable calculation. Excluding these outliers, the mean absolute error improves to 0.98 and 0.66 kcal mol−1 for remaining A1 and A2A ligand pairs, respectively. This is similar to the weighted mean absolute errors achieved using alchemical free energy calculations on two ligand series of the A2A receptor [31].

4. Conclusion

Using the TIES and ESMACS protocols, we have computed the binding free energies of a series of ligands at the A1 and A2A adenosine receptors, two GPCRs for which substantial quantities of structural and functional data exist. Our rankings for binding free energies determined by both ESMACS and TIES are in line with previous experiments, confirming our ability to use these protocols on GPCRs, which are much larger protein targets than used previously with these protocols. ESMACS predicts values that correlate well with experimentally determined binding free energy values for structurally diverse sets of ligands, confirming its value for the hit-to-lead phase of drug discovery. TIES is again found to be a powerful protocol for the accurate calculation of relative binding free energies between structurally similar ligands, and is superior to methods that do not use equivalent ensemble-based sampling techniques, such as FEP+ [8]. TIES is thus of considerable value in lead optimization; here, we demonstrated this in a relatively extreme case involving a complex protein class with large alchemical transformations and correspondingly small common substructures. Although drug design has numerous constraints, we conclude that this methodology provides a useful tool with which to inform structure-based drug discovery workflows for the development of novel GPCR therapeutics.

Acknowledgements

We thank the University College London High Performance Computing infrastructure for providing access to their GRACE and Legion platforms. We acknowledge the Leibniz Supercomputing Centre (LRZ) for providing access to SuperMUC, the Poznan Super Computing and Network Centre (PSNC) for providing access to Eagle, Cyfronet in Kraków for providing access to Prometheus, and the US National Centre for Supercomputing Applications for providing access to BlueWaters (NSF grant no. 1713749). These computers were all used to produce different parts of the computational work presented.

Data accessibility

This article has no additional data.

Authors' contributions

All authors participated in the design of the study and contributed to the writing of the paper. S.W., A.P. and F.H. carried out the simulation work and performed analysis. M.M. provided assistance with simulation work performed on Prometheus. A.T.N. and P.V.C. conceived and directed the project. All authors gave final approval for publication and agree to be held accountable for the work performed herein.

Competing interests

We declare we have no competing interests.

Funding

We thank the European Commision for Horizon 2020 funding for the Computational Biomedicine Centre of Excellence (grant nos. 675451 and 823712) and VECMA (grant no. 800925), the MRC Medical Bioinformatics project (MR/L016311/1), and the UCL Provost for special funding to P.V.C. A.P. was supported by the Biotechnology and Biological Sciences Research Council (grant number BB/M009513/1) and the London Interdisciplinary Bioscience PhD Consortium (LIDo).

References

  • 1.Altwaijry NA, Baron M, Wright DW, Coveney PV, Townsend-Nicholson A. 2017. An ensemble-based protocol for the computational prediction of helix-helix interactions in g protein-coupled receptors using coarse-grained molecular dynamics. J. Chem. Theory Comput. 13, 2254–2270. ( 10.1021/acs.jctc.6b01246) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Wan S, Bhati AP, Zasada SJ, Wall I, Green D. 2017. Rapid and reliable binding affinity prediction for analysis of Bromodomain inhibitors : a computational study. J. Chem. Theory Comput. 13, 784–795. ( 10.1021/acs.jctc.6b00794) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Wright DW, Hall BA, Kenway OA, Jha S, Coveney PV. 2014. Computing clinically relevant binding free energies of HIV-1 protease inhibitors. J. Chem. Theory Comput. 10, 1228–1241. ( 10.1021/ct4007037) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Potterton A, Husseini FS, Southey MWY, Bodkin MJ, Heifetz A, Coveney PV, Townsend-Nicholson A. 2019. Ensemble-based steered molecular dynamics predicts relative residence time of A2A receptor binders. J. Chem. Theory Comput. 15, 3316–3330. ( 10.1021/acs.jctc.8b01270) [DOI] [PubMed] [Google Scholar]
  • 5.Saleh N, Ibrahim P, Saladino G, Gervasio FL, Clark T. 2017. An efficient metadynamics-based protocol to model the binding affinity and the transition state ensemble of G-protein-coupled receptor ligands. J. Chem. Inf. Model. 57, 1210–1217. ( 10.1021/acs.jcim.6b00772) [DOI] [PubMed] [Google Scholar]
  • 6.Saleh N, et al. 2018. Multiple binding sites contribute to the mechanism of mixed agonistic and positive allosteric modulators of the cannabinoid CB1 receptor. Angew. Chem. Int. Ed. 57, 2580–2585. ( 10.1002/anie.201708764) [DOI] [PubMed] [Google Scholar]
  • 7.Sherborne B, et al. 2016. Collaborating to improve the use of free-energy and other quantitative methods in drug discovery. J. Comput. Aided Mol. Des. 30, 1139–1141. ( 10.1007/s10822-016-9996-y) [DOI] [PubMed] [Google Scholar]
  • 8.Wan S, Tresadern G, Pérez-Benito L, Vlijmen H, Coveney PV. 2020. Accuracy and precision of alchemical relative free energy predictions with and without replica-exchange. Adv. Theory Simul. 3, 1900195 ( 10.1002/adts.201900195) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Wan S, Bhati AP, Zasada SJ, Coveney PV. 2020. Rapid, accurate, precise and reproducible ligand-protein binding free energy prediction. Interface Focus 10, 20200007 ( 10.1098/rsfs.2020.0007) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Kollman PA, et al. 2000. Calculating structures and free energies of complex molecules: combining molecular mechanics and continuum models. Acc. Chem. Res. 33, 889–897. ( 10.1021/ar000033j) [DOI] [PubMed] [Google Scholar]
  • 11.Abel R, Young T, Farid R, Berne BJ, Friesner RA. 2008. Role of the active-site solvent in the thermodynamics of factor Xa ligand binding. J. Am. Chem. Soc. 130, 2817–2831. ( 10.1021/ja0771033) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Hansson T, Åqvist J. 1995. Estimation of binding free energies for HIV proprotein inhibitors by molecular dynamics simulations. Protein Eng. Des. 8, 1137 ( 10.1093/protein/8.11.1137) [DOI] [PubMed] [Google Scholar]
  • 13.Åqvist J. 1996. Calculation of absolute binding free energies for charged ligands and effect of long-range electrostatic interactions. J. Comp. Chem. 17, 1587 () [DOI] [Google Scholar]
  • 14.Elofsson L, Nilsson L. 1993. How consistent are molecular dynamics simulations? Comparing structure and dynamics in reduced and oxidized Escherichia coli thioredoxin. J. Mol. Biol. 233, 766–780. ( 10.1006/jmbi.1993.1551) [DOI] [PubMed] [Google Scholar]
  • 15.Mongan J, Case DA, McCammon JA. 2004. Constant pH molecular dynamics in generalized Born implicit solvent. J. Comput. Chem. 25, 2038–2048. ( 10.1002/jcc.20139) [DOI] [PubMed] [Google Scholar]
  • 16.Sadiq SK, Wright DW, Kenway OA, Coveney PV. 2010. Accurate ensemble molecular dynamics binding free energy of multidrug-resistance HIV-1 protease. J. Chem. Inf. Model. 50, 890–905. ( 10.1021/ci100007w) [DOI] [PubMed] [Google Scholar]
  • 17.Wan S, Coveney PV. 2011. Rapid and accurate ranking of binding affinities of epidermal growth factor receptor sequences with selected lung cancer drugs. J. R. Soc. Interface 8, 1114–1127. ( 10.1098/rsif.2010.0609) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Wan S, Knapp B, Wright DW, Deane CM, Coveney PV. 2015. Rapid, precise, and reproducible prediction of peptide – MHC binding affiinities from molecular dynamics that correlate well with experiment. J. Chem. Theory Comput. 11, 3346–3356. ( 10.1021/acs.jctc.5b00179) [DOI] [PubMed] [Google Scholar]
  • 19.Wan S, Bhati AP, Skerratt S, Omoto K, Shanmugasundaram V, Bagal SK, Coveney PV. 2017. Evaluation and characterization of Trk kinase inhibitors for the treatment of pain: reliable binding affinity predictions from theory and computation. J. Chem. Inf. Model. 57, 897–909. ( 10.1021/acs.jcim.6b00780) [DOI] [PubMed] [Google Scholar]
  • 20.Coveney PV, Wan S. 2016. On the calculation of equilibrium thermodynamic properties from molecular dynamics. Phys. Chem. Chem. Phys. 18, 30 236–30 240. ( 10.1039/C6CP02349E) [DOI] [PubMed] [Google Scholar]
  • 21.Bhati AP, Wan S, Hu Y, Sherborne B, Coveney PV. 2018. Uncertainty quantification in alchemical free energy methods. J. Chem. Theory Comput. 14, 2867–2880. ( 10.1021/acs.jctc.7b01143) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Wan S, Sinclair RC, Coveney PV.2020. Uncertainty quantification in classical molecular dynamics. (https://arxiv.org/abs/2006.07104. )
  • 23.Zwanzig RW. 1954. High-temperature equation of state by a perturbation method in non polar gases. J. Chem. Phys. 22, 1420 ( 10.1063/1.1740409) [DOI] [Google Scholar]
  • 24.Wang L, Friesner RA, Berne BJ. 2011. Replica exchange with solute scaling: a more efficient version of replica exchange with solute tempering (REST2). J. Phys. Chem. B. 115, 9431–9438. ( 10.1021/jp204407d) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Bhati AP, Wan S, Wright DW, Coveney PV. 2017. Rapid, accurate, precise and reliable relative free energy prediction using ensemble based thermodynamic integration. J. Chem. Theory Comput. 13, 210–222. ( 10.1021/acs.jctc.6b00979) [DOI] [PubMed] [Google Scholar]
  • 26.Bhati AP, Wan S, Coveney PV. 2019. Ensemble-based replica exchange alchemical free energy methods: the effect of protein mutations on inhibitor binding. J. Chem. Theory Comput. 15, 1265–1277. ( 10.1021/acs.jctc.8b01118) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Wright DW, Husseini F, Wan S, Meyer C, van Vlijmen H, Tresadern G, Coveney PV. 2020. Application of the ESMACS binding free energy protocol to a multi-binding site lactate dehydogenase A ligand dataset. Adv Theory Simul. 3, 1900194 ( 10.1002/adts.201900194) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Wright DW, Wan S, Meyer C, van Vlijmen H, Tresadern G, Coveney PV. 2019. Application of ESMACS binding free energy protocols to diverse datasets: Bromodomain-containing protein 4. Sci. Rep. 9, 6017 ( 10.1038/s41598-019-41758-1) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Sriram K, Insel PA. 2018. G protein-coupled receptors as targets for approved drugs: how many targets and how many drugs? Mol. Pharmacol. 93, 251–258. ( 10.1124/mol.117.111062) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Cappel D, Hall ML, Lenselink EB, Beuming T, Qi J, Bradner J, Sherman W. 2016. Relative binding free energy calculations applied to protein homology models. J. Chem. Inf. Model. 56, 2388–2400. ( 10.1021/acs.jcim.6b00362) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Lenselink EB, et al. 2016. Predicting binding affinities for GPCR ligands using free-energy perturbation. ACS Omega 1, 293–304. ( 10.1021/acsomega.6b00086) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Mobley DL, Klimovich PV. 2012. Perspective: alchemical free energy calculations for drug discovery. J. Chem. Phys. 137, 230901 ( 10.1063/1.4769292) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Boukharta L, Gutierrez-de-Teran H, Åqvist J. 2014. Computational prediction of alanine scanning and ligand binding energetics in G-protein coupled receptors. PLoS Comput. Biol. 10, e1003585 ( 10.1371/journal.pcbi.1003585) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Lee HS, Seok C, Im W. 2015. Potential application of alchemical free energy simulations to discriminate GPCR ligand efficacy. J. Chem. Theory Comput. 11, 1255–1266. ( 10.1021/ct5008907) [DOI] [PubMed] [Google Scholar]
  • 35.Frauenfelder H, Parak F, Young RD. 1988. Conformational substates in proteins. Annu. Rev. Biophys. Biophys. Chem. 17, 451–479. ( 10.1146/annurev.bb.17.060188.002315) [DOI] [PubMed] [Google Scholar]
  • 36.Frauenfelder H, Sligar SG, Wolynes PG. 1991. The energy landscapes and motions of proteins. Science 254, 1598–1603. ( 10.1126/science.1749933) [DOI] [PubMed] [Google Scholar]
  • 37.Linderstrom-Lang KU, Shellmann J. 1959. The enzymes. New York, NY: Academic Press. [Google Scholar]
  • 38.Robertson N, Jazayeri A, Errey J, Baig A, Hurrell E, Zhukov A, Langmead CJ, Weir M, Marshall FH. 2011. The properties of thermostabilised G protein-coupled receptors (StaRs) and their use in drug discovery. Neuropharmacology 60, 36–44. ( 10.1016/j.neuropharm.2010.07.001) [DOI] [PubMed] [Google Scholar]
  • 39.Vass M, Kooistra AJ, Yang D, Stevens RC, Wang M-W, Graaf C. 2018. Chemical diversity in the G protein-coupled receptor superfamily. Trends Pharmacol. Sci. 39, 494–510. ( 10.1016/j.tips.2018.02.004) [DOI] [PubMed] [Google Scholar]
  • 40.Park PS. 2012. Ensemble of G protein-coupled receptor active states. Curr. Med. Chem. 19, 1146–1154. ( 10.2174/092986712799320619) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Townsend-Nicholson A, Shine J. 1992. Molecular cloning and characterisation of a human brain A1 adenosine receptor cDNA. Brain Res. Mol. Brain Res. 16, 365–370. ( 10.1016/0169-328X(92)90248-A) [DOI] [PubMed] [Google Scholar]
  • 42.Townsend-Nicholson A, Schofield PR. 1994. A threonine residue in the seventh transmembrane domain of the human A1 adenosine receptor mediates specific agonist binding. J. Biol. Chem. 269, 2373–2376. [PubMed] [Google Scholar]
  • 43.Townsend-Nicholson A, Baker E, Schofield PR, Sutherland GR. 1995. Localization of the adenosine A1 receptor subtype gene (ADORA1) to chromosome 1q32.1. Genomics. 26, 423–425. ( 10.1016/0888-7543(95)80236-F) [DOI] [PubMed] [Google Scholar]
  • 44.Sadiq S, Wright DW, Watson SJ, Zasada S, Stoica I, Coveney PV. 2008. Automated molecular simulation based binding affinity calculator for ligand-bound HIV-1 protease. J. Chem. Inf. Model. 48, 1909 ( 10.1021/ci8000937) [DOI] [PubMed] [Google Scholar]
  • 45.Lebon G, Edwards PC, Leslie AGW, Tate CG. 2015. Molecular determinants of CGS21680 binding to the human adenosine A2A receptor. Mol. Pharmacol. 87, 907–915. ( 10.1124/mol.114.097360) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Carpenter B, Nehmé R, Warne T, Leslie AGW, Tate CG. 2016. Structure of the adenosine A2A receptor bound to an engineered G protein. Nature 536, 104–107. ( 10.1038/nature18966) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Draper-Joyce CJ, et al. 2018. Structure of the adenosine-bound human adenosine A1 receptor–Gi complex. Nature 558, 559–563. ( 10.1038/s41586-018-0236-6) [DOI] [PubMed] [Google Scholar]
  • 48.Munk C, et al. 2019. An online resource for GPCR structure determination and analysis. Nat. Methods 16, 151–162. ( 10.1038/s41592-018-0302-x) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Segala E, et al. 2016. Controlling the dissociation of ligands from the adenosine A2A receptor through modulation of salt bridge strength. J. Med. Chem. 59, 6470–6479. ( 10.1021/acs.jmedchem.6b00653) [DOI] [PubMed] [Google Scholar]
  • 50.Glukhova A, Thal DM, Nguyen AT, Vecchio EA, Jörg M, Scammells PJ, May LT, Sexton PM, Christopoulos A. 2017. Structure of the adenosine A1 receptor reveals the basis for subtype selectivity. Cell 168, 867–877.e13. ( 10.1016/j.cell.2017.01.042) [DOI] [PubMed] [Google Scholar]
  • 51.Brooks BR, et al. 2009. CHARMM: the biomolecular simulation program. J. Comp. Chem. 30, 1545–1614. ( 10.1002/jcc.21287) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Vanommeslaeghe K, et al. 2010. CHARMM general force field (CGenFF): a force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 31, 671–690. ( 10.1002/jcc.21367) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Jo S, Lim JB, Klauda JB, Im W. 2009. CHARMM-GUI membrane builder for mixed bilayers and its application to yeast membranes. Biophys. J. 97, 50–58. ( 10.1016/j.bpj.2009.04.013) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein ML. 1983. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935. ( 10.1063/1.445869) [DOI] [Google Scholar]
  • 55.Salomon-Ferrer R, Case DA, Walker RC. 2013. An overview of the Amber biomolecular simulation package. Wiley Interdiscip. Rev. Comput. Mol. Sci. 3, 198–210. ( 10.1002/wcms.1121) [DOI] [Google Scholar]
  • 56.Maier J, Martinez C, Kasavajhala J, Wickstrom L, Hauser K, Simmerling C. 2015. ff14SB: improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theory Comput. 11, 3696–3713. ( 10.1021/acs.jctc.5b00255) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Dickson CJ, Madej BD, Skjevik ÅA, Betz RM, Teigen K, Gould IR, Walker RC. 2014. Lipid14: the Amber lipid force field. J. Chem. Theory Comput. 10, 865–879. ( 10.1021/ct4010307) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Phillips JC, et al. 2005. Scalable molecular dynamics with NAMD. J. Comp. Chem. 26, 1781–1802. ( 10.1002/jcc.20289) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Darden T, York D, Pedersen L. 1993. The effect of long-range electrostatic interactions in simulations of macromolecular crystals: a comparison of the Ewald and truncated list methods. J. Chem. Phys. 99, 8345–8348. ( 10.1063/1.465608) [DOI] [Google Scholar]
  • 60.Davidchack RL, Handel R, Tretyakov MV. 2009. Langevin thermostat for rigid body dynamics. J. Chem. Phys. 130, 234101 ( 10.1063/1.3149788) [DOI] [PubMed] [Google Scholar]
  • 61.Feller SE, Zhang YH, Pastor RW, Brooks BR. 1995. Constant pressure molecular dynamics simulation: the Langevin piston method. J. Chem. Phys. 103, 4613–4621. ( 10.1063/1.470648) [DOI] [Google Scholar]
  • 62.Ryckaert JP, Ciccotti G, Berendsen HJC. 1977. Numerical integration of the Cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. J. Comp. Phys. 23, 327–341. ( 10.1016/0021-9991(77)90098-5) [DOI] [Google Scholar]
  • 63.Guo D, Xia L, Van Veldhoven JPD, Hazeu M, Mocking T, Brussee J, Ijzerman AP, Heitman LH. 2014. Binding kinetics of ZM241385 derivatives at the human adenosine A2A receptor. ChemMedChem. 9, 752–761. ( 10.1002/cmdc.201300474) [DOI] [PubMed] [Google Scholar]
  • 64.Guo D, Van Dorp EJH, Mulder-Krieger T, Van Veldhoven JPD, Brussee J, IJzerman AP, Heitman LH. 2013. Dual-point competition association assay: a fast and high-throughput kinetic screening method for assessing ligand-receptor binding kinetics. J. Biomol. Screen. 18, 309–320. ( 10.1177/1087057112464776) [DOI] [PubMed] [Google Scholar]
  • 65.Guo D, Mulder-Krieger T, IJzerman AP, Heitman LH. 2012. Functional efficacy of adenosine A2A receptor agonists is positively correlated to their receptor residence time. Br. J. Pharmacol. 166, 1846–1859. ( 10.1111/j.1476-5381.2012.01897.x) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Guo D, Dijksteel GS, Van Duijl T, Heezen M, Heitman LH, Ijzerman AP. 2016. Equilibrium and kinetic selectivity profiling on the human adenosine receptors. Biochem. Pharmacol. 105, 34–41. ( 10.1016/j.bcp.2016.02.018) [DOI] [PubMed] [Google Scholar]
  • 67.Xia L, de Vries H, IJzerman AP, Heitman LH. 2016. Scintillation proximity assay (SPA) as a new approach to determine a ligand's kinetic profile. A case in point for the adenosine A1 receptor. Purinergic Signal. 12, 115–126. ( 10.1007/s11302-015-9485-0) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Louvel J, Guo D, Soethoudt M, Mocking TAM, Lenselink EB, Mulder-Krieger T, Heitman LH, Ijzerman AP. 2015. Structure-kinetics relationships of Capadenoson derivatives as adenosine A1 receptor agonists. Eur. J. Med. Chem. 101, 681–691. ( 10.1016/j.ejmech.2015.07.023) [DOI] [PubMed] [Google Scholar]
  • 69.McNeely PM, Naranjo AN, Forsten-Williams K, Robinson AS. 2017. A2AR binding kinetics in the ligand depletion regime. SLAS Discov. 22, 166–175. ( 10.1177/1087057116667256) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Guo D, Venhorst SN, Massink A, Van Veldhoven JPD, Vauquelin G, Ijzerman AP, Heitman LH. 2014. Molecular mechanism of allosteric modulation at GPCRs: insight from a binding kinetics study at the human A1 adenosine receptor. Br. J. Pharmacol. 171, 5295–5312. ( 10.1111/bph.12836) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Shao Y, et al. 2006. Advances in methods and algorithms in a modern quantum chemistry program package. Phys. Chem. Chem. Phys. 8, 3172–3191. ( 10.1039/B517914A) [DOI] [PubMed] [Google Scholar]
  • 72.Wang J, Wang W, Kollman PA, Case DA. 2006. Automatic atom type and bond type perception in molecular mechanical calculations. J. Mol. Graph. Model. 25, 247260 ( 10.1016/j.jmgm.2005.12.005) [DOI] [PubMed] [Google Scholar]
  • 73.Wang JM, Wolf RM, Caldwell JW, Kollman PA, Case DA. 2004. Development and testing of a general amber force field. J. Comput. Chem. 25, 1157–1174. ( 10.1002/jcc.20035) [DOI] [PubMed] [Google Scholar]
  • 74.Frisch M, et al. 2016. Gaussian09, revision C.01. Wallingford, CT: Gaussian, Inc. [Google Scholar]
  • 75.Morris GM, Huey R, Lindstrom W, Sanner MF, Belew RK, Goodsell DS, Olson AJ. 2009. AutoDock4 and AutoDockTools4: automated docking with selective receptor flexibility. J. Comput. Chem. 30, 2785–2791. ( 10.1002/jcc.21256) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Pettersen E, Goddard T, Huang C, Couch G, Greenblatt D, Meng E, Ferrin TE. 2004. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612. ( 10.1002/jcc.20084) [DOI] [PubMed] [Google Scholar]
  • 77.Massova L, Kollman PA. 1999. Computational alanine scanning to probe protein-protein interactions: a novel approach to evaluate binding free energies. J. Am. Chem. Soc. 121, 8133–8143. ( 10.1021/ja990935j) [DOI] [Google Scholar]
  • 78.Case DA. 1994. Normal mode analysis of protein dynamics. Curr. Opin. Struct. Biol. 4, 285–290. ( 10.1016/S0959-440X(94)90321-2) [DOI] [Google Scholar]
  • 79.Tidor B, Karplus M. 1993. The contribution of cross-links to protein stability: a normal mode analysis of the configurational entropy of the native state. Proteins 1, 71–79. ( 10.1002/prot.340150109) [DOI] [PubMed] [Google Scholar]
  • 80.Mitternacht S. 2016. FreeSASA: an open source C library for solvent accessible surface area calculations. F1000Research 5, 189 ( 10.12688/f1000research.7931.1) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Lee B, Richards FM. 1971. The interpretation of protein structures: estimation of static accessibility. J. Mol. Biol. 55, 379–400. ( 10.1016/0022-2836(71)90324-X) [DOI] [PubMed] [Google Scholar]
  • 82.Straatsma TP, Berendsen HJC, Postma JPM. 1986. Free energy of hydrophobic hydration: a molecular dynamics study of nobel gases in water. J. Chem. Phys. 85, 6720–6727. ( 10.1063/1.451846) [DOI] [Google Scholar]
  • 83.Straatsma TP, Berendsen HJC. 1988. Free energy of ionic hydration: analysis of a thermodynamic integration technique to evaluate free energy differences by molecular dynamics simulation. J. Chem. Phys. 89, 5876–5886. ( 10.1063/1.455539) [DOI] [Google Scholar]
  • 84.Straatsma TP, McCammon JA. 1991. Multiconfigurational thermodynamic integration. J. Chem. Phys. 95, 1175 ( 10.1063/1.461148) [DOI] [Google Scholar]
  • 85.Straatsma TP, McCammon JA. 1992. Computational alchemy. Annu. Rev. Phys. Chem. 43, 407–435. ( 10.1146/annurev.pc.43.100192.002203) [DOI] [Google Scholar]
  • 86.Loeffler HH, Michel J, Woods C. 2015. FESetup: automating setup for alchemical free energy simulations. J. Chem. Inf. Model. 55, 2485–2490. ( 10.1021/acs.jcim.5b00368) [DOI] [PubMed] [Google Scholar]
  • 87.Zacharias M, Straatsma TP, McCammon JA. 1994. Separation-shifted scaling, a new scaling method for Lennard-Jones interactions in thermodynamic integration. J. Chem. Phys. 100, 9025–9031. ( 10.1063/1.466707) [DOI] [Google Scholar]
  • 88.Beutler TC, Mark AE, van Schaik RC, Gerber PR, van Gunsteren WF. 1994. Avoiding singularities and numerical instabilities in free energy calculations based on molecular simulations. Chem. Phys. Lett. 222, 529–539. ( 10.1016/0009-2614(94)00397-1) [DOI] [Google Scholar]
  • 89.Cook RD. 1975. Detection of influential observation in linear regression. Technometrics 19, 15–18. ( 10.1080/00401706.1977.10489493) [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This article has no additional data.


Articles from Interface Focus are provided here courtesy of The Royal Society

RESOURCES