Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2025 Jan 22.
Published in final edited form as: J Chem Inf Model. 2023 Dec 5;64(2):378–392. doi: 10.1021/acs.jcim.3c01513

Polarizable AMOEBA Model for Simulating Mg2+•Protein•Nucleotide Complexes

Julian M Delgado 1, Péter R Nagy 2, Sameer Varma 3
PMCID: PMC11345861  NIHMSID: NIHMS2014261  PMID: 38051630

Abstract

Molecular mechanics (MM) simulations have the potential to provide detailed insights into the mechanisms of enzymes that utilize nucleotides as cofactors. In most cases, the activities of these enzymes also require the binding of divalent cations to catalytic sites. However, modeling divalent cations in MM simulations has been challenging. The inclusion of explicit polarization was considered promising, but despite improvements over nonpolarizable force fields and despite the inclusion of “Nonbonded-fix (NB-fix)” corrections, errors in interaction energies of divalent cations with proteins remain large. Importantly, the application of these models fails to reproduce the experimental structural data on Mg2+·Protein·ATP complexes. Focusing on these complexes, here we provide a systematic assessment of the polarizable AMOEBA model and recommend critical changes that substantially improve its predictive performance. Our key results are as follows. We first show that our recent revision of the AMOEBA protein model (AMOEBABIO18-HFC), which contains high field corrections (HFCs) to induced dipoles, dramatically improves Mg2+–protein interaction energies, reducing the mean absolute error (MAE) from 17 to 10 kcal/mol. This further supports the general applicability of AMOEBABIO18-HFC. The inclusion of many-body NB-fix corrections further reduces MAE to 6 kcal/mol, which amounts to less than 2% error. The errors are estimated with respect to vdW-inclusive density functional theory that we benchmark against CCSD(T) calculations and experiments. We also present a new model of ATP with revised polarization parameters to better capture its high field response, as well as new vdW and dihedral parameters. The ATP model accurately predicts experimental Mg2+-ATP binding free energy in the aqueous phase and provides new insights into how Mg2+ associates with ATP. Finally, we show that molecular dynamics (MD) simulations of Mg2+·Kinase·ATP complexes carried out with these improvements lead to a better agreement in global and local catalytic site structures between MD and X-ray crystallography.

Graphical Abstract

graphic file with name nihms-2014261-f0011.jpg

INTRODUCTION

Mg2+ is the most abundant divalent cation in the intracellular fluid.13 As a hard metal, it interacts directly with water, carboxylates, phosphates, hydroxyls, and carbonyls present in proteins, nucleotides, nucleic acids, and lipids.47 It is shown to have metabolic, structural, and regulatory functions in many diverse physiological processes.1,3,4,6,810 For example, it modulates the membrane structure, binds to dissolved nucleotides, such as ATP, stabilizes DNA and RNA structures, and serves as a metal cofactor in various enzymes, including kinases, adenylyl cyclases, and helicases.

Molecular mechanics (MM) simulations can, in principle, provide detailed insights into such Mg2+-dependent processes. Here, we focus on MM simulations of Mg2+·Protein·Nucleotide complexes. To simulate such complexes reliably, the employed MM model must accurately describe a wide range of interactions, including Mg2+–water, Mg2+–protein, Mg2+–nucleotide, protein–water, nucleotide–water, and protein–nucleotide interactions.

MM models to describe protein–water interactions have been constructed carefully using reference data from gas- and condensed-phase experiments as well as high-level quantum mechanics (QM) calculations and are refined regularly to improve accuracy.1119 However, Mg2+ ions are not included during the parametrization of protein MM models. Mg2+ parameters are determined separately from reference data on Mg2+–water interactions,2024 where they have been shown to perform reasonably well. In simulations consisting of both proteins and Mg2+ ions, Mg2+–protein interactions are estimated by using a predefined set of mixing rules for Lennard-Jones (LJ) terms. This, however, does not guarantee the reliability of predicted Mg2+–protein interactions. In fact, previous studies,2536 including ours, show that the use of mixing rules can produce large errors in interactions of monovalent cations with proteins. Not surprisingly, for divalents like Ca2+, these errors are even larger31 and can be > 50 kcal/mol. Similar errors can be expected for Mg2+-protein interactions.

In recent years, errors in ion–protein interactions have been reduced to some extent in an a posteriori manner through the so-called “Nonbonded-fix (NB-fix)”-type approaches. In NB-fix, the parameters for ion–protein interactions obtained from mixing rules are substituted with those obtained from fitting to reference data on interactions of ions with small molecules representative of protein chemical groups.3033,3741 However, even after including NB-fix corrections, and in MM models whose multipoles can adjust to a local electric field (polarizable MM models), errors in interactions of proteins with monovalent and divalent cations can exceed 10 and 20 kcal/mol, respectively.31,36 This implies that to further improve accuracies, we need to look beyond NB-fix corrections.

Assessing the accuracy of Mg2+–nucleotide interactions poses an additional challenge related to the limited experimental data. The clearest evidence from experiments is that Mg2+ binds strongly to nucleotides in solution. Compilation of data from 34 different experiments,42 including NMR, calorimetry, spectrophotometry, and pH titration, which we summarize in Table S7 in the Supporting Information, shows that in the aqueous phase, Mg2+-ATP binding free energy lies between −7.2 and −9.9 kcal/mol with an average of −8.5 ± 0.7 kcal/mol. Not surprisingly, most enzymes that utilize nucleotides as cofactors or substrates are known to interact not just with the nucleotides but also with bound divalent cations, especially Mg2+ ions, that coordinate with both nucleotides and proteins.43

While the Mg2+-nucleotide binding free energy is known from experiments, the binding modes of Mg2+ to nucleotides remain ambiguous. Early NMR experiments in the aqueous phase suggest that Mg2+ coordinates with oxygens from only one of the three ATP phosphates.44,45 More recent 31P NMR experiments and 18O vibrational studies suggest that the predominant binding mode is bidentate; that is, Mg2+ coordinates with oxygens from two phosphates.46,47 Finally, reinterpretation of the original 31P data and newer 31P NMR experiments suggest that Mg2+-ATP coordination is predominantly tridentate.4749 In relation to these experiments, QM geometry optimizations conducted in implicit solvents suggest that the most stable binding mode is one in which oxygens from all three phosphates coordinate with Mg2+, followed by a bidentate binding mode.50

This ambiguity is further complicated by experimental structural data for proteins complexed with Mg2+ and nucleotides. Previous statistical analysis of structures from the protein data bank (PDB),51,52 as well as our own analysis (Figure S1 in the Supporting Information), reveals three predominant Mg2+-ATP binding modes, two modes in which Mg2+ coordinates with oxygens of two phosphates (βγ or αβ) and the third in which Mg2+ coordinates with oxygens of all the three phosphates (αβγ). None of the nonpolarizable MM models, including CHARMM and AMBER, have been shown to reproduce this structural data.52 The polarizable AMOEBA model does reproduce Mg2+-ATP binding free energy in the aqueous phase,53 but as shown in the Results and Discussion section, when it is employed to conduct molecular dynamics (MD) simulations of Mg2+·Kinase·ATP complexes, it fails to reproduce experimental structural data on Mg2+-ATP binding modes, ATP-protein binding modes, as well as ATP conformations within kinases. This happens despite the inclusion of NB-fix cross-terms for Mg2+–protein interactions. Perhaps, this is because large errors are still present in Mg2+–protein interactions, or perhaps, this results from a combination of errors in Mg2+–protein, Mg2+–nucleotide, and protein–nucleotide interactions. A further assessment is required.

There are two general approaches to addressing this problem. In one approach, the MM parameter space can be explored to directly target the experimental structures of Mg2+·Protein·Nucleotide complexes. Alternatively, Mg2+–protein, Mg2+–nucleotide, and protein–nucleotide interactions can be assessed separately, and then the improved models can be validated against the experimental structures of Mg2+·Protein·Nucleotide complexes. Here, we chose the latter approach. The advantage of the latter approach is that we do not rely on experimental structural data for potentially fixing the underlying physics of intermolecular interactions. Instead, we aim to assess and add/fix the fundamental physics of intermolecular interactions and then examine the extent to which such improvements modify the correspondence between simulated and experimental structures.

We focus on the assessment and improvement of these interactions in the polarizable AMOEBA model. We build on our earlier work3336 and preceding foundational developments of AMOEBA.5459 In an earlier work, we developed new Mg2+ LJ parameters that reproduce CCSD(T) data on Mg2+–water clusters and accurately predict the aqueous phase properties of Mg2+ (Table 1).34 In an earlier work, we also recalibrated the polarization and LJ interaction terms of carbonyls, hydroxyls, and carboxylate groups in proteins so that they responded equally well to both the high electric fields present near cations and lower electric fields. We had shown that the revised AMOEBA protein model, AMOEBABIO18-HFC, where HFC stands for high field correction, combined with many-body NB-fix (MBNB-fix) corrections to interactions between monovalent cations and proteins reduced errors in ion–protein interactions from 9.2 to 2.7 kcal/mol.36 The errors were estimated with respect to van der Waals-inclusive density functional theory (DFT) that was benchmarked against CCSD(T)/CBS and quantum Monte Carlo calculations as well as experiments.29,3335,60,61 We had also shown that the revised model retains its intrinsic reliability in predicting protein structure and dynamics in the condensed phase.36

Table 1.

Aqueous Phase Properties of Mg2+ Predicted by AMOEBABIO18 and Our Revised AMOEBABIO18-HFC22 Models34d

aqueous phase property experiment AMOEBABIO18 AMEOBABIO18-HFC22
hydration free energy (kcal/mol)  −207.2a  −200.9 ± 0.3  −206.4 ± 0.3
coordination number  6b  6  6
coordination first peak (Å)  2.0–2.12b  2.05  2.07
water residence time (μs)  1.5c  0.1  0.9
a

Taken from ref 62.

b

Taken from ref 63.

c

Taken from ref 64.

d

Hydration free energies are given for the substitution of 2LiCl → MgCl2. Water residence time is calculated using transition-state theory, where the energy barrier between the first and second shell is obtained from the potential of mean force as described in the Supporting Information. The Mg2+ parameters are provided in Table S1.

In this work, we first show that AMOEBABIO18-HFC22 dramatically improves the Mg2+–protein interaction energies, reducing mean absolute errors (MAE) from 17 to 10 kcal/mol. Since divalent cations were not considered during the calibration of AMOEBABIO18-HFC22, this result demonstrates the general applicability of our high-field corrections to AMOEBA. We then explore two different approaches to develop MBNB-fix cross-terms for Mg2+–protein interactions and show that our chosen approach further reduces the MAE to 6 kcal/mol. Next, we recalibrate the triphosphate parameters in ATP and evaluate its performance and predictions for describing Mg2+–nucleotide interactions in the aqueous phase. We show that the revised model continues to predict the experimental Mg2+-ATP binding free energy accurately. The revised model is also more flexible and shows high probabilities for Mg2+ to form mono-, bi-, and tridentate coordinations with ATP, as can be seen in NMR experiments discussed above. Then, we recalibrate protein–ATP interactions. Finally, we conduct MD simulations of two different Mg2+·Kinase·ATP complexes and evaluate the performance of the improved model against experimental structures. We show that our proposed improvements lead to a better agreement in local and global structures between MD and X-ray crystallography.

METHODS

QM Reference Data.

Reference energies and structures of ion–ligand and ligand–ligand clusters are obtained from the Perdew–Burke–Ernzerhof (PBE0)65,66 exchange–correlation density functional supplemented by Tkatchenko–Scheffler67 self-consistent corrections for dispersion. We refer to this dispersion-corrected DFT as PBE0+vdW. This hybrid functional mixes 25% exact exchange to alleviate the local or semilocal approximation of the DFT-based methods. This is particularly important when significant intramolecular charge transfer is involved, for example, for hydrogen-bonded systems or systems involving ions. All single-point calculations as well as geometry optimizations employing PBE0+vdW are performed with FHI-AIMS68 and with “really tight” basis sets. Total energies and electron densities are converged to 10−6 eV and 10−5 electrons, respectively, and geometries are optimized with a force tolerance of 10−3 eV/Å.

Our choice for PBE0+vdW stems from the three main reasons, discussed in more detail in our recent work.36 First, PBE0+vdW yields an accuracy of 0.3 kcal/mol in comparison to “gold standard” quantum chemical reference data for a wide range of intermolecular interactions in molecular dimers.69,70 Second, we have reported previously3335,60,61 that compared to quantum Monte Carlo calculations and CCSD(T) calculations in the complete basis set (CBS) limit, PBE0+vdW yields a MAE of 0.93 kcal/mol for interaction energies of monovalent cations (Na+, K+, and NH +4) with homogeneous clusters of various small molecules (waters, alcohols, amides, carboxylates, and aromatics). Finally, we have shown that with reference to LNO–CCSD(T), Mg2+ slightly overbinds water clusters by 1–2 kcal/mol, and within the Harmonic approximation, it yields gas-phase Mg2+-water binding free energies with a MAE of less than 1 kcal/mol with respect to experiments.35 Here, we further benchmark this method for interactions of Mg2+ ions with various neutral and negatively charged small molecules.

Reference CCSD(T) information to benchmark DFT + vdW is obtained using the local natural orbital (LNO) method7174 as implemented in the Mrcc package.75,76 Optimized geometries from PBE0+vdW are used for single-point energy LNO–CCSD(T) calculations. Extrapolation of LNO–CCSD(T) results toward the approximation-free CCSD(T) value and the corresponding local error estimates are performed using the Tight and very Tight LNO–CCSD(T) threshold sets72,73 according to the extrapolation scheme of ref 73. For the LNO–CCSD(T) computations, we employ Dunning’s correlation-consistent basis sets augmented with diffuse functions (aug-cc-pVXZ, X = T, Q, 5) for first row elements, and the corresponding core–valence basis sets77 for Mg2+. While deep-core electrons of all atoms are kept frozen, the subvalence electrons of Mg2+ have to be correlated for reliable results. Extrapolation toward the CBS limit of CCSD(T) results is performed via standard formulas,78,79 yielding CBS(X, X + 1) values from the aug-cc-pVXZ results. The remaining basis set incompleteness error (BSIE) of the final LNO–CCSD(T)/CBS(Q,5) interaction energies is estimated as the difference between the CBS(T,Q) and CBS(Q,5) results. The accuracy of the final tight-very tight LNO–CCSD(T)/CBS(Q,5) interaction energies can be conservatively characterized using a cumulative BSIE and local error estimate, which indicates that our reference results are within ±0.5 kcal/mol of the approximation-free CCSD-(T)/CBS ones for all studied complexes.

Induced dipoles and polarizability tensors are computed using second-order Møller–Plesset perturbation (MP2) theory80 implemented in Gaussian09.81 We use Dunning’s82 triple-ζ correlation consistent polarize valence only basis set augmented with diffuse functions (aug-cc-pVTZ), which has been shown to have comparable accuracy to CCSD(T) for calculation of molecular dipole moments and polarizabilities.83

MD Parameters.

All molecular dynamics (MD) simulations are carried out using Tinker9.84 Temperature is regulated using an extended ensemble approach85 and with a coupling constant of 0.1 ps. In isobaric simulations, pressure is regulated using a Monte Carlo approach86,87 and with a coupling constant of 0.1 ps. The equations of motion are integrated using the RESPA algorithm with an outer time step of 3 fs and an inner time step of 0.5 fs.88 Electrostatic interactions are computed using particle mesh Ewald (PME), with a direct space cutoff of 9 Å. Van der Waals interactions are computed for interatomic distances smaller than 9 Å. Both energy and virial are corrected with an isotropic long-range correction term.89 The convergence cutoff for induced dipoles is set at a value of 10−2 Debye in the condensed phase and 10−5 Debye in the gas phase. Other control functions and parameters are set to their default values. Energy optimizations employing MM models are performed using the optimize program in TINKER890 using an RMS potential gradient cutoff of 10−2 kcal mol−1 Å−1.

MD of Mg2+·Kinase·ATP Complexes.

The coordinates for the starting structure of glycogen synthase kinase-3 beta (GSK3β) are taken from PDB ID 1PYX, which was solved at a resolution of 2.4 Å.91 The coordinates for the starting structure of cyclin-dependent kinase 2 (CDK2) are taken from PDB ID 1B38 whose structure was solved at a resolution of 2 Å.92 In both cases, we simulate the monomeric forms of the enzyme, and in neither case, the enzymes are bound to substrates. Missing residues, all of which were >10 Å away from the ATP binding pocket, are added using Modeller.93 The longest contiguous stretch of residues that we add for CDK2 is from residue numbers 36–43, which belongs to a loop. In the case of GSK3β, the longest contiguous stretch of residues we add is from residue numbers 86–90 and that also belongs to a loop. Protein termini are capped using ACE and NME, and the protonation states of titratable residues are assigned their standard protonation states under physiological conditions. Missing hydrogens are added by optimizing hydrogen bond networks using PDB 2PQR.94 Crystal waters are retained, and additional water molecules are introduced to fill cubic boxes with edge sizes of 98.8 and 90.7 Å for GSK3β and CDK2, respectively. The GSK3β box contains 29,698 water molecules, and the CDK2 box contains 22,465 water molecules. A salt concentration of 20 mM is set by randomly replacing water molecules with appropriate numbers of Na+ and Cl ions. Additional 2 and 7 Cl ions in CDK2 and GSK3β boxes are added to make the systems charge neutral. Following energy minimization, the boxes are simulated under NPT conditions (1 atm, 298 K) with harmonic restraints applied to all protein backbone atoms, ATP, and Mg2+ ions. Initially, a spring constant of 30 kcal/mol/Å2 is used, which is decreased to 0 in steps of 6 kcal/mol/Å2 per 900 ps. After this, we perform unrestrained NPT simulations for 300 ns.

Double-Decoupling Method.

The standard free energy change (ΔFi) associated with the binding of Mg2+ to a specific site (i) on ATP in solution is computed as95,96

ΔFi=ΔFrstrATPMg(i)(aq)ΔFMg(aq)ΔFrstri (1)

Here, ΔFrstrATP-Mg(i)(aq) is the free energy change associated with decoupling the vdW and electrostatics interactions of Mg2+ from ATP and the solvent, while it is held restrained at binding site i.ΔFMg(aq) is the solvation energy of Mg2+ in the absence of ATP, and ΔFrstri is the contribution from the restraining potential that is subtracted from the solvation energy of Mg2+.

The electrostatic interactions of Mg2+ are decoupled from the rest of the system by linearly scaling the multipole and polarization parameters with a scaling factor (λ) that is changed from λ=1 to λ=0 with a constant step size of Δλ=0.1. As electrostatic interactions of Mg2+ are decoupled, we also turn on restraints that keep the Mg2+ bound to site i. vdW interactions of Mg2+ are decoupled using the following set of scaling values λ={1,0.9,0.8,0.75,0.7,0.65,0.62,0.6,0.55,0.5,0.4,0}. Therefore, a total of 22 independent simulations with different values of λ for electrostatics and vdW are run for each system. We use Bennet acceptance ratio97 to calculate the free energy difference between neighboring λ. Since we are using Ewald summation to compute the electrostatics with periodic boundary conditions, we ensure that the charge of the system remains neutral by simultaneously decoupling 2Cl ions present in solution.35 Therefore, eq 1 is rewritten as

ΔFi=ΔFrstrATPMg(i)Cl2(aq)ΔFMgCl2(aq)ΔFrstri (2)

In all cases, we apply restraints using a single harmonic potential defined as u(r)=krr02, where k is the spring constant and r0 is the equilibrium distance that we chose separately for each binding mode as the average distance observed in 30 ns of unrestrained MD in the λ={1, 1} state. For monodentate binding modes, restraint is applied between Mg2+ and the coordinating phosphate oxygen. For multi-dentate binding modes, the restraint is applied between Mg2+ and the centers of masses of the coordinating phosphate oxygens. The contribution from the restraining potential is estimated analytically as

ΔFrstri=kbTln(4πc0r2eβu(r)dr) (3)

where c is the unit concentration. To assess the effect of the choice of the spring constant, we compute ΔFi for the original model using two different spring constants, k={5, 10}kcal/mol/2. We note that ΔFi differences by less than 0.2kcal/mol. The values we present in the Results and Discussion section are computed using k=5kcal/mol/2.

The error associated with the binding free energy ΔFibind of each mode is computed as δΔFibind=jδjΔFibind2, where δjΔFibind is the Monte Carlo error for each window j for each mode i. We keep track of both the free energy and associated error as a function of simulation time, and based on this, we conduct 15 ns sampling for each λ value to get errors below 0.5kcal/mol

Multiparameter Optimization.

MM parameters are optimized using the Nelder–Mead simplex-based algorithm as implemented in the ParOpt software.98,99 The Nelder–Mead coefficients are set to their standard values for reflection (α=1.0), contraction (β=0.5), expansion (γ=2.0), and size (δ=0.5). For each optimization, a random point within the domain of the parameter space is selected, and the algorithm is iterated until the simplex is converged; that is, when the rootmean-square distance between the simplex vertices and the centroid is below 105 or until 1000 steps are completed. The parameters with the lowest value for the target function are selected after the algorithm is run starting at many different initial values. The parameter domain that we explore as well as the number of optimizations we perform are context dependent and, therefore, are discussed at appropriate places in the Results and Discussion section.

RESULTS AND DISCUSSION

This section is organized as follows. We first assess and propose improvements to the Mg2+–protein interactions. To account for high electric field response of phosphates, we present a recalibration of nucleotide triphosphate parameters. Then, we propose a new set of Mg2+–phosphate interaction cross-terms and assess the model’s performance in predicting Mg2+-ATP binding in the aqueous phase. Finally, we propose a new set of nucleotide–protein interaction cross-terms and assess the performance of the revised model in simulating Mg2+·Kinase·ATP complexes.

Mg2+–Protein Interactions.

To assess the performance of AMOEBA in predicting Mg2+–protein interaction energies, we extract Mg2+ clusters from PDB structures and examine how AMOEBA performs in predicting interaction energies between Mg2+ ions and the rest of the cluster. We extracted clusters from the PDB in an unbiased manner as follows. First, we download all X-ray structures from the PDB that contain Mg2+ ions and have resolutions better than 1.5 Å.4 From these structures, we extracted all atoms that are within 6 Å from Mg2+. Even if a single atom of an amino acid is within this cutoff, the entire amino acid is included as part of the cluster. Then, we discard all clusters that contain nonprotein or nonwater atoms or more than one Mg2+ ion. At this stage, we are left with 29 clusters. These clusters contain a single Mg2+ ion, peptide fragments, and water. We assume the default protonation state at physiological pH for titratable residues as well as for water molecules. This approximation does not affect the estimated errors as the same assumption is made in both QM and MM calculations. Peptide fragments in each cluster are then analyzed, and if their ends are separated by less than two amino acids, the connecting amino acids are also included in the cluster, even if the connecting amino acids are outside the 6 A cutoff. The peptide fragment ends are then capped with ACE and NME, and all missing hydrogens, including those on water, are added. The clusters are then energy-minimized using the original AMOEBA model with position restraints applied to atoms that were resolved in PDB structures.

The smallest cluster in our set consists of 65 atoms, and the largest cluster consists of 285 atoms. We note that in 18 of these 29 clusters, Mg2+ is 5-fold or 6-fold coordinated. In the remaining clusters, the smaller coordination is perhaps due to missing water molecules, although we understand that coordination number is not just a property of the ion but is also influenced by the environment.100 These clusters also differ from each other in terms of the numbers of hydroxyls, carboxylates, and carbonyls as well as the number of waters in the Mg2+ ion’s first coordination shell (see Table S2 in the Supporting Information for details). We note, however, that the hydroxyl oxygen is underrepresented in this cluster set, but this is expected for Mg2+ based on a previous PDB survey.101 Additionally, we note that these clusters also provide a good distribution of geometries around energy minima. The peak of the distribution of distances between Mg2+ and their first shell coordinators is close to the optimal distance in Mg2+–water clusters,34,63 and both the repulsive hard-wall and first shell outer boundary distances are represented (see Figure S3 in the Supporting Information).

Using each of these clusters, we determine the interaction energies (ΔE) between Mg2+ and everything else in the clusters. Figure 1 compares the interaction energies obtained from the original AMOEBA model against those obtained from our benchmarked QM method (PBE0+vdW). We note that predictions from the original model correlate excellently with reference QM in the entire range of interaction energies [−210, −820] kcal/mol. However, both the MAE of 17 kcal/mol and the maximum error of 53 kcal/mol are very high. The magnitudes of these errors are similar to that noted for the CHARMM DRUDE MM model in the case of Ca2+ ions, despite the inclusion of NB-fix corrections.31 When considering only those clusters in which Mg2+ is 6-fold coordinated, the MAE of 22 kcal/mol is even larger. We provide an assessment of 6-fold coordinated clusters separately because 6-fold coordination is known to be the preferred coordination of Mg2+ ions when they bind proteins.5

Figure 1.

Figure 1.

Performance of AMOEBA models in predicting Mg2+–protein interaction energies. The scatter plot compares the interaction energies computed using benchmarked QM (PBE0+vdW), against four AMOEBA models: AMOEBA,59 AMOEBABIO18,102 AMOEBABIO18-HFC22,36 and AMOEBABIO18-HFC22 + MBNB-fix that includes Mg2+–protein MBNB-fix corrections. See the related text for explanations of these models. Next to the scatter plot, we show the distribution of absolute errors in the predictions of AMOEBA models computed with respect to PBE0+vdW. The mean and median are represented by dashed and solid lines, respectively.

Figure 1 also shows the performance of another AMOEBA model, which we refer to as the “AMOEBABIO18” model. The modifications that this model has concerning Mg2+–protein interactions are revised Glu/Asp carboxylate parameters, modified Mg2+ polarizability, and LJ parameters and explicit NB-fix cross-terms for interactions of carboxylates with Mg2+ and amines.41 The AMOEBABIO18 model performs better than the original model, with a smaller MAE of 10 kcal/mol and also a smaller MAE of 13 kcal/mol if only 6-fold clusters are considered. The maximum error of 45 kcal/mol is also lower compared with the original model.

Next, we determine the performance of our revised AMOEBABIO18-HFC22 model.36 This model has new Mg2+–water parameters that reproduce CCSD(T) cluster data and accurately predict aqueous phase properties of Mg2+ (Table 1).35 Additionally, the revised model has new parameters for hydroxyls, carbonyls, and carboxylates in proteins that were shown to perform well at the high electric fields present near cations.36 However, the revised model does not have any NB-fix corrections for Mg2+ interactions with proteins. We note that even without NB-fix corrections, it performs better than the original and AMOEBABIO18 models. The MAE reduces to 10 kcal/mol for all clusters and 9 kcal/mol for 6-fold clusters. Additionally, the maximum error is now reduced dramatically to 22 kcal/mol.

To further reduce this error, we develop ligand-specific NB-fix-type LJ cross-terms for Mg2+. The ligands are carbonyls (Asn/Gln/backbone/ACE/NME), carboxylates (Asp/Glu), and hydroxyls (Tyr/Ser/Thr). The typical strategy to develop these cross-terms is to use reference energies from two-body interactions.3032,3744 For example, reference energies between Mg2+ and NMA can be computed at different distances and used as target data to optimize the Mg2+–carbonyl LJ cross-terms. This strategy, however, does not account for many-body cooperativity effects. To account for them, we have shown that LJ cross-terms can also be determined from reference data on many-body clusters.99

To examine whether we should use the traditional two-body cluster approach or use higher-order clusters, we carried out the following exercise. We first determine Mg2+–carbonyl and Mg2+–carboxylate LJ cross-terms using reference data on Mg2+–NMA and Mg2+–acetate dimers, respectively, and use them to predict interaction energies in 6-fold clusters made up of waters, NMA, and acetate. Our set of 6-fold clusters consists of 100 geometries belonging to 10 unique chemical compositions shown in Figure S4 of the Supporting Information. Alternatively, we determine cross-terms using the same 6-fold clusters we used for testing the results of the former strategy and use them to predict dimer energies. In all cases, the reference energies are determined from PBE0+vdW. For Mg2+–acetate and Mg2+–NMA dimers, we also benchmark PBE0+vdW against LNO–CCSD(T)/CBS (Figure 2a) and find that it performs excellently. The NMA and acetate parameters are taken from our earlier work that perform better at reproducing high-field reference data.36 In both strategies, we use the Nelder–Mead algorithm for optimizing LJ cross-terms using a parameter range of [3.0, 4.2] Å and [0.1, 0.9] kcal/mol for R0 and ϵ, respectively. We carried out 100 different optimizations starting with different seeds, which produced 100 different sets of parameters. We chose the parameter set that yields the smallest error with respect to the target energies.

Figure 2.

Figure 2.

(a) Distance-dependent interaction energies of Mg2+ with acetate and NMA. “Recal.+2B” refers to our revised AMOEBA model for small molecules (AMOEBA09-HFC22)36 along with Mg2+–carbonyl and Mg2+–carboxylate LJ cross-terms determined using Mg2+–NMA and Mg2+–acetate dimers. “Recal.+6F” refers to our revised AMOEBA model for small molecules along with Mg2+–carbonyl and Mg2+–carboxylate LJ cross-terms determined using 6-fold clusters of NMA, acetate, and water. (b) Distributions of the absolute error in interaction energies of Mg2+ ions with 6-fold clusters computed using the “Recal.+2B” and “Recal.+6F” models.

The results of this exercise are shown in Figure 2. We make three key observations. First, LJ cross-terms determined using two-body clusters do not perform as well compared to those determined using 6-fold clusters in estimating the interaction energies of 6-fold clusters. Second, the LJ cross-terms determined using 6-fold clusters perform poorly in predicting the interaction energies of Mg2+-NMA/acetate dimers. This means that cross-terms determined by using higher order clusters should be used with caution to compute energies of lower order clusters and vice versa. Third, even when considering only 6-fold clusters, we are unable to reduce the MAE below 5 kcal/mol, which means that we may have reached the predictive limit of AMOEBA’s functional form. Together, these observations imply that there is still essential physics missing from the AMOEBA model. Nevertheless, these results suggest that since we intend to use Mg2+ for simulating protein binding in 6-fold coordination geometries, we need to determine cross-terms using higher order clusters.

The other issue concerns whether LJ cross-terms should be determined by using small-molecule representatives of proteins or directly from protein clusters. Previously, we had chosen to do this using protein clusters, as the multipoles in small molecules slightly differ from their representative chemical groups in proteins.36 We use the interaction energies from the 6-fold PDB clusters as target data and determine Rij0 and ϵij such that they minimize the root-mean-square error (RMSE) to this target. The best parameter set obtained after the exhaustive search is provided in Table S3 of the Supporting Information. We note that the biggest difference between the terms calculated from the mixing rules relates to making the energy well ϵij deeper. Figure 1 shows that for 6-fold clusters, these new LJ cross-terms further reduce MAE to 6 kcal/mol and the maximum error to 12 kcal/mol. Even when all the clusters are taken into account, the MAE remains around 6 kcal/mol, and the overall distribution of errors also shrinks, except for a single case whose error lies around 17 kcal/mol.

Overall, our revised parameters dramatically reduce MAE in Mg2+–protein interaction energies from 17 to 6 kcal/mol and maximum error from 53 to 17 kcal/mol. For 6-fold clusters that are expected to be common in Mg2+·Protein·Nucleotide complexes,101 MAE reduces from 22 to 6 kcal/mol and the maximum error reduces from 53 to 12 kcal/mol.

Polarization, LJ, and Torsional Parameters of the Triphosphate.

In earlier work, we showed that the accuracy of an MM model to describe cation–ligand interactions depends strongly on the ability of ligands to respond to the high electric field present near cations.3336 To assess this in AMOEBA’s ATP model,53 we compute the induced dipole moment of dimethyl phosphate (DMP) as a function of its distance from a unit positive point charge. DMP is also used in the parametrization of other MM models of nucleotides, including CHARMM and AMBER.103105 We did this calculation in two ways. In the first approach, we compute DMP’s field response using AMOEBA’s DMP model.106 We note that the atomic polarizabilities of AMOEBA’s DMP reproduce the molecular polarizability tensor obtained from the MP2 theory. Our calibration of atomic polarizabilities against MP2 yields very similar values compared to the original DMP’s model. In the second approach, we transfer the polarizabilities of AMOEBA’s triphosphate group, which is the same for all the three phosphate groups, to AMOEBA’s DMP model. The results of these calculations are shown in Figure 3 and compared to reference values obtained from MP2 theory. We find that DMP’s polarization model outperforms the triphosphate’s polarization model in predicting both low and high electric field responses. We, therefore, transfer the atomic polarizabilities of DMP to ATP’s triphosphate group, as listed in Table S4 of the Supporting Information.

Figure 3.

Figure 3.

Induced dipole moment of DMP as a function of distance (|r|) from a unit-positive point charge. The primary component of the induced dipole, μind, which is parallel to the distance vector, is shown. DMP-pol refers to the set of atomic polarizabilities in AMOEBA’s DMP model,106 and TP-pol refers to the unique set of atomic polarizabilities in all ATP/GTP’s triphosphate groups.53

Since we modify triphosphate’s polarization term, we will need to recalibrate triphosphate’s interaction with water. To do this, we extract 10 random snapshots from an MD simulation of ATP in water. Then, we discard all water molecules, except those that coordinate directly with ATP’s triphosphate group. From these ATP–water clusters, we computed interaction energies between ATP and all water molecules. We use the interaction energies from the original model as reference and recalibrate the LJ parameters of the triphosphate oxygens to minimize the RMS error with respect to the reference energies. After recalibration, the RMS error is 0.04 kcal/mol/water. We also note that the new parameters are only slightly different from the original ones and are provided in Table S4 of the Supporting Information. To further validate the new parameters, we compute interaction energies of triphosphate with waters placed at varying distances from the triphosphate oxygens and compare them against PBE0+vdW. To avoid protonation of triphosphate in QM calculations, we perform these calculations in the presence of all waters within the triphosphate’s first coordination shell. The results are shown in Figure S5 of the Supporting Information. We find that the recalibrated model performs well against QM in describing the triphosphate–water interaction energies.

Since polarization and LJ parameters are modified, we examine whether dihedral parameters also need to be revised. Since the triphosphate is defined using two dihedrals, we generate a two-dimensional potential energy surface (2D-PES). We rotate dihedral angles Oα-Pβ-Oβ-Pγ(ψ) and Pα-Oα-PβOβ(ϕ) in increments of 15° and generate 576 geometries. We energy-minimize these geometries with dihedral restraints using the original AMOEBA model. We then compute the total potential energy of each geometry and scale the PES by subtracting the energy of the most stable geometry. Figure 4 compares the PES computed from the original model against that obtained from PBE0+vdW. We note that the original model deviates from QM by an average of 6.3 kcal/mol. Furthermore, the energetic barrier around ψ=ϕ=0 is substantially weaker in the original model. Additionally, the relatively more stable geometries are more localized in ϕ, which perhaps enhances the torsional rigidity of the original model.

Figure 4.

Figure 4.

2D-PES of triphosphate dihedrals Oα-Pβ-Oβ-Pγ(ψ) and Pα-Oα-Pβ-Oβ(ϕ). The QM2D-PES is computed using PBE0+vdW. “Original” refers to the ATP parameters developed by Walker53 et al.,41 and “Revised” refers to the modifications of the triphosphate parameters in this work.

To recalibrate the dihedrals, we construct an error function that includes not only the 2D-PES determined above but also a set of four optimized structures, each in the local minima of the cis–cis, cis–trans, trans–cis, and trans–trans conformations, as shown in Figure S6 of the Supporting Information. We carry out a total of 100 Nelder–Mead optimizations starting from randomly selected simplexes and explore a parameter range of [−20, 20] kcal/mol for the coefficients of the dihedral term. The overall parameter space explored is shown in Figure S7 of the Supporting Information. The parameter set that minimizes the error function reduces the PES MAE from 6.3 to 1.4 kcal/mol, and the largest rmsd in structure drops from 0.9 to 0.5 Å. Both the original and revised parameters for the triphosphate are provided in Tables S4 and S5 of the Supporting Information. Figure 4 shows the 2D-PES generated from the final set of polarization, LJ and dihedral parameters. We note that the main features of the 2D-PES that were concerning in the original model are now resolved.

Mg2+–Triphosphate Interactions.

To determine Mg2+–triphosphate LJ cross-terms, we make a minor adjustment to the protocol used above for determining Mg2+–protein interactions. Instead of using triphosphates, we will use their small-molecule representative DMP. The main reason we do not use triphosphates is because they carry a high negative charge of −4 eu. This high negative charge will produce long electron density tails in gas-phase QM calculations that are likely to be much shorter in the condensed phase due to intermolecular electron repulsion. Since our goal is to get simulations right in the condensed phase, we decided not to capture the electrostatic effect of the long tails in the gas phase. Therefore, we avoid using triphosphates to obtain LJ cross-terms and instead use DMP.

We obtain LJ cross-terms with DMP using reference energies and structures associated with the following substitution reaction

[MgW6]2++n(DMP)[Mg(DMP)nW6n]2++nW (4)

where n={1, 2}. Table 2 lists the substitution energies computed using different methods. We first note excellent agreement between PBE0+vdW and LNO–CCSD(T). The large discrepancies that we find in the original model are partially mediated by the use of our new Mg2+ parameters and are further reduced by introducing LJ cross-terms between Mg2+ and the phosphate oxygens. We also note that all models correctly capture the optimized geometry obtained from QM (Figure S8 and Supporting Information). The LJ cross-terms are provided in Table S6 of the Supporting Information. We note that transferring Mg2+–DMP LJ cross-terms to Mg2+–triphosphate LJ interactions can lead to errors due to differences between triphosphate and DMP multipoles. However, the differences in multipoles are small, and we expect the associated error to be on the order of 1–2 kcal/mol.

Table 2.

Substitution Energies in kcal/mol Computed for Eq 4 Using Different Methods

method n = 1 n = 2 MAE
LNO–CCSD(T) −202.0 −315.9
PBE0+vdW −201.9 −316.8 0.5
AMOEBABIO18 −214.3 −330.5 13.5
AMOEBABIO18-HFC22 −209.8 −323.1 7.5
AMOEBABIO18-HFC22 + NB-fix −206.5 −315.5 2.4

Mg2+-ATP Binding in Bulk Water.

After introducing modifications to the triphosphate group and its interaction with Mg2+, we now predict the structure and thermodynamics of Mg2+-ATP binding in the aqueous phase.

We first identify “stable” binding modes of Mg2+-ATP complexes in the aqueous phase. Overall, there are 15 possible coordination binding modes between Mg2+ and the different oxygen atoms of ATP (see Figure S9 in the Supporting Information). We solvate optimized geometries of each of these 15 6-fold coordinated binding modes in separate 70 Å cubic boxes of water with 100 mM of KCl. The choice of the specific ionic strength is to emulate experimental conditions in which KCl is used as a buffer for determining Mg2+-ATP binding free energies.42 Note that there are two additional K+ values in each box to balance the net charge. We first conduct 3 ns long NPT simulations (1 atm, 298 K) with harmonic distance restraints (25 kcal/mol/Å2) between Mg2+ and its six coordinators. We then reduce these harmonic restraints over the next 1.2 ns of the NPT simulations. Then, we carry out an additional 300 ps of NPT simulations to determine average box volumes. Next, we select snapshots from the final 300 ps trajectories that have volumes closest to the average volumes and use them to initiate 15 unrestrained NVT runs, one for each binding mode. We carry out NVT simulations for 30 ns.

Analysis of these trajectories reveals the following seven stable binding modes in the revised ATP model: three monodentate modes (α,β, and γ), three bidentate modes (αβ,βγ, and αγ), and one tridentate mode (αγγ). These binding modes are shown in Figure 5a. Note that αγ and αγγ are shown together as we find them to switch frequently between each other. Mg2+ ions in the remaining eight modes switch coordinations to form one of these seven binding modes (see Figure S10 in the Supporting Information). In comparison to the revised model, the original model shows only four binding modes, α,β,γ, and αγ consistent with a previous study.53 These four binding modes are subsets of the seven binding modes observed with the revised model. This perhaps means that our revised model is more flexible in that it can adopt more conformations in the presence of Mg2+.

Figure 5.

Figure 5.

(a) Predicted modes of binding of Mg2+ to ATP in solution. Mg2+ is shown as pink spheres, and the truncated adenosine groups are colored green. The binding free energies of Mg2+ to each mode ΔFi are provided in units of kcal/mol. “Original” refers to the ATP parameters developed by Walker53 et al.41 and the Mg2+ parameters used in that development. “Revised” refers to the modifications to triphosphate and Mg2+–triphosphate parameters introduced in this work and the Mg2+ parameters we developed previously.35 (b) Mg2+-ATP binding free energies from different models are compared to experiment. Note that estimates from the revised model are predictions, as they are a direct outcome of the model that we revised solely based on QM targets consisting of local interactions. This is not the case with the original model. Note also that two estimates are provided for the original model, one taken from Walker et al.53 and the other determined in this work. The gray circles are estimates from seven experimental methods, compiled by others,42 and summarized in Table S7 in the Supporting Information. The dashed black line is the average of all experiments, and the green shaded region represents the associated standard deviation.

To quantify Mg2+-ATP affinity in each of these binding modes, we compute standard binding free energies using the double decoupling method95,96 (details provided in Methods section). The results are listed in Figure 5a. We make two main observations. First, revision of the model changes the relative strengths of binding modes seen in the original model. Specifically, in the original model, the energetically most favored binding mode is α, followed by αγ and γ. However, in the revised model, αγ/αγγ is most favorable, followed by βγ and γ. In the revised model, we find that any interaction involving the γ oxygen leads to stronger binding. Second, while the original model did not produce any of the three binding modes (αβ,βγ, and αβγ) preferentially seen in PDB structures of Mg2+·Protein·ATP complexes, the revised model suggests that the βγ mode does exist in the aqueous phase with high probability. The other prominent bidentate binding mode, the αβ mode also exists in the aqueous phase, but with lower probability compared to the βγ mode. We do not observe the third binding mode observed in PDB structures, the αβγ mode, which suggests that it perhaps results from complexation with protein.

Figure 5b shows the total Mg2+-ATP binding free energy, estimated by Boltzmann averaging the standard binding free energies of all modes ΔF=kbTlnieβΔFi. The standard error, as shown by Walker et al.,53 is δΔF=iδΔFi2, where δΔFi is the error of each mode. We first note excellent agreement between the previously calculated ΔF of 7.2±1.8kcal/mol53 and the one we obtained with the same original model (7.4±0.7kcal/mol) but following a systematic protocol to identify binding modes. The revised model yields a ΔF of -9.1±0.8kcal/mol. All of these values are within the experimental range compiled by Goldberg and Tewari42 and summarized in Table S7 in the Supporting Information. Note that the predictions from the revised model are a direct outcome of the revisions that we made to the model using QM targets consisting only of local interactions.

Protein–Triphosphate Interactions.

To model binding of nucleotides to proteins, we also need to assess the accuracies of protein–triphosphate interactions. Figure 6 shows the interaction energies between DMP and Lys/Arg for different distances between the charged groups. For DMP–Arg interactions, we find overall good agreement between PBE0+vdW and estimates from both AMOEBABIO18 (MAE = 1.7 kcal/mol) and revised AMOEBABIO18-HFC22 (MAE = 1.9 kcal/mol) models. The inclusion of NB-fix cross-terms between phosphate oxygen and Arg guanidinium nitrogen has only a small impact on MAE. Therefore, we decided to stick to the LJ parameters computed from mixing rules. In the case of DMP–Lys interactions, the MAE is 3.2 kcal/mol for the AMOEBABIO18 model and 3.4 kcal/mol for the AMOEBABIO18-HFC22 model. The inclusion of LJ cross-terms reduced the MAE to 0.5 kcal/mol. These cross-terms are listed in Table S6 of the Supporting Information.

Figure 6.

Figure 6.

DMP–lysine and DMP–arginine interaction energies for nine different dimer geometries. The dimer geometries differ from each other mainly in terms of their distances ranging between 3.4 and 5 Å.

MD Simulations of Mg2+·Kinase·ATP Complexes.

To assess the performance of our revised model, we perform MD simulations of two protein kinases, one complexed with ATP and two Mg2+ ions and the other complexed with ATP and one Mg2+ ion. The kinase structure complexed with two Mg2+ ions is that of glycogen synthase kinase-3 beta (GSK3β), and the kinase structure complexed with one Mg2+ ion is that of cyclin-dependent kinase 2 (CDK2). We perform 300 ns long MD simulations using both the “original” and the “revised” models. The “original” model refers to the AMOEBABIO18 model with ATP parameters developed in 2021.53 The revised model refers to the AMOEBABIO18-HFC22 model, along with all the revisions made in this work.

Figure 7 shows the root-mean-square deviation (rmsd) of backbone atoms determined with respect to the X-ray structures. Overall, we find that the revised model exhibits a smaller deviation relative to X-ray structures. The difference in rmsd between the two models emerges primarily from the structures of loops, which remain closer to the X-ray structure in the revised model. This is deduced quantitatively from examining rmsds computed without including residues that belong to loops, and in these calculations, the rmsds in both models are similar. This is also clear from the visual inspection of the superimposed structures shown in Figure 7. This is consistent with what we noted previously36 in the case of another protein where the original model produced more flexible loops than the revised model. We had also noted previously that the revised model performed better than the original model at reproducing second-order NMR parameters.36 Of particular interest is the Gly-rich loop in GSK3β that interacts directly with bound ATP and is known to modulate its catalytic activity.107 We find that in both the original and the revised models, the Gly-rich loop is flexible, but in the original model, we see large deviations from the X-ray structure. In the revised model, the fluctuations are around the X-ray structure. This difference is not noted in the case of CDK2, which suggests that the differences noted in the case of GSK3β′s Gly-rich loop may be due to the enhanced electrostatics from the second Mg2+ ion, although further investigation is necessary. Overall, the revised model performs better at reproducing the global structures of the complexes.

Figure 7.

Figure 7.

Effect of model revision on the global structure of the (a) CDK2 and (b) GSK3β complexes. The figures on the left show backbone rmsds with and without the inclusion of residues that belong to loops. The figures on the right show 20 equally spaced snapshots from simulations superimposed over X-ray structures (green) used for starting simulations. Mg2+ ions are shown as spheres.

Next, we evaluated the coordination environments of Mg2+ ions. Figure 8 shows the inner shell (<3 Å) coordination numbers of Mg2+. In CDK2’s X-ray structure, Mg2+ is resolved in the αβγ binding mode to ATP. Mg2+ is 6-fold coordinated, and the remaining three coordinators are N132, D145, and one water. In the original model, Mg2+ continues to coordinate with N132, D145, and one water but its binding to ATP switches to a αγ mode, dropping the coordination number from 6 to 5. In the revised model, Mg2+ also continues to coordinate with N132, D145, and one water. The binding to ATP also switches, but to a αγγ/ααγ mode, leaving the coordination number around 6.

Figure 8.

Figure 8.

Coordination numbers of Mg2+ ions in simulations compared to the X-ray structures. The figures on the left show total coordination numbers as functions of simulation time. The dashed lines show averages over the entire lengths of the simulations. The text on the right shows probabilities of Mg2+ coordination with different ligands. “X” and “–” imply whether the coordination is present or absent in X-ray structures.

In GSK3β’s X-ray structure, neither of the two Mg2+ ions is resolved in 6-fold coordinated states: one is 4-fold coordinated and the other is 5-fold coordinated. One Mg2+ coordinates ATP in the αγ mode and the other Mg2+ coordinates ATP in the βγ mode. The original model retains the αγ binding mode, but βγ switches to the β mode. Both Mg2+ ions have average coordination numbers close to 6, and in both cases, the new coordinators that Mg2+ ions pick up to fill their inner shells are waters. The revised model also retains the αγ binding mode of Mg2+ to ATP, but the βγ binding mode switches to the γ mode. Both Mg2+ ions also have average coordination numbers close to 6, but in contrast to the original model, the new coordinators that Mg2+ ions pick up to fill their inner shells are not waters but protein groups. One Mg2+ picks up Q185, and the other picks up D181. To further access the geometry of Mg2+ ions, we calculate the distance between the two Mg2+ ions. This applies only to GSK3β. In the X-ray structure, this distance is 4.1. In both models, the average distance is a bit larger and is around 5 (see Figure S11 in the Supporting Information). Nevertheless, we do note that the revised model does sample short distances in the interval of [3.6–4.3] , which the original model does not. Overall, while both models only partially retain the Mg2+ coordination geometries seen in X-ray structures, the revised model performs better at retaining coordination numbers.

Next, we assess polar contacts between ATP and protein. These are summarized in Figure 9 and are computed using a protocol described in our earlier work.108 We focus our attention on the three main parts of ATP: the nitrogen atoms from the base (N1 and N6), the hydroxyl oxygens from the sugar (O2′ and O3′), and the triphosphate oxygens (OA, OB, and OG). Overall, we note that while both models retain the contacts of adenosine base, the revised model performs discernibly better at retaining the contacts of sugar hydroxyls and triphosphate oxygens.

Figure 9.

Figure 9.

Probabilities of polar contacts between ATP and the protein in simulations compared to those observed in X-ray structures. Lines denote contacts between donors and acceptors and the symbol “o–o” denotes salt bridges. The ATP atoms are shown as circles filled in blue, and protein atoms are shown as circles filled in orange. The probability of a given contact is color-coded on the lines connecting the circles. Contacts with a probability lower than 0.2 are not shown.

Finally, we examine the conformation of the triphosphate group in the protein binding site. Figure 10 shows a 2D histogram of the dihedral angles of O3α-Pβ-O3β-Pγ and Pα-O3α-Pβ-O3β. In both complexes, our revisions to AMOEBA shift the distributions away from the centers ψ=ϕ=0 and toward larger angles. Specifically, in CDK2, the two density maxima in the original model shift from around (ψ,ϕ)=(3030,30,30 to 60,30,90,30. In GSK3β, a single density maximum around (45,45) splits into two density maxima around 75,60,90,45. Comparing these 2D histograms to Figure 4 might suggest that the revisions to the triphosphate dihedral potentials contribute to this shift in dihedral distributions. However, note that substantial changes are also made to protein polar groups as well as to interactions between Mg2+, proteins, and nucleotides, and so we consider this shift in dihedral distributions to be a combined effect of all the changes introduced in the MM model. Importantly, in both cases, our revisions shift the distributions discernibly closer to the angles observed in X-ray structures. We also examine distributions in the triphosphate Pα-Pβ-Pγ angle and find that both models perform equally well against the angles noted in the X-ray structures.

Figure 10.

Figure 10.

2D histogram of the triphosphate dihedral angles of O3α-Pβ-O3β-Pγ(ψ) and Pα-O3α-Pβ-O3β(ϕ). The red triangles represent the dihedral angles from different X-ray structures from the PDB, and the blue triangles represent the X-ray structure used for these simulations. Triangles labeled “CDK2 + Cyclin A3” and “CDK2 + KAP” are in X-ray structures bound to cyclin A3 and KAP, respectively.

Taken together, we note that the revised model performs better than the original model in reproducing the global X-ray structure as well as the local geometries at catalytic sites, including Mg2+ coordinations, ATP coordinations, and ATP conformations. However, there are still differences between local geometries predicted by the revised model and the X-ray structure.

CONCLUSIONS

We present an improved polarizable AMOEBA model for simulating Mg2+·Protein·Nucleotide complexes. We follow an approach in which we first make specific improvements to Mg2+–protein, Mg2+–nucleotide, and protein–nucleotide interactions, assessing them individually by using benchmarked QM and experimental data. We then evaluate the performance of these improvements by carrying out MD simulations of Mg2+·Kinase·ATP complexes and comparing MD structural data against X-ray structural data.

We first show that our revised AMOEBA polarizable model for proteins, AMOEBABIO18-HFC2236 that responds better to high electric fields and was built on the strong foundations of the AMOEBABIO18 protein model, does better at predicting Mg2+–protein interactions. Inclusion of many-body LJ cross-terms (MBNB-fix) further improves the accuracy and brings the MAE down to 6 kcal/mol. Clearly, there is still room for improvement in Mg2+–protein interactions to bring MAE down to benchmarked DFT accuracy or to the almost DFT level accuracy of 2.7 kcal/mol that we achieved previously36 for interactions of proteins with monovalent cations. Note that employing LJ cross-terms developed using reference data from two-body interactions, which is a typical strategy employed in the development of polarizable and nonpolarizable models, yields larger errors in many-body interactions. Conversely, employing LJ cross-terms developed using reference data from many-body interactions yields larger errors in two-body interactions. Therefore, the many-body LJ cross-terms that we provide should not be used for studying two-body Mg2+–protein interactions. This also implies that there is still some essential physics that is missing in the description of these interactions.

We also present a new set of the molecular polarizability, LJ terms, and torsional angles of the triphosphate group, as well as new Mg2+–phosphate cross-terms that accurately predict Mg2+-ATP binding free energies in the aqueous phase, without the need for any a posteriori adjustments. The manner in which Mg2+ associates with ATP in the aqueous phase (Mg2+-ATP binding modes) remains ambiguous from experiments, and there is supporting evidence from NMR that Mg2+ can form mono-, bi-, and tridentate coordinations with ATP. Our revised model predicts the existence of all these Mg2+-ATP binding modes with high probabilities.

Analysis of the X-ray structures of Mg2+·Protein·ATP complexes reveals three preferred Mg2+-ATP binding modes, βγ,αβ, and αβγ. In the absence of the protein, the revised model predicts two of these binding modes, βγ and αβ with the βγ mode having a much higher probability compared to that of the αβ mode. Since we do not observe the third αβγ mode in the absence of the protein, our model suggests that perhaps the protein electrostatic environment, as well as steric hindrance of the binding site, allows this specific coordination once Mg2+·ATP is bound.

Finally, MD simulations of two different Mg2+·Kinase·ATP complexes show that the revisions that we introduce in Mg2+–protein, Mg2+–nucleotide, and protein–nucleotide interactions substantially improve the correspondence between MD and X-ray structures. We note improvements in both global and local catalytic site structures. In terms of local structures, we note improvements in predictions of Mg2+-ATP binding modes, ATP-protein binding modes, as well as ATP orientations within kinases.

However, there are still discernible differences in local catalytic site geometries between the revised model and X-ray structures. Most notable is the revised model’s inability to retain βγ or αβγ binding modes of Mg2+ to ATP, although it does maintain the αγ binding mode. These modes are also not retained by the original model. However, we note that these modes do not switch to binding modes that are more probable in the aqueous phase. The βγ mode switches to the γ mode, although both the βγ and γ modes are equally likely in the aqueous phase. The tridentate αβγ mode remains tridentate but fluctuates between αγγ and ααγ modes, although the ααγ mode is not observed in the aqueous phase. This implies that the protein is strongly influencing the formation of these modes, and perhaps the errors that are still present in the Mg2+–protein interactions are somehow influencing the switching of these Mg2+-ATP binding modes in the catalytic site. It remains to be seen whether further improvements in Mg2+–protein interactions close this gap between MD simulations and X-ray crystallography.

Supplementary Material

Supporting Information

ACKNOWLEDGMENTS

The authors thank Prof. Sagar Pandit for his insightful feedback. The authors acknowledge the use of computer time from Research Computing at USF and Hungarian HPC Infrastructure at NIIF Institute. Funding for this study was provided by the National Institute of Health (grant numbers R01 GM118697 and R01 GM147210), ERC (starting grant number 101076972), the National Research, Development, and Innovation Office (grant numbers FK142489 and KKP126451), and the János Bolyai Research Scholarship of the Hungarian Academy of Sciences.

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jcim.3c01513.

Statistical analyses of PDBs of Mg2+–ATP–protein and Mg2+–protein complexes; water residence time calculations; Mg2+ parameters in different versions of AMOEBA force fields; structures of Mg2+ complexes with small molecules; new parameters for triphosphate; parameter space explored for fitting dihedral parameters; optimized geometries of triphosphate and Mg2+–DMP–water clusters; comparison of triphosphate–water interactions between AMOEBA and QM; new cross-terms for Mg2+–protein, Mg2+–ATP, and ATP–protein interactions; summary of experimental estimates of hydration free energies of ATP; optimized geometries of 6-fold Mg2+–ATP–water clusters; statistics of Mg2+-ATP binding modes in the aqueous phase; and Mg2+–Mg2+ distance in GSK3β MD simulations (PDF)

Complete contact information is available at: https://pubs.acs.org/10.1021/acs.jcim.3c01513

The authors declare no competing financial interest.

Contributor Information

Julian M. Delgado, Department of Molecular Biosciences, University of South Florida, Tampa, Florida 33620, United States

Péter R. Nagy, Department of Physical Chemistry and Materials Science, Faculty of Chemical Technology and Biotechnology, Budapest University of Technology and Economics, Budapest H-1111, Hungary; HUN-REN-BME Quantum Chemistry Research Group, Budapest H-1111, Hungary; MTA-BME Lendület Quantum Chemistry Research Group, Budapest H-1111, Hungary

Sameer Varma, Department of Molecular Biosciences, University of South Florida, Tampa, Florida 33620, United States; Department of Physics, University of South Florida, Tampa, Florida 33620, United States.

Data Availability Statement

Software packages, Tinker9, FHI-aims, Gaussian9, and Mrcc are available at https://github.com/TinkerTools/tinker9, https://fhi-aims.org, https://gaussian.com, and https://www.mrcc.hu/, respectively. New parameters for the AMOEBA force field are provided in the Supporting Information, and the updated AMOEBA force field PRM file can be downloaded from http://labs.cas.usf.edu/cbb/research.htm. PDB structures used in this study are tabulated in the Supporting Information.

REFERENCES

  • (1).Wolf FI; Torsello A; Fasanella S; Cittadini A Cell physiology of magnesium. Mol. Aspects Med 2003, 24, 11–26. [DOI] [PubMed] [Google Scholar]
  • (2).Romani AM Cellular magnesium homeostasis. Arch. Biochem. Biophys 2011, 512, 1–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Romani AMP Encyclopedia of Metalloproteins; Kretsinger RH, Uversky VN, Permyakov EA, Eds.; Springer New York: New York, NY, 2013; pp 1255–1264. [Google Scholar]
  • (4).Black C; Huang H-W; Cowan J Biological coordination chemistry of magnesium, sodium, and potassium ions. Protein and nucleotide binding sites. Coord. Chem. Rev 1994, 135–136, 165–202. [Google Scholar]
  • (5).Tang S; Yang JJ Encyclopedia of Metalloproteins; Kretsinger RH, Uversky VN, Permyakov EA, Eds.; Springer New York: New York, NY, 2013; pp 1243–1250. [Google Scholar]
  • (6).Hartwig A. Role of magnesium in genomic stability. Mutat. Res 2001, 475, 113–121. [DOI] [PubMed] [Google Scholar]
  • (7).Le CTM; Houri A; Balage N; Smith BJ; Mechler A Interaction of small ionic species with phospholipid membranes: The role of metal coordination. Front. Mater. Sci 2019, 5, 80. [Google Scholar]
  • (8).Dunn MJ; Grant R Red blood cell calcium and magnesium: effects upon sodium and potassium transport and cellular morphology. Biochim. Biophys. Acta, Biomembr 1974, 352, 97–116. [DOI] [PubMed] [Google Scholar]
  • (9).Romero PJ The role of membrane-bound magnesium in the permeability of ghosts to K+. Biochim. Biophys. Acta, Biomembr 1974, 339, 116–125. [DOI] [PubMed] [Google Scholar]
  • (10).Wolf FI; Trapani V Cell (patho) physiology of magnesium. Clin. Sci 2008, 114, 27–35. [DOI] [PubMed] [Google Scholar]
  • (11).Ponder JW; Case DA Protein Simulations. Advances in Protein Chemistry; Academic Press, 2003; Vol. 66, pp 27–85. [DOI] [PubMed] [Google Scholar]
  • (12).Hornak V; Abel R; Okur A; Strockbine B; Roitberg A; Simmerling C Comparison of multiple Amber force fields and development of improved protein backbone parameters. Proteins: Struct., Funct., Bioinf 2006, 65, 712–725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).Cieplak P; Dupradeau F-Y; Duan Y; Wang J Polarization effects in molecular mechanical force fields. J. Condens. Matter Phys 2009, 21, 333102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Cisneros GA; Karttunen M; Ren P; Sagui C Classical Electrostatics for Biomolecular Simulations. Chem. Rev 2014, 114, 779–814. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Baker CM Polarizable force fields for molecular dynamics simulations of biomolecules. Wiley Interdiscip. Rev. Comput. Mol. Sci 2015, 5, 241–254. [Google Scholar]
  • (16).Lopes PEM; Guvench O; MacKerell AD Current Status of Protein Force Fields for Molecular Dynamics Simulations. Methods Mol. Biol 2015, 1215, 47–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (17).Lemkul JA; Huang J; Roux B; MacKerell AD An Empirical Polarizable Force Field Based on the Classical Drude Oscillator Model: Development History and Recent Applications. Chem. Rev 2016, 116, 4983–5013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Nerenberg PS; Head-Gordon T New developments in force fields for biomolecular simulations. Curr. Opin. Struct. Biol 2018, 49, 129–138. [DOI] [PubMed] [Google Scholar]
  • (19).Jing Z; Liu C; Cheng SY; Qi R; Walker BD; Piquemal J-P; Ren P Polarizable Force Fields for Biomolecular Simulations: Recent Advances and Applications. Annu. Rev. Biophys 2019, 48, 371–394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Jiao D; King C; Grossfield A; Darden TA; Ren P Simulation of Ca2+ and Mg2+ solvation using polarizable atomic multipole potential. J. Phys. Chem. B 2006, 110, 18553–18559. [DOI] [PubMed] [Google Scholar]
  • (21).Yu H; Whitfield TW; Harder E; Lamoureux G; Vorobyov I; Anisimov VM; MacKerell AD Jr.; Roux B Simulating monovalent and divalent ions in aqueous solution using a Drude polarizable force field. J. Chem. Theory Comput 2010, 6, 774–786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Li P; Roberts BP; Chakravorty DK; Merz KM Jr. Rational design of particle mesh Ewald compatible Lennard-Jones parameters for+ 2 metal cations in explicit solvent. J. Chem. Theory Comput 2013, 9, 2733–2748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Grotz KK; Cruz-León S; Schwierz N Optimized magnesium force field parameters for biomolecular simulations with accurate solvation, ion-binding, and water-exchange properties. J. Chem. Theory Comput 2021, 17, 2530–2540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (24).Allnér O; Nilsson L; Villa A Magnesium ion–water coordination and exchange in biomolecular simulations. J. Chem. Theory Comput 2012, 8, 1493–1502. [DOI] [PubMed] [Google Scholar]
  • (25).Warren GL; Patel S Comparison of the Solvation Structure of Polarizable and Nonpolarizable Ions in Bulk Water and Near the Aqueous Liquid-Vapor Interface. J. Phys. Chem. C 2008, 112, 7455–7467. [Google Scholar]
  • (26).Varma S; Rempe SB Multibody Effects in Ion Binding and Selectivity. Biophys. J 2010, 99, 3394–3401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).Rogers DM; Beck TL Quasichemical and structural analysis of polarizable anion hydration. J. Chem. Phys 2010, 132, 014505. [DOI] [PubMed] [Google Scholar]
  • (28).Varma S; Rogers DM; Pratt LR; Rempe SB Design principles for K+ selectivity in membrane transport. J. Gen. Physiol 2011, 138, 279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (29).Rossi M; Tkatchenko A; Rempe SB; Varma S Role of methyl-induced polarization in ion binding. Proc. Natl. Acad. Sci. U.S.A 2013, 110, 12978–12983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (30).Savelyev A; MacKerell AD Balancing the Interactions of Ions, Water, and DNA in the Drude Polarizable Force Field. J. Phys. Chem. B 2014, 118, 6742–6757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (31).Li H; Ngo V; Da Silva MC; Salahub DR; Callahan K; Roux B; Noskov SY Representation of Ion-Protein Interactions Using the Drude Polarizable Force-Field. J. Phys. Chem. B 2015, 119, 9401–9416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (32).Savelyev A; MacKerell AD Competition among Li+, Na+, K+, and Rb+ Monovalent Ions for DNA in Molecular Dynamics Simulations Using the Additive CHARMM36 and Drude Polarizable Force Fields. J. Phys. Chem. B 2015, 119, 4428–4440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (33).Wineman-Fisher V; Al-Hamdani Y; Addou I; Tkatchenko A; Varma S Ion-Hydroxyl Interactions: From High-Level Quantum Benchmarks to Transferable Polarizable Force Fields. J. Chem. Theory Comput 2019, 15, 2444–2453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (34).Wineman-Fisher V; Al-Hamdani Y; Nagy PR; Tkatchenko A; Varma S Improved description of ligand polarization enhances transferability of ion–ligand interactions. J. Chem. Phys 2020, 153, 094115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35).Wineman-Fisher V; Delgado JM; Nagy PR; Jakobsson E; Pandit SA; Varma S Transferable interactions of Li+ and Mg2+ ions in polarizable models. J. Chem. Phys 2020, 153, 104113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (36).Delgado JA; Wineman-Fisher V; Pandit S; Varma S Inclusion of High-Field Target Data in AMOEBA’s Calibration Improves Predictions of Protein–Ion Interactions. J. Chem. Inf. Model 2022, 62, 4713–4726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (37).Baker CM; Lopes PEM; Zhu X; Roux B; MacKerell AD Accurate Calculation of Hydration Free Energies using Pair-Specific Lennard-Jones Parameters in the CHARMM Drude Polarizable Force Field. J. Chem. Theory Comput 2010, 6, 1181–1198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (38).Fyta M; Netz RR Ionic force field optimization based on single-ion and ion-pair solvation properties: Going beyond standard mixing rules. J. Chem. Phys 2012, 136 (12), 124103. [DOI] [PubMed] [Google Scholar]
  • (39).Mamatkulov S; Fyta M; Netz RR Force fields for divalent cations based on single-ion and ion-pair properties. J. Chem. Phys 2013, 138 (2), 024505. [DOI] [PubMed] [Google Scholar]
  • (40).Yoo J; Aksimentiev A Improved Parametrization of Li+, Na+, K+, and Mg2+ Ions for All-Atom Molecular Dynamics Simulations of. Nucleic Acid Systems. Chem. Phys. Lett 2012, 3, 45–50. [Google Scholar]
  • (41).Jing Z; Qi R; Liu C; Ren P Study of interactions between metal ions and protein model compounds by energy decomposition analyses and the AMOEBA force field. J. Chem. Phys 2017, 147 (16), 161733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (42).Goldberg RN; Tewari YB Thermodynamics of the disproportionation of adenosine 5’-diphosphate to adenosine 5’-triphosphate and adenosine 5’-monophosphate: I. Equilibrium model. Biophys. Chem 1991, 40, 241–261. [DOI] [PubMed] [Google Scholar]
  • (43).Saris N-EL; Mervaala E; Karppanen H; Khawaja JA; Lewenstam A Magnesium: an update on physiological, clinical and analytical aspects. Clin. Chim. Acta 2000, 294, 1–26. [DOI] [PubMed] [Google Scholar]
  • (44).Cohn M; Hughes TR Jr. Nuclear magnetic resonance spectra of adenosine di-and triphosphate: II. Effect of complexing with divalent metal ions. Biophys. J 1962, 237, 176–181. [PubMed] [Google Scholar]
  • (45).Tran-Dinh S; Neumann JA 31P-NMR study of the interaction of Mg2+ ions with nucleoside diphosphates. Nucleic Acids Res. 1977, 4, 397–403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (46).Ramirez F; Marecek JF Coordination of magnesium with adenosine 5’-diphosphate and triphosphate. Biochim. Biophys. Acta, Bioenerg 1980, 589, 21–29. [DOI] [PubMed] [Google Scholar]
  • (47).Takeuchi H; Murata H; Harada I Interaction of adenosine 5’-triphosphate with Mg2+: vibrational study of coordination sites by use of 18O-labeled triphosphates. J. Am. Chem. Soc 1988, 110, 392–397. [Google Scholar]
  • (48).Kuntz G; Swift T Contrasting structures of magnesium and calcium adenosine-triphosphate complexes as studied by nuclear relaxation. Fed. Proc 1973, 32, 546–552. [Google Scholar]
  • (49).Bishop EO; Kimber SJ; Orchard D; Smith BE A 31P-NMR study of mono-and dimagnesium complexes of adenosine 5’-triphosphate and model systems. Biochim. Biophys. Acta, Bioenerg 1981, 635, 63–72. [DOI] [PubMed] [Google Scholar]
  • (50).Dudev T; Grauffel C; Lim C How native and alien metal cations bind ATP: Implications for lithium as a therapeutic agent. Sci. Rep 2017, 7, 42377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (51).Komuro Y; Re S; Kobayashi C; Muneyuki E; Sugita Y CHARMM force-fields with modified polyphosphate parameters allow stable simulation of the ATP-bound structure of Ca2+-ATPase. J. Chem. Theory Comput 2014, 10, 4133–4142. [DOI] [PubMed] [Google Scholar]
  • (52).Buelens FP; Leonov H; de Groot BL; Grubmüller H ATP–Magnesium Coordination: Protein Structure-Based Force Field Evaluation and Corrections. J. Chem. Theory Comput 2021, 17, 1922–1930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (53).Walker B; Jing Z; Ren P Molecular dynamics free energy simulations of ATP: Mg2+ and ADP: Mg2+ using the polarisable force field AMOEBA. Mol. Simul 2021, 47, 439–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (54).Ren P; Ponder JW Polarizable atomic multipole water model for molecular mechanics simulation. J. Phys. Chem. B 2003, 107, 5933–5947. [Google Scholar]
  • (55).Grossfield A; Ren P; Ponder JW Ion Solvation Thermodynamics from Simulation with a Polarizable Force Field. J. Am. Chem. Soc 2003, 125, 15671–15682. [DOI] [PubMed] [Google Scholar]
  • (56).Ren P; Ponder JW Temperature and pressure dependence of the AMOEBA water model. J. Phys. Chem. B 2004, 108, 13427–13437. [Google Scholar]
  • (57).Ponder JW; Wu C; Ren P; Pande VS; Chodera JD; Schnieders MJ; Haque I; Mobley DL; Lambrecht DS; DiStasio RA Jr.; et al. Current status of the AMOEBA polarizable force field. J. Phys. Chem. B 2010, 114, 2549–2564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Ren P; Wu C; Ponder JW Polarizable Atomic Multipole-Based Molecular Mechanics for Organic Molecules. J. Chem. Theory Comput 2011, 7, 3143–3161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (59).Shi Y; Xia Z; Zhang J; Best R; Wu C; Ponder JW; Ren P Polarizable Atomic Multipole-Based AMOEBA Force Field for Proteins. J. Chem. Theory Comput 2013, 9, 4046–4063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (60).Rahman S; Wineman-Fisher V; Al-Hamdani Y; Tkatchenko A; Varma S Predictive QM/MM Modeling of Modulations in Protein–Protein Binding by Lysine Methylation. J. Mol. Biol 2021, 433, 166745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (61).Rahman S; Wineman-Fisher V; Nagy PR; Al-Hamdani Y; Tkatchenko A; Varma S Methyl-induced Polarization Destabilizes Non-covalent Interactions of N-methylated Lysines. Chem.—Eur. J 2021, 27, 11005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (62).Marcus Y. Ions in Solution and Their Solvation; John Wiley & Sons, 2015. [Google Scholar]
  • (63).Ohtaki H; Radnai T Structure and dynamics of hydrated ions. Chem. Rev 1993, 93, 1157–1204. [Google Scholar]
  • (64).Bleuzen A; Pittet P-A; Helm L; Merbach AE Water exchange on magnesium (II) in aqueous solution: a variable temperature and pressure 17O NMR study. Magn. Reson. Chem 1997, 35, 765–773. [Google Scholar]
  • (65).Perdew JP; Burke K; Ernzerhof M Generalized Gradient Approximation Made Simple. Phys. Rev. Lett 1996, 77, 3865–3868. [DOI] [PubMed] [Google Scholar]
  • (66).Adamo C; Barone V Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys 1999, 110, 6158–6170. [Google Scholar]
  • (67).Tkatchenko A; Scheffler M Accurate molecular van der Waals interactions from ground-state electron density and free-atom reference data. Phys. Rev. Lett 2009, 102, 073005. [DOI] [PubMed] [Google Scholar]
  • (68).Blum V; Gehrke R; Hanke F; Havu P; Havu V; Ren X; Reuter K; Scheffler M Ab initio molecular simulations with numeric atom-centered orbitals. Comput. Phys. Commun 2009, 180, 2175–2196. [Google Scholar]
  • (69).DiLabio GA; Otero-de-la Roza A Noncovalent interactions in density functional theory. Rev. Comput. Chem 2016, 29, 1–97. [DOI] [PubMed] [Google Scholar]
  • (70).Marom N; Tkatchenko A; Rossi M; Gobre VV; Hod O; Scheffler M; Kronik L Dispersion Interactions with Density-Functional Theory: Benchmarking Semiempirical and Interatomic Pairwise Corrected Density Functionals. J. Chem. Theory Comput 2011, 7, 3944–3951. [DOI] [PubMed] [Google Scholar]
  • (71).Nagy PR; Kállay M Optimization of the linear-scaling local natural orbital CCSD(T) method: Redundancy-free triples correction using Laplace transform. J. Chem. Phys 2017, 146, 214106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (72).Nagy PR; Samu G; Kallay M Optimization of the linear-scaling local natural orbital CCSD(T) method: Improved algorithm and benchmark applications. J. Chem. Theory Comput 2018, 14, 4193–4215. [DOI] [PubMed] [Google Scholar]
  • (73).Nagy PR; Kállay M Approaching the basis set limit of CCSD(T) energies for large molecules with local natural orbital coupled-cluster methods. J. Chem. Theory Comput 2019, 15, 5275–5298. [DOI] [PubMed] [Google Scholar]
  • (74).Gyevi-Nagy L; Kállay M; Nagy PR Integral-direct and parallel implementation of the CCSD(T) method: Algorithmic developments and large-scale applications. J. Chem. Theory Comput 2020, 16, 336. [DOI] [PubMed] [Google Scholar]
  • (75).Kallay M; Nagy PR; Mester D; Rolik Z; Samu G; Csontos J; Csóka J; Szabó PB; Gyevi-Nagy L; Hégely B; et al. The MRCC program system: Accurate quantum chemistry from water to proteins. J. Chem. Phys 2020, 152, 074107. [DOI] [PubMed] [Google Scholar]
  • (76).Kállay M. et al. Mrcc, a quantum chemical program suite. See https://www.mrcc.hu/ (accessed August 1, 2023).
  • (77).Prascher BP; Woon DE; Peterson KA; Dunning TH; Wilson AK Gaussian basis sets for use in correlated molecular calculations. VII. Valence, core-valence, and scalar relativistic basis sets for Li, Be, Na, and Mg. Theor. Chem. Acc 2011, 128, 69–82. [Google Scholar]
  • (78).Karton A; Martin JML Comment on: “Estimating the Hartree–Fock limit from finite basis set calculations” [Jensen F (2005) Theor Chem Acc 113:267]. Theor. Chem. Acc 2006, 115, 330–333. [Google Scholar]
  • (79).Helgaker T; Klopper W; Koch H; Noga J Basis-set convergence of correlated calculations on water. J. Chem. Phys 1997, 106, 9639–9646. [Google Scholar]
  • (80).Møller C; Plesset MS Note on an approximation treatment for many-electron systems. Phys. Rev 1934, 46, 618–622. [Google Scholar]
  • (81).Frisch MJ; et al. Gaussian 09. Revision E.01; Gaussian Inc.: Wallingford CT, 2009;. [Google Scholar]
  • (82).Dunning TH Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys 1989, 90, 1007–1023. [Google Scholar]
  • (83).Hickey AL; Rowley CN Benchmarking quantum chemical methods for the calculation of molecular dipole moments and polarizabilities. J. Phys. Chem. A 2014, 118, 3678–3687. [DOI] [PubMed] [Google Scholar]
  • (84).Wang Z; Ponder JW Tinker9: Next Generation of Tinker with GPU Support, 2021. https://github.com/TinkerTools/tinker9.
  • (85).Bussi G; Donadio D; Parrinello M Canonical sampling through velocity rescaling. J. Chem. Phys 2007, 126, 014101. [DOI] [PubMed] [Google Scholar]
  • (86).Åqvist J; Wennerström P; Nervall M; Bjelic S; Brandsdal BO Molecular dynamics simulations of water and biomolecules with a Monte Carlo constant pressure algorithm. Chem. Phys. Lett 2004, 384, 288–294. [Google Scholar]
  • (87).Chow K-H; Ferguson DM Isothermal-isobaric molecular dynamics simulations with Monte Carlo volume sampling. Comput. Phys. Commun 1995, 91, 283–289. [Google Scholar]
  • (88).Tuckerman M; Berne BJ; Martyna GJ Reversible multiple time scale molecular dynamics. J. Chem. Phys 1992, 97, 1990–2001. [Google Scholar]
  • (89).Allen MP; Tildesley DJ Computer Simulation of Liquids; Oxford University Press, 2017. [Google Scholar]
  • (90).Rackers JA; Wang Z; Lu C; Laury ML; Lagardère L; Schnieders MJ; Piquemal J-P; Ren P; Ponder JW Tinker 8: software tools for molecular design. J. Chem. Theory Comput 2018, 14, 5273–5289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (91).Bertrand J; Thieffine S; Vulpetti A; Cristiani C; Valsasina B; Knapp S; Kalisz H; Flocco M Structural characterization of the GSK-3β active site using selective and non-selective ATP-mimetic inhibitors. J. Mol. Biol 2003, 333, 393–407. [DOI] [PubMed] [Google Scholar]
  • (92).Brown NR; Noble ME; Lawrie AM; Morris MC; Tunnah P; Divita G; Johnson LN; Endicott JA Effects of phosphorylation of threonine 160 on cyclin-dependent kinase 2 structure and activity. J. Biol. Chem 1999, 274, 8746–8756. [DOI] [PubMed] [Google Scholar]
  • (93).šali A; Blundell TL Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol 1993, 234, 779–815. [DOI] [PubMed] [Google Scholar]
  • (94).Dolinsky TJ; Nielsen JE; McCammon JA; Baker NA PDB2PQR: an automated pipeline for the setup of Poisson–Boltzmann electrostatics calculations. Nucleic Acids Res. 2004, 32, W665–W667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (95).Gilson MK; Given JA; Bush BL; McCammon JA The Statistical-Thermodynamic Basis for Computation of Binding Affinities: A Critical Review. Biophys. J 1997, 72, 1047–1069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (96).Hamelberg D; McCammon JA Standard free energy of releasing a localized water molecule from the binding pockets of proteins: double-decoupling method. J. Am. Chem. Soc 2004, 126, 7683–7689. [DOI] [PubMed] [Google Scholar]
  • (97).Bennett CH Efficient estimation of free energy differences from Monte Carlo data. J. Comput. Phys 1976, 22, 245–268. [Google Scholar]
  • (98).Fogarty JC; Chiu S-W; Kirby P; Jakobsson E; Pandit SA Automated optimization of water-water interaction parameters for a coarse-grained model. J. Phys. Chem. B 2014, 118, 1603–1611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (99).Saunders M; Wineman-Fisher V; Jakobsson E; Varma S; Pandit SA High-Dimensional Parameter Search Method to Determine Force Field Mixing Terms in Molecular Simulations. Langmuir 2022, 38, 2840–2851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (100).Varma S; Rempe S Structural transitions in ion coordination driven by changes in competition for ligand binding. J. Am. Chem. Soc 2008, 130, 15405–15419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (101).Dudev T; Lin; Dudev M; Lim C First- second shell interactions in metal binding sites in proteins: A PDB survey and DFT/CDM calculations. J. Am. Chem. Soc 2003, 125, 3168–3180. [DOI] [PubMed] [Google Scholar]
  • (102).Liu C; Qi R; Wang Q; Piquemal J-P; Ren P Capturing many-body interactions with classical dipole induction models. J. Chem. Theory Comput 2017, 13, 2751–2761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (103).Foloppe N; MacKerell AD Jr. All-atom empirical force field for nucleic acids: I. Parameter optimization based on small molecule and condensed phase macromolecular target data. J. Comput. Chem 2000, 21, 86–104. [Google Scholar]
  • (104).Pastor R; MacKerell A Jr. Development of the CHARMM force field for lipids. J. Phys. Chem. Lett 2011, 2, 1526–1532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (105).Cornell WD; Cieplak P; Bayly CI; Gould IR; Merz KM; Ferguson DM; Spellmeyer DC; Fox T; Caldwell JW; Kollman PA A second generation force field for the simulation of proteins, nucleic acids, and organic molecules. J. Am. Chem. Soc 1995, 117, 5179–5197. [Google Scholar]
  • (106).Zhang C; Lu C; Wang Q; Ponder JW; Ren P Polarizable multipole-based force field for dimethyl and trimethyl phosphate. J. Chem. Theory Comput 2015, 11, 5326–5339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (107).Bossemeyer D. The glycine-rich sequence of protein kinases: a multifunctional element. Trends Biochem. Sci 1994, 19, 201–205. [DOI] [PubMed] [Google Scholar]
  • (108).Delgado JM; Duro N; Rogers DM; Tkatchenko A; Pandit S; Varma S Molecular basis for higher affinity of SARS-CoV-2 spike RBD for human ACE2 receptor. Proteins: Struct., Funct., Bioinf 2021, 89, 1134–1144. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supporting Information

Data Availability Statement

Software packages, Tinker9, FHI-aims, Gaussian9, and Mrcc are available at https://github.com/TinkerTools/tinker9, https://fhi-aims.org, https://gaussian.com, and https://www.mrcc.hu/, respectively. New parameters for the AMOEBA force field are provided in the Supporting Information, and the updated AMOEBA force field PRM file can be downloaded from http://labs.cas.usf.edu/cbb/research.htm. PDB structures used in this study are tabulated in the Supporting Information.

RESOURCES