ACS Publications. Most Trusted. Most Cited. Most Read
Relative Binding Free Energy Calculations for Ligands with Diverse Scaffolds with the Alchemical Transfer Method
My Activity

Figure 1Loading Img
  • Open Access
Computational Chemistry

Relative Binding Free Energy Calculations for Ligands with Diverse Scaffolds with the Alchemical Transfer Method
Click to copy article linkArticle link copied!

  • Solmaz Azimi
    Solmaz Azimi
    Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United States
    Ph.D. Program in Biochemistry, The Graduate Center of the City University of New York, New York, New York 10016, United States
    More by Solmaz Azimi
  • Sheenam Khuttan
    Sheenam Khuttan
    Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United States
    Ph.D. Program in Biochemistry, The Graduate Center of the City University of New York, New York, New York 10016, United States
  • Joe Z. Wu
    Joe Z. Wu
    Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United States
    Ph.D. Program in Chemistry, The Graduate Center of the City University of New York, New York, New York 10016, United States
    More by Joe Z. Wu
  • Rajat K. Pal
    Rajat K. Pal
    Roivant Sciences, Inc., Boston, Massachusetts 02210, United States
    More by Rajat K. Pal
  • Emilio Gallicchio*
    Emilio Gallicchio
    Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United States
    Ph.D. Program in Chemistry, The Graduate Center of the City University of New York, New York, New York 10016, United States
    Ph.D. Program in Biochemistry, The Graduate Center of the City University of New York, New York, New York 10016, United States
    *E-mail: [email protected]
Open PDFSupporting Information (1)

Journal of Chemical Information and Modeling

Cite this: J. Chem. Inf. Model. 2022, 62, 2, 309–323
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jcim.1c01129
Published January 6, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

We present an extension of the alchemical transfer method (ATM) for the estimation of relative binding free energies of molecular complexes applicable to conventional, as well as scaffold-hopping, alchemical transformations. Named ATM-RBFE, the method is implemented in the free and open-source OpenMM molecular simulation package and aims to provide a simpler and more generally applicable route to the calculation of relative binding free energies than what is currently available. ATM-RBFE is based on sound statistical mechanics theory and a novel coordinate perturbation scheme designed to swap the positions of a pair of ligands such that one is transferred from the bulk solvent to the receptor binding site while the other moves simultaneously in the opposite direction. The calculation is conducted directly in a single solvent box with a system prepared with conventional setup tools, without splitting of electrostatic and nonelectrostatic transformations, and without pairwise soft-core potentials. ATM-RBFE is validated here against the absolute binding free energies of the SAMPL8 GDCC host–guest benchmark set and against protein–ligand benchmark sets that include complexes of the estrogen receptor ERα and those of the methyltransferase EZH2. In each case the method yields self-consistent and converged relative binding free energy estimates in agreement with absolute binding free energies and reference literature values, as well as experimental measurements.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2022 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Relative binding free energy (RBFE) calculations are a critical component of modern structure-based drug discovery efforts. (1−3) Generally, the goal of these calculations is to estimate the difference of the standard binding free energies of two molecular ligands to the same receptor more directly than taking the difference of the corresponding absolute binding free energies (ABFEs). While a variety of approaches to obtain RBFEs have been proposed, most large-scale deployments to date employ a thermodynamic cycle in which one ligand is alchemically transformed into another ligand while both are bound to the receptor and, separately, while both are in the solvent bulk. The difference of the free energy changes of these two transformations yields the RBFE of the ligand pair. (4−6)
In RBFE methods, the transformations from one ligand into another are termed alchemical as they are based on progressively modifying the Hamiltonian of the system through a series of artificial chemical states that interpolate between the physical end states corresponding to the two complexes. These alchemical transformations are commonly carried out by parameter interpolation, in which the parameters of the energy function (partial charges, Lennard-Jones parameters, force constants, soft-core parameters, etc.) are progressively changed by means of an alchemical progress parameter λ. (7,8) Generally, the implementation of parameter interpolation methods require specific customization of the energy routines of the molecular dynamics engine, especially when soft-core pairwise potentials are used to treat end-point singularities. (9,10)
Various implementations of alchemical transformations are commonly utilized. (11,12) For one, single-topology implementations map a selected set of atoms of the first ligand to corresponding atoms of the second ligand. The first set of atoms is then transformed into the other by interpolating the force field parameters assigned to these atoms. (13) In the case of two ligands with different atom counts, the atoms of one ligand that do not correspond in the other ligand are converted to and from dummy atoms. (14) In dual-topology implementations by the means of congeneric pairing, a whole functional group of one ligand is typically converted to that of the second ligand by converting the first group to a dummy group and converting the second from a dummy group to the target group simultaneously. (7)
Single- and dual-topology approaches have distinct advantages and limitations that make them more or less preferable depending on the specific application. The correct treatment of dummy atoms in single-topology methods can be particularly subtle. (14) In dual-topology approaches, the mode of attachment of the dummy group to the ligand scaffold can also require care in order to avoid artifacts in the resulting free energies. (15,16) These and other implementation issues can be quite complex to address especially when considered in the context of hybrid approaches that involve both single- and dual-topology transformations, (17) as well as the more specialized separated-topology approaches. (18) In all of these methods, specialized system setup tools are employed to deal with the nonchemical molecular topologies to treat, for example, dummy atoms and provide ways to make sure that the atoms of the groups being transformed in dual topologies do not interact with each other. To avoid numerical instabilities, the alchemical transformations are also typically split into distinct simulation legs to separately treat van der Waals and electrostatic interactions. (5,8)
The setup of an RBFE calculation is further complicated by the preparation and simulation of multiple systems, particularly if the calculation is based on two simulation systems─the receptor complex (with the two ligands) in the solvent and, separately, the ligands in the solvent. RBFE implementations based on separate hydration and receptor-binding systems, for example, are not naturally suited for ligand pairs with different net charges. (19) In these cases, the change of system charge during the alchemical transformation causes issues with the treatment of long-range electrostatic interactions and introduces artifacts in the free energy estimates unless complex correction factors are introduced. (20) To overcome some of these challenges, coupled transformations have also been used in which the change in the net charge of the ligands is counterbalanced by the opposite change of charge of an ion to a solvent molecule or to another ion. (21,22)
Scaffold hopping transformations that, unlike the simpler terminal R-group modifications, include structural changes in the chemical topologies of the two ligands, such as breaking and forming rings, reducing the size of rings, and modifying linker groups, are particularly challenging with traditional RBFE approaches. Wang. et al. (23) introduced a customized soft-bond stretch potential to address the numerical instabilities encountered when forming and breaking covalent bonds in a single-topology context. More recently, Zou et al. (24) presented a method that uses a series of restraining potentials to address the same problem.
As a result of the above-discussed issues and other complexities associated with RBFE methods (25) and despite the strong need in structure-based drug discovery, software tools in this arena remain of limited availability to many practitioners in the field. This deficiency is particularly felt in academic settings where commercial solutions might be prohibitively expensive or not sufficiently customizable, and the substantial amount of human resources and the specialized expertise required to develop and deploy RBFE solutions anew are not available. As an example, there are only a handful of open-source academic codes, at various stages of development, potentially applicable to scaffold-hopping RBFE transformations. (26−28)
In an attempt to address some of the aforementioned challenges, in this work we present a streamlined RBFE protocol based on the alchemical transfer method (ATM). (29) The method aims to provide a simpler and more generally applicable route to relative binding free energy calculations than what is currently available. Like the parent ATM absolute binding free energy (ATM-ABFE) approach and unlike conventional parameter interpolation-based methods, the ATM-RBFE protocol is based on a straightforward coordinate transformation perturbation. As described here for this method, and similar to the double-system/single-box method (30,31) for ABFE calculations, the receptor and the two ligands are initially placed in a single solvated simulation box in such a way that one ligand is bound to the receptor and the other is placed in an arbitrary position in the solvent bulk. Molecular dynamics simulations are then conducted with a λ-dependent alchemical potential energy function based on swapping the positions of the two ligands.
The ATM-RBFE protocol proposed here has the following notable features.
  • It is applicable to simple terminal R-group transformations, as well as to scaffold-hopping transformations to connect ligand pairs that do not share the same topology.

  • It does not require splitting the alchemical transformations into electrostatic and nonelectrostatic steps, and it does not require the implementation of soft-core pair potentials.

  • It does not require modifications of the core energy routines of the molecular dynamics engine. It uses only the energies and forces returned by the unmodified existing routines used for molecular dynamics time propagation.

  • By design, it is applicable with any potential energy function, including conventional fixed-charge molecular mechanics force fields with and without long-range electrostatics, as well as many-body potentials, such as polarizable, (32,33) quantum-mechanical, (34−36) machine learning potentials, (37,38) implicit solvation models, (39) and coarse-grained potentials, (40) which are generally poorly supported by free energy alchemical protocols.

  • The molecular simulation system contains only chemically valid topologies without dummy atoms and is prepared with conventional tools.

  • Because perturbation energies are λ-independent, the alchemical potential energies can be evaluated algebraically rather than by rescoring trajectories through calling the energy routines of the MD engine, making the method easy to implement in conjunction with advanced conformational sampling algorithms, such as Hamiltonian replica exchange, (41) and with multistate free energy analysis tools. (42,43)

  • It is amenable to simple, compact, and self-contained software implementations. The method presented here is implemented into a freely available and open-source plugin of the popular OpenMM library for molecular simulations. (44)

The work is organized as follows. We first describe the formulation of the ABFE and RBFE ATM protocols. A rigorous derivation of the method from a well-established statistical mechanics theory of noncovalent bimolecular binding is presented in the Supporting Information. Here we also introduce a restraining potential that maintains ligand alignment in order to enhance the convergence of the calculations for similar ligand pairs and we show mathematically that these calculations yield unbiased binding free energy estimates. We then present the application of the ATM-RBFE method to a set of host–guest as well as protein–ligand complexes. Specifically, we compute all pairwise relative binding free energies for the SAMPL8 GDCC host–guest sets and for proposed agonists of the estrogen receptor α nuclear receptor (ERα), as well as the inhibitors of the lysine N-methyltransferase 6 enzyme (EZH2). Even though a majority of the ligand pairs considered in this work are related by challenging scaffold-hopping transformations, we show that the ATM-RBFE protocol yields relative binding free estimates that are not only self-consistent and in good agreement with absolute binding free energy estimates but are also consistent with literature reference values and with experimental measurements. This work paves the way to streamlined and more readily available RBFE estimation tools in structure-based drug discovery and other fields.

Theory and Methods

Click to copy section linkSection link copied!

Alchemical Transfer Method for the Estimation of Absolute Binding Free Energies

The alchemical transfer method (ATM for short) is a computational protocol first introduced for the estimation of absolute binding free energies (ABFEs). (29) The ATM-ABFE protocol has been tested and validated on strict host–guest benchmarks. (29,45) The theoretical underpinnings of the approach are described in detail in refs (46), (47), and (29). The derivation of the ATM-ABFE method from statistical mechanics principles is also reviewed in the Supporting Information. (16,29,48−51) The central feature of the method (see Figure 1) is an alchemical perturbation based on a coordinate transformation that translates the position of ligand A, for instance, from the solvent bulk to the binding site of the receptor R. The potential energy functions of the system before and after the translation, UR+A and URA, respectively, are combined into a λ-dependent potential function such that the system is progressively transformed, in an alchemical sense, from the unbound state to the bound state along the alchemical pathway. The ATM-ABFE alchemical pathway does not connect the bound and unbound states directly. Rather, the bound and unbound states are each transformed into a common alchemical intermediate state at the midpoint of the transformation. The absolute binding free energy is then obtained by the difference of the free energy changes of these two transformations. (29) This feature of the method is maintained in the RBFE extension described below (see also Figure 2).

Figure 1

Figure 1. Illustration of the ATM-RBFE calculation setup that consists of displacing one ligand from the binding site to the solvent bulk along a translation vector d (in green) while simultaneously translating the second ligand from the solvent bulk to the binding site along the same vector. The protein receptor (ERα) is shown in cartoon representation colored by secondary structure. The ligands are shown in van der Waals representation.

Figure 2

Figure 2. Free energy diagram for an ATM-RBFE calculation, which consists of two independent legs that are connected to a single alchemical intermediate state. The alchemical calculation for leg 1 begins at λ = 0, in which ligand A is bound to the binding site of the receptor R and ligand B is dissociated in the solvent bulk, and ends at λ = 1/2 at the alchemical intermediate (denoted by R(AB)1/2 + (BA)1/2), in which A and B are simultaneously present at 50% strength in the binding site and solvent bulk. The alchemical calculation for leg 2 begins with ligand B bound to the binding site and ligand A in the solvent bulk and ends at same alchemical intermediate. ΔG1 and ΔG2 correspond to the free energy changes along each respective ATM leg. The relative binding free energy, ΔΔGb°(B,A), of ligand B with respect to ligand A is the difference between the free energies of legs 1 and 2.

The ATM-ABFE method introduces several advantageous features in comparison to conventional alchemical binding free energy methods. Among others, it requires only one simulation system free of alchemical topologies and is prepared with conventional tools. It does not require soft-core pair potentials and does not require modifications of the energy routines of the MD engine. It does not require splitting the binding free energy calculations in receptor and solvation legs and into electrostatic and nonelectrostatic steps. In this work, the method is extended to the estimation of relative binding free energies while retaining all of these favorable features.

Alchemical Transfer Method for the Estimation of Relative Binding Free Energies

The ATM relative binding free energy protocol (ATM-RBFE for short) is conceptually similar to the ATM-ABFE protocol described above. While in ATM-ABFE, a single ligand A is translated from the solvent bulk to the binding site, ATM-RBFE considers two ligands A and B, at approximately distance d from each other, one placed in the binding site of the receptor R and the other placed in the solvent bulk, plus a coordinate transformation that swaps their positions (Figure 1).
The rigorous derivation of the method from statistical mechanics principle is outlined in the Supporting Information. The main outcome is an ensemble average expression for the ratio of the binding constants of ligand B relative to ligand A for the receptor R:
(1)
where
(2)
is the ATM-RBFE perturbation energy (see below), RA + B indicates the state in which ligand A is bound to the receptor and ligand B is in the solvent bulk, and RB + A denotes the state obtained by swapping their positions (see Figure 1). The potential energies of these states are
(3)
and
(4)
respectively. In these expressions, xA, xB, and xR denote the internal coordinates of the two ligands and the receptor, respectively, and rS are the coordinates of the solvent molecules. cA and ωA denote the external degrees of freedom of ligand A relative to the receptor or, in other words, the position and orientation of ligand A relative to the coordinate frame of the receptor. The quantity cA denotes a position within the binding site of the receptor. Similar quantities refer to ligand B except that cB* denotes the initial position of ligand B in the solvent bulk (see the Supporting Information).
The translation vector d encodes the coordinate transformation that is at the basis of the method (see Figure 1). The state RB + A is obtained from the state RA + B by displacing ligand A so that, after the translation, it is at position cA + d in the solvent bulk and, concomitantly, ligand B, starting at position cB* in the solvent bulk, is translated in the opposite direction to reach the position cB = cB*d in the receptor binding site. The perturbation energy (eq 2) is the change in potential energy upon this coordinate transformation. Finally, the ensemble average in eq 1 is conducted in the RA + B state, that is in the equilibrium ensemble of conformations obtained when ligand A is bound to the receptor and ligand B is in the solvent bulk.
Equation 1, or its equivalent formulation in terms of the relative binding free energies
(5)
is the formal expression of the ATM-RBFE method. While formally exact, the direct evaluation of the ensemble average in eq 5 is numerically unstable. The following sections describe the alchemical stratification techniques, the ligand alignment restraints, and the thermodynamic cycle that implement eq 5 in practice.

Ligand Alignment Restraints

While eq 5 is formally exact, it converges slowly near the RA + B end-point because translating ligand B in the binding site from the solvent bulk and translating ligand A in the solvent bulk from the binding site starting from configurations sampled from the RA + B state is likely to produce severe clashes with either receptor atoms or solvent molecules. For ligands with similar shape, the occurrences of severe clashes with nearby atoms would be considerably reduced if the two ligands land in a similar location and orientation when their positions are swapped. In order to improve convergence, we therefore introduce restraint potentials that help to maintain the positions and orientations of the two ligands in approximate alignment. The potentials are based on the distance and angles of coordinate reference frames fixed on each ligand defined by three chosen reference atoms (Figure S1). In the Supporting Information we show that the ligand alignment potentials do not bias the relative binding free energy estimates (5) as long as they are symmetric with respect to the exchange of ligand labels.
The ligand alignment restraint is implemented as the sum of position and orientation terms. The positional restraint is a 3D harmonic potential between the origins of the reference frames of ligand A and B (Figure S1) centered on the translation vector d
(6)
As defined, the positional restraint keeps the separation between the two ligands near the translation vector d so that they land in similar positions when they are translated in and out of the binding site (Figure 1). As shown in the Supporting Information, the potential function (eq 6) satisfies the required symmetry properties.
The orientation restraints are implemented by two potentials. One potential, named Uθ, is based on the angle θ between the z-axis of the reference frame of ligand B relative to that of ligand A (Figure S1), which is naturally symmetric with respect to the exchange of the two frames. The second restraint potential is based on the dihedral angle ψ in Figure S1. Because the ψ angle is not symmetric with respect to the exchange of ligand labels, we use a symmetrized restraining potential of the form
(7)
where Uψ is the restraining potential below, ψBA and ψAB are the ψ angles of the reference frame of ligand B with respect to A and of ligand A with respect to B, respectively (Figure S1). Each orientation restraint potential is based on the cosine of the corresponding angle. For example,
(8)
and similarly for the ψ angle, where kθ (or kψ in the case of the ψ angle) is a factor that controls the strength of the restraint.
The reference atoms and values of the force constants used in this work are described in Computational Details for each system considered. Preliminary guidelines on how to choose the ligand reference atoms are discussed later. The effects of dialing in and out the alignment potentials have been studied numerically for one host–guest system (see Results).

Computation Protocol and Software Implementation

While eq 5 is formally exact, its direct implementation in terms of the exponential average of the perturbation energy is numerically unstable. (16) Instead, the free energy change is computed by well-established protocols based on stratification techniques using a sequence of alchemical states defined by a λ-dependent Hamiltonian (52) and multistate thermodynamic reweighting methods. (42,43)
The specific alchemical protocol we employ is described in detail in previous publications. (29,46,47) Briefly, to smoothly interpolate between the RA + B and RB + A states, we employ a linear λ-dependent alchemical potential energy function of the form
(9)
where x represents the set of atomic coordinates of the system (the internal coordinates of the ligands and the receptor, the external coordinates of the ligands, and the coordinates of the solvent), URA+B(x) represents the potential energy of the system before the application of the coordinate transformations (eq 3), the perturbation energy u(x) is defined by eq 2 and Wλ(u) is the generalized soft plus alchemical potential function
(10)
The parameters λ2, λ1, α, u0, and w0 are functions of λ formulated, as described in detail in refs (46) and (47), to soften entropic bottlenecks along the alchemical thermodynamic path and to enhance the rate of binding/unbinding transitions. (46,47) The function
(11)
with
(12)
and
(13)
is the soft-core perturbation energy function designed to avoid singularities near the initial state of the alchemical transformation. (46) The parameters umax, uc, and a are set to cap the perturbation energy u(x) to a maximum positive value without affecting it away from the singularity. The specific values of u0, umax, and of the scaling parameter a used in this work are listed in the Computational Details.
The standard linear alchemical interpolation potential
(14)
is a special case of eq 10 for λ1 = λ2 = λ. With two exceptions (see Computational Details), the RBFE calculations reported in this work employed the linear alchemical potential.
Similar to the ABFE protocol, (29) the RBFE alchemical protocol involves two free energy calculations, defined as two distinct legs (Figure 2). The initial state of the first leg corresponds to the physical state, corresponding to the potential energy function URA+B, in which ligand A is bound to the receptor and ligand B is in the solvent bulk. The initial state of the second leg, described by the potential energy function URB+A, is the physical state in which the roles of the two ligands are reversed so that ligand B is bound to the receptor and ligand A is in the solvent bulk. The alchemical potential energy function for the second leg is the same as the one for leg 1 (eq 9) but with the A and B labels reversed. The initial states of the two legs are connected to the alchemical intermediate state defined by the potential function in eq 9 at λ = 1/2. As discussed in ref (29), the alchemical intermediate state corresponds to a nonphysical symmetric mixture of the two end-states corresponding to the potential energy function
(15)
that describes a state in which the ligands, while not interacting with each other, are interacting at 50% strength with the binding site and solution environments. The difference of the free energy changes, ΔG1 and ΔG2, of leg 1 and leg 2 respectively, gives the relative binding free energy of the two ligands (Figure 2), or in other words:
(16)
In practice, the RBFE alchemical calculation for leg 1, for example, consists of a series of molecular dynamics simulations with the potential function (9) distributed at increasing values of λ varying from 0 to 1/2 corresponding to ligand A bound and ligand B unbound and to the alchemical intermediate, respectively. At each MD step, the perturbation energy (2) is obtained by displacing ligand A into the solvent bulk by translating its position along the vector d while simultaneously translating ligand B to the receptor binding site in the opposite direction.
The calculation procedure for each free energy leg is implemented in the OpenMM molecular dynamics engine (see Figure 3) by means of an add-on plugin. (53) The plugin includes a modified Langevin MD integrator that calls OpenMM’s energy/forces calculation routine before and after swapping the positions of the two ligands. The energies and forces computed by each call are then merged according to eq 9 to obtain the forces at the current value of λ. For example, denoting by FRA+B and FRB+A, the forces before and after the swap respectively, the atomic forces at λ in the case of the linear alchemical potential (14) are
(17)
The values of the perturbation energies are saved at regular intervals and processed by conventional multistate reweighting analysis (43) to compute the free energy of each leg. The calculation for leg 2 is performed similarly to leg 1 by swapping the identities of ligands A and B.

Figure 3

Figure 3. Schematic flow diagram of the MD software implementation of the ATM-RBFE method. At each time step, a modified MD integrator routine gives the current coordinates of the system, in which ligand A is bound to the receptor and ligand B is in the solvent bulk, to the MD engine energy/forces calculation routine (RA + B state, left side of the diagram). The coordinates of the system are then transformed to swap the positions of ligands A and B so that B becomes bound to the receptor and A is now present in the solvent bulk (RB + A state, on the right branch of the diagram) and are given to the energy/force calculation routine. The resulting sets of energies and forces are merged according to eq 9 by the energy/force merging routine and fed again to the MD integrator to initiate the subsequent time step. The blue-colored energy/forces calculations routines are used from the MD engine unmodified. The red-colored routines are customized for ATM.

Molecular Systems

We report here the results of validation tests of the ATM-RBFE protocol on the SAMPL8 GDCC host–guest set (45,54) and on two benchmark sets of protein–ligand complexes that include the nuclear transcription factor estrogen receptor α (ERα) and the methyltransferase enhancer of zeste homologue 2 (EZH2) enzyme. (23,55)
The SAMPL8 GDCC host–guest set consists of five small molecule guests binding to two host receptors (Figure 4). The hosts and the guests are known to form complexes with 1:1 stoichiometry under experimental conditions. The GDCC hosts (named TEMOA and TEETOA) are cup-shaped molecules with deep hydrophobic cavities that accommodate the nonpolar end of each guest. The hosts differ in the nature of the side chains that surround the rim of their respective cavities (methyl groups in TEMOA and ethyl groups in TEETOA). In the host–guest complexes, the nonpolar end of the guest occupies the cavity of the host and the charged/polar groups of the guest tend to orient toward the solvent. In this work, all pairwise relative binding free energies for this set were compared to the differences of the ATM absolute binding free energy (ABFE) estimates that were obtained previously. (45,56) The ABFE estimates for this set are considered particularly trustworthy references because they closely agree with independent potential of mean force estimates and with the experimentally measured binding free energies. (45) As shown in Figure 4, the majority of the pairwise RBFE calculations in this set involve scaffold hopping transformations. For example, all of the transformations to or from the G3 guest, among others, involve the deletion of a ring or the conversion of an aromatic 6-membered ring to an aliphatic 5-membered ring. With the exception of G2, all guests are modeled in their anionic, deprotonated forms. The G2 guest is modeled in the neutral, protonated state that, according to our previous work, (45) contributes the most to binding to both TEMOA and TEETOA hosts. Because all of the ligand pair transformations involving G2 also involve a change of the net charge of the ligand, this set tests the ability of the ATM-RBFE protocol to handle changes in net charge in addition to changes of ligand topology.

Figure 4

Figure 4. ATM binding free energy cycle for the SAMPL8 GDCC benchmark set that includes two hosts, TEMOA (A) and TEETOA (B). A representative complex of each host bound to the guest G1 is shown at the bottom of each panel. Binding free energy estimates in kcal/mol are illustrated alongside arrows connecting each ligand pair transformation (top of each panel). Relative binding free energy estimates are represented in blue, and the difference of the absolute binding free energy for each guest pair are represented in red. These values are also tabulated in Table 1. On the host and guest structures, red corresponds to oxygen atoms and white to hydrogen atoms. Carbon atoms are represented in gray in the host structures and green in the guest structures.

The ERα benchmark set consists of four molecular complexes of the ERα receptor with a series of benzopyran agonists (55) developed for the treatment of some forms of cancer. The four ligands (named 2e, 2d, 3a, and 3b) bind in a similar fashion to a large solvent-occluded cavity of the receptor (Figure 5). The ligands share a similar topology consisting of a common scaffold along the long molecular axis ending in two polar phenol groups, flanked by a variable side group with either a five- or six-membered ring. Alchemical RBFE transformations in this set can be distinctly classified as either straightforward terminal R-group modifications (such as 3a to 3b) or more challenging scaffold-hopping transformations involving ring expansion (such as 2e to 3a). The common long molecular axis and the conserved portion of the side group provide a natural definition of the reference frame to the four ligands (Figure 5.) In this work, all ERα ligands were modeled in their neutral forms.

Figure 5

Figure 5. Relative binding free energy calculations for the ERα complexes. A representative complex of ERα bound to a ligand is demonstrated in the top left. The alignment frame used to apply a restraining potential to the positions and orientations of each ligand pair is illustrated in the bottom left. Relative free energy calculations for each ligand transformation are presented by arrows connecting each ligand pair (right). Free energy estimates in kcal/mol are color-coordinated according to the method: those computed by ATM-RBFE are in black, those obtained experimentally are in red, and those reported in literature are in blue. The same values are reported in Table 2. In the ligand structures, green represents carbon atoms; red, oxygen; and cyan, fluorine.

The EZH2 benchmark set consists of four molecular complexes of the methyltransferase EZH2 enzyme, an established cancer target, with indazole inhibitors (compounds 22 and 27 in Figure 6) and their more potent acyclic alkylated amine derivatives (compounds 29 and 31). (57) Transformations involving ligands in each subset (22–27 and 29–31) involve a conversion of a cyclopentane ring to a tetrahydropyrane ring which can be treated as a conventional terminal R-group modification. (23) However, all other conversions are between pairs of cyclic and acyclic ligands that involve the breaking of a ring and require more complex scaffold-hopping alchemical transformations. This EZH2 data set is an interesting benchmark because it is an example in which ligand rigidification, a commonly employed lead optimization strategy, actually leads to poorer binding. The data set also displays significant nonadditive effects that are difficult to predict. For example, decyclization in the presence of the tetrahydropyrane yields a much larger benefit than in the case of the cyclopentane substituent (Figure 6).

Figure 6

Figure 6. Relative binding free energy calculations for the EZH2 complexes. A representative complex of EZH2 bound to a ligand is demonstrated in the top left. The alignment frame used to apply a restraining potential to the positions and orientations of each ligand pair is illustrated in the bottom left. Relative free energy calculations for each ligand transformation are presented by arrows connecting each ligand pair (right). Free energy estimates in kcal/mol are color-coordinated according to the method: those computed by ATM-RBFE are in black, those obtained experimentally are in red, and those reported in literature are in blue. The same values are reported in Table 3. In the ligand structures, green represents carbon atoms; red, oxygen; blue, nitrogen; brown, bromine; and white, hydrogen.

Computational Details

The host–guest systems were prepared from the MOL2 files provided by the SAMPL8 organizers (54) as previously described. (45) All five guests were first aligned relative to each other, in terms of relative position and orientation, and then manually placed into the inner cavity of each host using Maestro (Schrödinger Inc.). The structures of the ERα and EZH2 protein complexes were also prepared in Maestro starting from the provided files. (23) Force-field parameter assignments and solvation of the molecular systems were performed using AmberTools 19. The GAFF1.8/AM1-BCC force field parameters (58,59) were assigned for the host–guest systems while Amber ff14SB parameters (15,60) were assigned to the protein receptors. The complexes were assembled using the LEaP program in AmberTools 19. The LEaP setup consisted of loading the receptor and a pair of aligned ligands, of which one was selected to be translated along the diagonal of a hypothetical solvent box so as to be positioned outside of the binding site cavity (Figure 1). This translation is approximately 38 Å in length, a value that was employed for both the GDCC host–guest and the protein–ligand systems to ensure that the displaced ligand in the solvent bulk is separated from the receptor by at least three layers of water molecules. Each system comprising of a bound complex and a displaced ligand, was solvated with a 10 Å buffer and the resulting rectangular solvent box was neutralized by the addition of an appropriate number of sodium or chloride ions.
The complexes were energy minimized and thermalized at 300 K. Starting from the bound state at λ = 0, the minimized systems were annealed to the symmetric alchemical intermediate at λ = 1/2 for 250 ps using the ATM alchemical potential energy function for leg 1 (eq 14). This step provides a suitable initial configuration of the system without severe unfavorable repulsion interactions at the alchemical intermediate state to seed the molecular dynamics simulations. The host molecules were loosely restrained with a flat-bottom harmonic potential of force constant 25.0 kcal/(mol Å2), and a tolerance of 1.5 Å was set on the heavy atoms at the lower cup of the molecule (the first 40 atoms of the host as listed in the provided files). The same positional restraints were applied to the Cα atoms of the protein receptors.
With some exceptions (see below), the linear alchemical potential, Wλ(u) = λusc(u) was used for all alchemical calculations with 11 λ-states uniformly distributed between λ = 0 and 1/2 for each of the two ATM legs. The calculation for the more challenging test without alignment restraints from TEMOA-G4 to TEMOA-G1 and for the EZH2 complexes employed the softplus alchemical potential (9) with the parameters listed in Table 5. The purpose of implementing the softplus alchemical potential for these transformations was to achieve convergence similar in time to that of the other transformations that employed the linear potential.
Table 1. Relative Binding Free Energies of the TEMOA and TEETOA Host–Guest Complexes
complex pairΔG1a,bΔG2a,cΔΔGb°(RBFE)a,dΔΔGb°(ABFE)a,eΔΔGb°(expt)a,f,g
TEMOA-G1-G2P41.68 ± 0.2147.34 ± 0.30–5.66 ± 0.37–6.39 ± 0.88 
TEMOA-G1-G320.67 ± 0.1823.01 ± 0.24–2.34 ± 0.30–1.55 ± 0.391.18 ± 0.22
TEMOA-G1-G415.98 ± 0.3018.45 ± 0.39–2.47 ± 0.49–1.92 ± 0.42–0.76 ± 0.22
TEMOA-G1-G514.53 ± 0.2416.39 ± 0.39–1.86 ± 0.46–0.99 ± 0.420.29 ± 0.22
TEMOA-G2P-G349.52 ± 0.3045.59 ± 0.363.93 ± 0.474.84 ± 0.87 
TEMOA-G2P-G445.19 ± 0.2742.05 ± 0.273.14 ± 0.384.47 ± 0.88 
TEMOA-G2P-G544.07 ± 0.2440.22 ± 0.273.85 ± 0.365.40 ± 0.88 
TEMOA-G3-G416.62 ± 0.3916.95 ± 0.39–0.33 ± 0.55–0.37 ± 0.39–1.94 ± 0.14
TEMOA-G3-G516.82 ± 0.4816.80 ± 0.450.02 ± 0.660.56 ± 0.39–0.89 ± 0.14
TEMOA-G4-G510.80 ± 0.2710.35 ± 0.240.44 ± 0.360.93 ± 0.42–1.05 ± 0.14
TEETOA-G1-G2P40.97 ± 0.2748.52 ± 0.27–7.55 ± 0.38–6.88 ± 0.67 
TEETOA-G1-G321.64 ± 0.2423.24 ± 0.27–1.60 ± 0.36–0.58 ± 0.66 
TEETOA-G1-G416.20 ± 0.3017.70 ± 0.39–1.50 ± 0.49–1.44 ± 0.660.2 ± 0.28
TEETOA-G1-G514.51 ± 0.2415.70 ± 0.39–1.19 ± 0.46–1.75 ± 0.651.17 ± 0.22
TEETOA-G2P-G351.89 ± 0.2744.94 ± 0.276.95 ± 0.386.30 ± 0.61 
TEETOA-G2P-G446.33 ± 0.3040.95 ± 0.275.38 ± 0.405.44 ± 0.61 
TEETOA-G2P-G544.46 ± 0.2439.08 ± 0.275.38 ± 0.365.13 ± 0.60 
TEETOA-G3-G417.90 ± 0.3018.82 ± 0.24–0.92 ± 0.38–0.86 ± 0.62 
TEETOA-G3-G516.69 ± 0.4517.09 ± 0.48–0.40 ± 0.66–1.17 ± 0.61 
TEETOA-G4-G510.43 ± 0.3010.47 ± 0.24–0.04 ± 0.38–0.31 ± 0.611.15 ± 0.22
a

In kcal/mol, errors are reported as 3 times the standard deviation.

b

This work, ATM-RBFE leg 1.

c

This work, ATM-RBFE leg 2.

d

This work, from the difference of the leg 1 and leg 2 free energies.

e

From the differences of absolute binding free energies from Table 4. (45)

f

From the differences of experimental binding free energies from Table 4. (45)

g

Experimental binding free energy values for the guest G2 in the protonated form (G2P) are unknown; TEETOA-G3 is a nonbinder experimentally.

Table 2. Relative Binding Free Energies of the ERα Complexes
complex pairΔG1a,bΔG2a,cΔΔGb°(RBFE)a,dΔΔGb°(expt)eΔΔGb°(lit.)a,f
2d–2e12.97 ± 0.2515.29 ± 0.28–2.32 ± 0.38–0.66–1.20 ± 0.12
2d–3a17.13 ± 0.3314.71 ± 0.362.42 ± 0.49>2.313.77 ± 0.30
2d–3b16.03 ± 0.3016.42 ± 0.36–0.39 ± 0.471.781.37 ± 0.36
2e–3a17.68 ± 0.2912.89 ± 0.274.79 ± 0.40>2.975.26 ± 0.30
2e–3b16.55 ± 0.2614.46 ± 0.262.09 ± 0.372.442.86 ± 0.33
3a–3b15.09 ± 0.2417.77 ± 0.29–2.68 ± 0.38<−0.53–2.36 ± 0.33
a

In kcal/mol, errors are reported as 3 times the standard deviation.

b

This work, ATM-RBFE leg 1.

c

This work, ATM-RBFE leg 2.

d

This work, from the difference of the leg 1 and leg 2 free energies.

e

Experimental RBFE values from ref (23), experimental errors were not reported.

f

From ref (23).

Table 3. Relative Binding Free Energies of the EZH2 Complexes
complex pairΔG1a,bΔG2a,cΔΔGb°(RBFE)a,dΔΔGb°(expt)a,eΔΔGb°(lit.)a,f
22–2719.90 ± 0.3019.46 ± 0.300.44 ± 0.421.32 ± 0.30.71 ± 0.21
22–2921.96 ± 0.3623.39 ± 0.33–1.43 ± 0.48–0.58 ± 0.5–0.45 ± 0.84
22–3122.84 ± 0.3625.80 ± 0.36–2.96 ± 0.51–0.58 ± 0.5–1.18 ± 0.75
27–2923.96 ± 0.3325.62 ± 0.36–1.66 ± 0.48–1.90 ± 0.4–2.17 ± 0.78
27–3123.28 ± 0.3627.67 ± 0.33–4.38 ± 0.48–1.90 ± 0.4–2.51 ± 0.75
29–3118.10 ± 0.3019.52 ± 0.30–1.41 ± 0.420.00 ± 0.6–0.34 ± 0.21
a

In kcal/mol, errors are reported as 3 times the standard deviation.

b

This work, ATM-RBFE leg 1.

c

This work, ATM-RBFE leg 2.

d

This work, from the difference of the leg 1 and leg 2 free energies.

e

From ref (57), errors as originally reported.

f

From ref (23), errors are reported as 3 times the standard deviation.

Table 4. Experimental and ATM Absolute Binding Free Energy Estimates for the Host–Guest Complexes from Reference (45)
complexΔG°(expt)a,bΔGb°(ABFE)a,c
TEMOA-G1–6.96 ± 0.2–6.71 ± 0.30
TEMOA-G2PNAd–13.10 ± 0.83
TEMOA-G3–5.78 ± 0.1–8.26 ± 0.25
TEMOA-G4–7.72 ± 0.1–8.63 ± 0.30
TEMOA-G5–6.67 ± 0.1–7.70 ± 0.30
TEETOA-G1–4.49 ± 0.1–1.07 ± 0.34
TEETOA-G2PNAd–7.95 ± 0.28
TEETOA-G3NBe–1.65 ± 0.30
TEETOA-G4–4.47 ± 0.2–2.51 ± 0.30
TEETOA-G5–3.32 ± 0.1–2.82 ± 0.28
a

In kcal/mol, computational uncertainties are reported as 3 times the standard deviation.

b

Experimental binding free energy. (54)

c

Calculated binding free energy. (45)

d

The experimental binding free energy of a specific protonation state is not available.

e

Binding not observed.

Table 5. Alchemical Schedule of the Softplus Alchemical Potential for the Two Legs of the TEMOA-G1 to TEMOA-G4 Transformation without Alignment Restraints and for the Alchemical Transformations for the EZH2 Complexes
λλ1λ2αau0bw0b
0.000.000.000.101500
0.050.000.050.101350
0.100.000.100.101200
0.150.000.150.101050
0.200.000.200.10900
0.250.000.250.10750
0.300.100.300.10600
0.350.200.350.10400
0.400.300.400.10400
0.450.400.450.10400
0.500.500.500.10400
a

In (kcal/mol)−1.

b

In kcal/mol.

The soft-core perturbation energy eq 11 was used for all calculations with umax = 200 kcal/mol, uc = 100 kcal/mol, and a = 1/16. The ligand was sequestered within the binding site by means of a flat-bottom harmonic potential between the centers of mass of the receptor and the ligand with a force constant of 25 kcal/mol Å2 applied for separation greater than 4.5 Å. Perturbation energy samples and trajectory frames were saved every 10 ps. Asynchronous Hamiltonian replica exchanges (41) in λ-space were performed every 10 ps. The Langevin thermostat with a time constant of 2 ps was used to maintain the temperature at 300 K. Each replica was simulated for a minimum of 20 ns for the host–guest systems and for 10 ns for the protein–ligand systems. Binding free energies and the corresponding uncertainties were computed from the perturbation energy samples using UWHAM (43) after discarding the first half of the trajectory.
Alignment restraints were imposed using three reference atoms of the ligands. For the SAMPL8 guests, alignment settings were customized for each pair to maintain an approximate alignment between the polar groups oriented toward the solvent and the body of the guest oriented toward the binding cavity of the host. The values of the force constants for the restraining potentials in eqs 6 and 8 were adjusted to allow for more flexibility between more dissimilar guests. Here, kr varied from 25 to 0.5 kcal/mol·Å2, kθ varied from 50 to 10 kcal/mol, and kψ varied from 50 to 0 kcal/mol. The reference atoms for the ERα and EZH2 ligands are illustrated in Figures 5 and 6, respectively. The corresponding force constants were set to kr = 2.5 kcal/mol/Å2, kθ = 10 kcal/mol, and kψ = 10 kcal/mol. The specific alignment settings for each system can be found in the publicly posted input files. (61)
The alchemical molecular dynamics simulations were performed with the OpenMM 7.3 (44) MD engine and the ATM integrator plugin (53) using the OpenCL platform. The ASyncRE software, (41) customized for OpenMM and ATM, (62) was used for the Hamiltonian replica exchange in λ space for each ATM leg. Simulation input files for this work are freely available. (61) The replica exchange simulations were performed on the XSEDE Comet and Expanse GPU HPC cluster at the San Diego Supercomputing Center, each using four GPUs per node.

Results

Click to copy section linkSection link copied!

SAMPL8 Host–Guest Systems

The calculated relative binding free energies for the SAMPL8 host–guest systems are reported in Table 1 and Figure 7. These values are compared to the corresponding absolute binding free energy estimates that are reported in ref (45) and reproduced in Table 4. The results obtained from tests conducted with alignment restraints on the TEMOA-G1-G4 transformation are shown in Table 6.

Figure 7

Figure 7. Comparison of the relative binding free energy estimates against the differences of the corresponding absolute values for the SAMPL8 benchmark set (Table 1). The line represents perfect agreement. The root-mean-square deviation between the two sets is 0.8 kcal/mol within statistical uncertainty.

Table 6. Calculated Binding Free Energy of TEMOA-G4 Relative to TEMOA-G1 with Various Alignment Restraints
protocolΔG1a,bΔG2a,cΔΔGb°(RBFE)a,d
ABFE differencee  –1.92 ± 0.42
RBFE full restraintse15.98 ± 0.3018.45 ± 0.39–2.47 ± 0.48
RBFE pos. restraints15.26 ± 0.2017.54 ± 0.27–2.28 ± 0.34
RBFE no restraints20.94 ± 0.2623.39 ± 0.30–2.45 ± 0.40
a

In kcal/mol, errors are reported as 3 times the standard deviation.

b

ATM-RBFE leg 1.

c

ATM-RBFE leg 2.

d

From the difference of the leg 1 and leg 2 free energies.

e

From Table 1.

As shown in Table 1, the calculated RBFE values (fourth column) for the 20 pairs of complexes, ten for each SAMPL8 host, are in good agreement with the corresponding ABFE estimates (fifth column). The root-mean-square deviation between the two sets of free energy differences is only 0.8 kcal/mol, a value well within statistical uncertainties. Only for three of the 20 pairs (TEMOA-G1-G3, TEMOA-G2P-G4, and TEMOA-G2P-G5), the deviation between the RBFE and ABFE binding free energy estimates falls slightly outside the uncertainty bounds. As two of the three cases have the largest ABFE statistical uncertainty of the set, we suspect that these inconsistencies are caused by systematic bias in the ABFE estimates obtained previously. (45) The agreement with the experimental RBFEs (sixth column in Table 1) is not as good, reflecting the mixed level of prediction accuracy achieved with this force field combination. (45,63) The comparison with experimental measurements is also incomplete due to the lack of binding free energy estimates for TEETOA-G3 and for all complexes involving the protonated form of the G2 guest.
The statistical uncertainties of the RBFE estimates are found to be consistently smaller than those of the differences of ABFE estimates. For example, the ATM-RBFE uncertainty for the large G1 to G3 scaffold-hopping transformation with TEETOA is 0.36 kcal/mol compared to an uncertainty of 0.66 kcal/mol from the differences of the corresponding ATM-ABFE estimates. The RBFE estimates are also found to have a high level of self-consistency with an average 3-point cycle closure error of only 0.25 kcal/mol. The latter value is computed by evaluating the free energy change along all of the triangular cycles in the network of transformations in Figure 4.
As illustrated in Table 1, the RBFE estimates are obtained from the differences between the free energy changes for the two legs (ΔG1 and ΔG2 in Table 1). The free energy changes for each leg tend to be of moderate magnitude (between 10 and 20 kcal/mol in magnitude) for the transformations between guests with the same net charge compared to the much larger free energy changes in ABFE calculations, which range between 40 and 50 kcal/mol. (45) This trend, which is remarkably maintained even for large scaffold-hopping transformations such as G1 to G3, can be justified by the large desolvation penalty incurred from binding a negatively charged guest to the host that cancels out in the RBFE calculations. Conversely, the trend is reversed for the RBFE transformations involving the neutral protonated G2 guest (G2P in Tables 1 and 4), in which the ABFE leg free energies tend to be significantly smaller than for the RBFE transformations. Evidently, for these specific cases, the ABFE desolvation penalty is much smaller, whereas the RBFE transformations tend to contend with a change of net charge. Nevertheless, even in these cases, the RBFE calculations yield consistent free energy estimates with reasonably small statistical uncertainties.

Test of the Alignment Restraints

As illustrated in Theory and Methods, the restraining potentials to keep the positions and orientations of the ligands in approximate alignment are designed to improve the convergence of the binding free energy estimates without biasing the results. We tested this fact numerically on the RBFE calculation from G1 to G4 for the TEMOA host. The results, shown in Table 6, largely confirm the theoretical expectations. The RBFE estimates obtained without alignment restraints (that is with only binding site restraints) agree with the estimates involving both full alignment restraints and only positional restraints (that is by turning off orientation restraints) within statistical uncertainty.
In addition to confirming theory, this data shows that, despite the similarities between the G1 and G4 scaffolds (Figure 4), strict alignment between two ligands is not necessarily beneficial for the RBFE calculation. For example, the estimate obtained without orientation restraints shows slightly smaller free energy changes for the two legs with smaller statistical uncertainties compared to that with full restraints. This suggests that allowing for more flexibility can lower the free energy of the intermediate state in which the two ligands simultaneously occupy the binding cavity and thus facilitate convergence.
The RBFE calculation without restraints, while also agreeing with the other estimates and with moderate uncertainty, suffers large entropic bottlenecks at the initial states of each leg (not shown). This entropy loss corresponds to the loss of translational and rotational degrees of freedom upon binding (46) and requires the use of the optimized softplus alchemical potential (47) to reach convergence (see Computational Details).
While more investigation on the role of the alignment restraints is needed, the results obtained so far suggest that some level of alignment restraints is beneficial to reach rapid convergence of the RBFE estimates with a general-purpose linear alchemical potential.

Protein–Ligand Complexes

The ATM-RBFE values calculated for the six pairs of ERα complexes (Figure 5) and the six pairs of EZH2 complexes (Figure 6) are shown in Tables 2 and 3, respectively, compared to the corresponding experimental values (55,57) and the literature estimates reported by Wang et al. (23) Most of the RBFE calculations in these sets involve scaffold hopping transformations. In the ERα set, with the exception of one (2d–2e in Figure 5), all transformations involve ring reduction or expansion. The EZH2 set is characterized by ring closure/opening scaffold-hopping modifications.
The ATM-RBFE estimates for the protein–ligand systems are in reasonable agreement with the measured inhibition potencies. The calculations correctly predict that for ERα, ligand 2e, which contains a 5-membered ring, is the strongest binder and 3a, which contains a 6-membered ring, is the weakest binder. The EZH2 set shows that, counterintuitively, rigidification of the ligand scaffold by cyclization does not lead to better binding. For example, transformations from either ligand 27 or 22 to either ligand 31 or 29 in Figure 6, involve the opening of the five-membered indazole ring, a modification that yields a more favorable binding free energy than the counter modification of closing the groups to render the indazole ring. In agreement with experiments (column five in Table 3), the ATM-RBFE calculations indeed show a gain in binding affinity for the EZH2 ligand set (Figure 6) when the core indazole ring is opened especially in the presence of the tetrahydropyran ring (transformation 27 to 31) relative to the cyclopentane ring (transformation 22 to 29).
The ATM-RBFE protein–ligand binding free energy estimates obtained here are in good quantitative agreement with the computational results by Wang et al. (23) even though these were obtained with a different force field model (OPLS3 in the case of Wang et al. and Amber ff14SB and GAFF2 in this work). The estimates obtained here have similar uncertainties as the literature values and also display a good level of self-consistency with small average 3-point cycle closure errors (0.1 kcal/mol for the ERα set and 0.6 kcal/mol for the EZH2 set), which are well within the statistical level of confidence for the calculations.

Discussion

Click to copy section linkSection link copied!

Binding free energy estimation remains one of the most challenging problems in modern computational chemistry. This is particularly so for the conformationally variable macromolecular targets relevant to structure-based drug discovery. Relative binding free energy (RBFE) methods based on alchemical transformations address some of these challenges by directly comparing pairs of ligands that share similar binding modes thereby bypassing, for example, the modeling of the transition from the apo to the holo state of the receptor necessary for absolute binding free energy estimation (ABFE). However, alchemical RBFE methods based on conventional single- and dual-topology setups have been traditionally limited to simple R-group modifications, which considerably impedes their range of applicability.
Recently, advanced alchemical scaffold-hopping methods have been introduced that allow for more complex changes in molecular topologies involving the breaking and forming of covalent bonds. Wang et al. (23) introduced a customized soft bond stretch potential to smoothly form and break bonds. More recently, Zou et al. (24) introduced a scheme based on auxiliary restraining potentials to achieve the same goal. Multisite λ-dynamics has been presented as an alternative to scaffold hopping transformations involving ring opening and ring expansion/reduction. (64) These advances require a considerable amount of development and testing effort and are currently only available on commercial platforms, namely the Schrödinger FEP+, the XTalPi XFEP, and the BIOVIA Discovery Studio platforms, respectively. As academic efforts have lagged behind in this arena, (26,27,32,65) open-source software packages have not yet reached a comparable level of robustness for production applications.
Scaffold-hopping transformations are only one of the many challenges that prevent the widespread implementation and use of alchemical RBFE tools. Among others, single- and dual-topology RBFE implementations generally require the customization of energy and force routines of the MD engine, the design of pairwise soft-core potentials, and the treatment of dummy groups and their attachment to the core of the ligand. In addition, RBFE calculations generally require the mapping of the corresponding atoms of the ligands, the setup of separate simulations for the bound and unbound states, and the treatment of changes in net charge of the ligands. (7,8) The substantial intellectual and development efforts involved in the correct design, testing, and applications of alchemical RBFE software tools can be discouraging to many practitioners.
In this work, we have presented a streamlined RBFE protocol based on the alchemical transfer method (ATM) (29) that promises to remove many of these obstacles while retaining a similar level of computational efficiency as the most advanced RBFE methods. Rather than mutating one ligand into another by interpolating force field parameters, as commonly done, (7) the ATM-RBFE protocol implements a simple coordinate transformation that transfers one ligand from the solvent bulk into the binding site while the other ligand is simultaneously transferred in the opposite direction. Similar to the double-system single box approach, (30,66) one advantage of this scheme is that it requires only one simulation system as opposed to two distinct ones for the solvation and receptor-bound mutations typical of standard approaches. (8) In addition to simplifying the system setup, using a single simulation box mitigates the artifacts observed in relative free energy estimates of ligands with different net charges that are sometimes addressed by correction factors (20,67) or by the coupled charging/decharging of counterions. (68) Instead, the ATM-RBFE protocol, based on the swapping of positions of the ligands in a single simulation box, maintains charge neutrality to help minimize finite-size effects (69) even when the two ligands have different net charges.
ATM does not require modifications of energy routines because the method is not dependent on soft-core pair potentials and nonphysical chemical topologies with dummy atoms. As a result, ATM can be more easily implemented as a controlling routine on top of existing force routines of MD engines (see Figure 3). The results presented in this work have been obtained using a plugin (53) of the open-source OpenMM molecular simulation library (44) that implements ATM in terms of a customized Langevin MD integrator. More recently, we have reimplemented the method as an OpenMM Force plugin (70) that yields more flexibility in terms of choice of MD integrators, as well as greater performance. The implementation of ATM does not require modifications of the energy routines on OpenMM. As described in Computation Protocol and Software Implementation, the method is based on collecting the potential energies and forces before and after the application of the coordinate transformations and combining them according to eq 9. On the downside, the repeated calculation of the energy and forces causes ATM to be slower than standard MD by approximately a factor of 2.
An important consequence of ATM’s design is that the method is compatible with any potential energy function, including many-body potential functions such as polarizable, (32,33,71) quantum-mechanical, (34−36) and artificial neural network (37,38) potentials and implicit solvent models. (72) In this work, for example, we validated the ATM-RBFE protocol with a particle mesh Ewald (PME) (73) long-range electrostatic potential without approximations and without customizations of the PME routines. We believe that the decoupling of the alchemical formulation and the potential energy function is an important feature of ATM that will help extend the range of applicability of alchemical free energy estimation tools.
ATM has its roots in the development of a statistical mechanics formulation of alchemical binding, (74) and the discovery of coordinate-based free energy perturbation schemes and novel alchemical potential functions capable of removing sampling bottlenecks along the alchemical thermodynamic path. (46,75) The method has been recently employed to the modeling of hydration (47) and molecular binding (29,45) with explicit solvation. In the present work, we combine ATM with a ligand alignment protocol that utilizes a restraining potential to model relative binding free energies. We tested the method on three stringent benchmarks that include scaffold-hopping and net charge modifications obtaining good agreement with ABFE calculations, literature values, and experimental observations.
Unlike single- and double-topology RBFE approaches, (7) in which the two ligands share all of some of the molecular topology, the ATM-RBFE protocol employs a representation where all the degrees of freedom of the two ligands are represented explicitly. Alignment restraining potentials are imposed on the relative positions and orientations of the ligand pairs to prevent one ligand from drifting away from the other and, by doing so, slowing down convergence. We have proven mathematically that the form of the alignment potential we employ does not introduce biases. We have also validated this assessment numerically. Restraining potentials that keep ligand pairs aligned have been previously used in relative binding free energy calculations; (76) however, to our knowledge, this is the first work that derives the requirements to ensure that they do not bias free energy estimates (see the Supporting Information).
While the ligand alignment potentials do not formally bias the free energy estimates, the results indicate that the parameters of restraining potentials can have a substantial effect on the convergence rate of the calculation. For pairs of small and rigid ligands, a variety of sensible choices of the alignment frames are likely to lead to nearly equivalent outcomes. In the case of the SAMPL8 guests, we found that alignment potentials designed to keep the charged ends of the two ligands together is a more useful technique than those designed to align the hydrophobic ends, presumably because the former preserved strong interactions with water molecules. For larger and more flexible ligand pairs, the alignment restraints can be imposed on a common rigid core, as we have done for the EZH2 set (Figure 6), leaving flexible side chains free to move. As in other applications, conformational rearrangements of side chains can hurt the rate of convergence if not sampled thoroughly. Out of concern that rearrangement may cause large motions of the axes of the ligands’ reference frames, we did not attempt to align the overall position and orientation of the two ligands by including reference atoms from the flexible side chains (see Figure 6). Although we have not seen occurrences in this work, we expect that there might not be good alignment choices for pairs of highly flexible and very dissimilar ligands. In such limiting cases it could be better to leave out alignment restraining potentials altogether, or to consider separate ABFE calculations for each ligand.
One interesting and promising feature of the protein–ligand RBFE calculations is that the magnitudes and the level of convergence of the free energy changes in each leg of the relative binding free energy calculation do not appear to depend strongly on the size of the alchemical transformation. For example, in the case of the ERα system in Figure 5, the large scaffold-hopping transformation from 2d to 3a, which involves a ring expansion, as well as the replacement and the displacement of a nonpolar R-group with a polar one, presents leg free energies ΔG1 and ΔG2 that are only slightly larger in magnitude and with similar uncertainties than those of the 2d to 2e transformation, which involves only a straightforward mutation of the three terminal atoms. Because the magnitudes of the leg free energies reflect the free energy of the alchemical intermediate relative to those of the physical end states, this result suggests that the ATM-RBFE alchemical pathway can avoid high free energy barriers typical of double-decoupling methods that rely on vacuum intermediates. (7,8,18)
Unlike double-decoupling approaches, (7) and similar to physical pathway methods, (77−80) and single-box alchemical approaches, (30) in ATM the unbound state of the complex is one in which the dissociated ligand is placed at a finite distance from the receptor. The effect of residual receptor–ligand interactions and correlations on the binding free energy estimates has not been comprehensively studied; however comparative efforts among computational methods and with experiments show that it is likely small. (29,45,81) In ATM, the dissociated ligand is positioned as far from the receptor as the simulation box size allows (Figure 1) (and never with fewer than three hydration layers separating it from receptor atoms). Hence, we tend to consider the small residual receptor–ligand correlations as finite size effects which can be controlled and made arbitrarily small, by increasing the simulation box size.

Conclusions

Click to copy section linkSection link copied!

We present the alchemical transfer method for the streamlined calculation of relative binding free energies of molecular complexes (ATM-RBFE). The work paves the way for a wider availability of advanced alchemical modeling tools for the community. The method is based on a coordinate transformation that swaps the positions of the ligands in a pair along an alchemical pathway. The method is applicable to both standard and scaffold hopping transformations. ATM-RBFE requires only one simulation system that is free of alchemical topologies and prepared with conventional tools, it does not require soft-core pair potentials and it does not require modifications of the energy routines of the MD engine. Moreover, ATM-RBFE does not require splitting the binding free energy calculations in receptor and solvation legs, nor into electrostatic and nonelectrostatic steps. We successfully validated the method on the SAMPL8 GDCC host–guest set and on two protein–ligand benchmark sets involving challenging scaffold hopping transformations that have so far required advanced tools of limited availability. The software implementation of ATM-RBFE, currently based on the highly performance OpenMM MD engine, is fully open-source and freely available. Future work will focus on reinforcing the limited validation tests conducted here with large-scale protein–ligand binding assessments benchmarks, on disseminating guidelines for the use of ATM-RBFE and on deployments on community MD engines.

Data and Software Availability

Click to copy section linkSection link copied!

The software used for this study and the simulation input files are freely available on Github. (53,61,62) The molecular dynamics trajectories and the data sets generated and/or analyzed during the current study are available from the corresponding author on reasonable request.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jcim.1c01129.

  • Review of theory underpinning the alchemical transfer method for absolute binding free energy estimation and derivation of the statistical mechanical extension for the relative binding free energy estimation with ligand alignment restraints (PDF)

Terms & Conditions

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

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Emilio Gallicchio - Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United StatesPh.D. Program in Chemistry, The Graduate Center of the City University of New York, New York, New York 10016, United StatesPh.D. Program in Biochemistry, The Graduate Center of the City University of New York, New York, New York 10016, United StatesOrcidhttps://orcid.org/0000-0002-2606-4913 Email: [email protected]
  • Authors
    • Solmaz Azimi - Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United StatesPh.D. Program in Biochemistry, The Graduate Center of the City University of New York, New York, New York 10016, United StatesOrcidhttps://orcid.org/0000-0002-9069-2914
    • Sheenam Khuttan - Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United StatesPh.D. Program in Biochemistry, The Graduate Center of the City University of New York, New York, New York 10016, United StatesOrcidhttps://orcid.org/0000-0002-8873-6359
    • Joe Z. Wu - Department of Chemistry, Brooklyn College of the City University of New York, Brooklyn, New York 11210, United StatesPh.D. Program in Chemistry, The Graduate Center of the City University of New York, New York, New York 10016, United StatesOrcidhttps://orcid.org/0000-0003-0681-8711
    • Rajat K. Pal - Roivant Sciences, Inc., Boston, Massachusetts 02210, United StatesOrcidhttps://orcid.org/0000-0001-9860-466X
  • Author Contributions

    S.A., S.K., and J.Z.W. contributed equally to this work.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We acknowledge support from the National Science Foundation (NSF CAREER 1750511 to E.G.). Molecular simulations were conducted on the Comet and Expanse GPU clusters at the San Diego Supercomputing Center supported by NSF XSEDE award TG-MCB150001.

References

Click to copy section linkSection link copied!

This article references 81 other publications.

  1. 1
    Jorgensen, W. L. The many roles of computation in drug discovery. Science 2004, 303, 18131818,  DOI: 10.1126/science.1096361
  2. 2
    Abel, R.; Wang, L.; Harder, E. D.; Berne, B.; Friesner, R. A. Advancing drug discovery through enhanced free energy calculations. Acc. Chem. Res. 2017, 50, 16251632,  DOI: 10.1021/acs.accounts.7b00083
  3. 3
    Armacost, K. A.; Riniker, S.; Cournia, Z. Novel directions in free energy methods and applications. J. Chem. Inf. Model. 2020, 60, 1,  DOI: 10.1021/acs.jcim.9b01174
  4. 4
    Wang, L.; Wu, Y.; Deng, Y.; Kim, B.; Pierce, L.; Krilov, G.; Lupyan, D.; Robinson, S.; Dahlgren, M. K.; Greenwood, J.; Romero, D. L.; Masse, C.; Knight, J. L.; Steinbrecher, T.; Beuming, T.; Damm, W.; Harder, E.; Sherman, W.; Brewer, M.; Wester, R.; Murcko, M.; Frye, L.; Farid, R.; Lin, T.; Mobley, D. L.; Jorgensen, W. L.; Berne, B. J.; Friesner, R. A.; Abel, R. Accurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force Field. J. Am. Chem. Soc. 2015, 137, 26952703,  DOI: 10.1021/ja512751q
  5. 5
    Cournia, Z.; Allen, B.; Sherman, W. Relative binding free energy calculations in drug discovery: recent advances and practical considerations. J. Chem. Inf. Model. 2017, 57, 29112937,  DOI: 10.1021/acs.jcim.7b00564
  6. 6
    Cournia, Z.; Allen, B. K.; Beuming, T.; Pearlman, D. A.; Radak, B. K.; Sherman, W. Rigorous free energy simulations in virtual screening. J. Chem. Inf. Model. 2020, 60, 41534169,  DOI: 10.1021/acs.jcim.0c00116
  7. 7
    Mey, A. S. J. S.; Allen, B. K.; Macdonald, H. E. B.; Chodera, J. D.; Hahn, D. F.; Kuhn, M.; Michel, J.; Mobley, D. L.; Naden, L. N.; Prasad, S.; Rizzi, A.; Scheen, J.; Shirts, M. R.; Tresadern, G.; Xu, H. Best Practices for Alchemical Free Energy Calculations [Article v1.0]. Living J. Comput. Mol. Sci. 2020, 2, 18378,  DOI: 10.33011/livecoms.2.1.18378
  8. 8
    Lee, T.-S.; Allen, B. K.; Giese, T. J.; Guo, Z.; Li, P.; Lin, C.; McGee, T. D., Jr; Pearlman, D. A.; Radak, B. K.; Tao, Y.; Tsai, H.-C.; Xu, H.; Sherman, W.; York, D. M. Alchemical Binding Free Energy Calculations in AMBER20: Advances and Best Practices for Drug Discovery. J. Chem. Inf. Model. 2020, 60, 55955623,  DOI: 10.1021/acs.jcim.0c00613
  9. 9
    Steinbrecher, T.; Joung, I.; Case, D. A. Soft-core potentials in thermodynamic integration: Comparing one- and two-step transformations. J. Comput. Chem. 2011, 32, 32533263,  DOI: 10.1002/jcc.21909
  10. 10
    Lee, T.-S.; Lin, Z.; Allen, B. K.; Lin, C.; Radak, B. K.; Tao, Y.; Tsai, H.-C.; Sherman, W.; York, D. M. Improved Alchemical Free Energy Calculations with Optimized Smoothstep Softcore Potentials. J. Chem. Theory Comput. 2020, 16, 55125525,  DOI: 10.1021/acs.jctc.0c00237
  11. 11
    Kim, S.; Oshima, H.; Zhang, H.; Kern, N. R.; Re, S.; Lee, J.; Roux, B.; Sugita, Y.; Jiang, W.; Im, W. CHARMM-GUI free energy calculator for absolute and relative ligand solvation and binding free energy simulations. J. Chem. Theory Comput. 2020, 16, 72077218,  DOI: 10.1021/acs.jctc.0c00884
  12. 12
    Zhang, H.; Kim, S.; Giese, T. J.; Lee, T.-S.; Lee, J.; York, D. M.; Im, W. CHARMM-GUI Free Energy Calculator for Practical Ligand Binding Free Energy Simulations with AMBER. J. Chem. Inf. Model. 2021, 61, 41454151,  DOI: 10.1021/acs.jcim.1c00747
  13. 13
    Liu, S.; Wu, Y.; Lin, T.; Abel, R.; Redmann, J. P.; Summa, C. M.; Jaber, V. R.; Lim, N. M.; Mobley, D. L. Lead optimization mapper: automating free energy calculations for lead optimization. J. Comput.-Aided Mol. Des. 2013, 27, 755770,  DOI: 10.1007/s10822-013-9678-y
  14. 14
    Fleck, M.; Wieder, M.; Boresch, S. Dummy Atoms in Alchemical Free Energy Calculations. J. Chem. Theory Comput. 2021, 17, 44034419,  DOI: 10.1021/acs.jctc.0c01328
  15. 15
    Zou, J.; Tian, C.; Simmerling, C. Blinded prediction of protein-ligand binding affinity using Amber thermodynamic integration for the 2018 D3R grand challenge 4. J. Comput.-Aided Mol. Des. 2019, 33, 10211029,  DOI: 10.1007/s10822-019-00223-x
  16. 16
    Gallicchio, E. In Computational Peptide Science: Methods and Protocols; Simonson, T., Ed.; Methods in Molecular Biology; Springer Nature, 2021.
  17. 17
    Jiang, W.; Chipot, C.; Roux, B. Computing relative binding affinity of ligands to receptor: An effective hybrid single-dual-topology free-energy perturbation approach in NAMD. J. Chem. Inf. Model. 2019, 59, 37943802,  DOI: 10.1021/acs.jcim.9b00362
  18. 18
    Rocklin, G. J.; Mobley, D. L.; Dill, K. A. Separated topologiesA method for relative binding free energy calculations using orientational restraints. J. Chem. Phys. 2013, 138, 085104,  DOI: 10.1063/1.4792251
  19. 19
    Chen, W.; Deng, Y.; Russell, E.; Wu, Y.; Abel, R.; Wang, L. Accurate calculation of relative binding free energies between ligands with different net charges. J. Chem. Theory Comput. 2018, 14, 63466358,  DOI: 10.1021/acs.jctc.8b00825
  20. 20
    Rocklin, G. J.; Mobley, D. L.; Dill, K. A.; Hünenberger, P. H. Calculating the binding free energies of charged species based on explicit-solvent simulations employing lattice-sum methods: An accurate correction scheme for electrostatic finite-size effects. J. Chem. Phys. 2013, 139, 184103,  DOI: 10.1063/1.4826261
  21. 21
    Dixit, S. B.; Chipot, C. Can absolute free energies of association be estimated from molecular mechanical simulations? The biotin-streptavidin system revisited. J. Phys. Chem. A 2001, 105, 97959799,  DOI: 10.1021/jp011878v
  22. 22
    Chen, W.; Wallace, J. A.; Yue, Z.; Shen, J. K. Introducing titratable water to all-atom molecular dynamics at constant pH. Biophys. J. 2013, 105, L15L17,  DOI: 10.1016/j.bpj.2013.06.036
  23. 23
    Wang, L.; Deng, Y.; Wu, Y.; Kim, B.; LeBard, D. N.; Wandschneider, D.; Beachy, M.; Friesner, R. A.; Abel, R. Accurate modeling of scaffold hopping transformations in drug discovery. J. Chem. Theory Comput. 2017, 13, 4254,  DOI: 10.1021/acs.jctc.6b00991
  24. 24
    Zou, J.; Li, Z.; Liu, S.; Peng, C.; Fang, D.; Wan, X.; Lin, Z.; Lee, T.-S.; Raleigh, D. P.; Yang, M.; Simmerling, C. Scaffold Hopping Transformations Using Auxiliary Restraints for Calculating Accurate Relative Binding Free Energies. J. Chem. Theory Comput. 2021, 17, 37103726,  DOI: 10.1021/acs.jctc.1c00214
  25. 25
    Loeffler, H. H.; Bosisio, S.; Duarte Ramos Matos, G.; Suh, D.; Roux, B.; Mobley, D. L.; Michel, J. Reproducibility of free energy calculations across different molecular simulation software packages. J. Chem. Theory Comput. 2018, 14, 55675582,  DOI: 10.1021/acs.jctc.8b00544
  26. 26
    Jespers, W.; Esguerra, M.; Åqvist, J.; Gutiérrez-de Terán, H. QligFEP: an automated workflow for small molecule free energy calculations in Q. J. Cheminf 2019, 11, 26,  DOI: 10.1186/s13321-019-0348-5
  27. 27
    Vilseck, J. Z.; Sohail, N.; Hayes, R. L.; Brooks, C. L., III Overcoming challenging substituent perturbations with multisite λ-dynamics: a case study targeting β-secretase 1. J. Phys. Chem. Lett. 2019, 10, 48754880,  DOI: 10.1021/acs.jpclett.9b02004
  28. 28
    Ligandswap. https://siremol.org/pages/apps/ligandswap.html (accessed September 12, 2021).
  29. 29
    Wu, J. Z.; Azimi, S.; Khuttan, S.; Deng, N.; Gallicchio, E. Alchemical Transfer Approach to Absolute Binding Free Energy Estimation. J. Chem. Theory Comput. 2021, 17, 3309,  DOI: 10.1021/acs.jctc.1c00266
  30. 30
    Gapsys, V.; Michielssens, S.; Peters, J. H.; de Groot, B. L.; Leonov, H. Molecular Modeling of Proteins; Springer, 2015; pp 173209.
  31. 31
    Macchiagodena, M.; Pagliai, M.; Karrenbrock, M.; Guarnieri, G.; Iannone, F.; Procacci, P. Virtual Double-System Single-Box: A Nonequilibrium Alchemical Technique for Absolute Binding Free Energy Calculations: Application to Ligands of the SARS-CoV-2 Main Protease. J. Chem. Theory Comput. 2020, 16, 71607172,  DOI: 10.1021/acs.jctc.0c00634
  32. 32
    Harger, M.; Li, D.; Wang, Z.; Dalby, K.; Lagardère, L.; Piquemal, J.-P.; Ponder, J.; Ren, P. Tinker-OpenMM: Absolute and relative alchemical free energies using AMOEBA on GPUs. J. Comput. Chem. 2017, 38, 20472055,  DOI: 10.1002/jcc.24853
  33. 33
    Panel, N.; Villa, F.; Fuentes, E. J.; Simonson, T. Accurate PDZ/peptide binding specificity with additive and polarizable free energy simulations. Biophys. J. 2018, 114, 10911102,  DOI: 10.1016/j.bpj.2018.01.008
  34. 34
    Beierlein, F. R.; Michel, J.; Essex, J. W. A simple QM/MM approach for capturing polarization effects in protein- ligand binding free energy calculations. J. Phys. Chem. B 2011, 115, 49114926,  DOI: 10.1021/jp109054j
  35. 35
    Lodola, A.; De Vivo, M. Adv. Protein Chem. Struct. Biol.; Elsevier, 2012; Vol. 87; pp 337362.
  36. 36
    Hudson, P. S.; Woodcock, H. L.; Boresch, S. Use of interaction energies in QM/MM free energy simulations. J. Chem. Theory Comput. 2019, 15, 46324645,  DOI: 10.1021/acs.jctc.9b00084
  37. 37
    Smith, J. S.; Nebgen, B. T.; Zubatyuk, R.; Lubbers, N.; Devereux, C.; Barros, K.; Tretiak, S.; Isayev, O.; Roitberg, A. E. Approaching coupled cluster accuracy with a general-purpose neural network potential through transfer learning. Nature Commun. 2019, 10, 18,  DOI: 10.1038/s41467-019-10827-4
  38. 38
    Rufa, D. A.; Macdonald, H. E. B.; Fass, J.; Wieder, M.; Grinaway, P. B.; Roitberg, A. E.; Isayev, O.; Chodera, J. D. Towards chemical accuracy for alchemical free energy calculations with hybrid physics-based machine learning/molecular mechanics potentials. bioRxiv , 2020, DOI: 10.1101/2020.07.29.227959
  39. 39
    Zhang, B.; Kilburg, D.; Eastman, P.; Pande, V. S.; Gallicchio, E. Efficient Gaussian Density Formulation of Volume and Surface Areas of Macromolecules on Graphical Processing Units. J. Comput. Chem. 2017, 38, 740752,  DOI: 10.1002/jcc.24745
  40. 40
    Spiriti, J.; Subramanian, S. R.; Palli, R.; Wu, M.; Zuckerman, D. M. Middle-way flexible docking: Pose prediction using mixed-resolution Monte Carlo in estrogen receptor α. PloS One 2019, 14, e0215694,  DOI: 10.1371/journal.pone.0215694
  41. 41
    Gallicchio, E.; Xia, J.; Flynn, W. F.; Zhang, B.; Samlalsingh, S.; Mentes, A.; Levy, R. M. Asynchronous replica exchange software for grid and heterogeneous computing. Comput. Phys. Commun. 2015, 196, 236246,  DOI: 10.1016/j.cpc.2015.06.010
  42. 42
    Shirts, M. R.; Chodera, J. D. Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 2008, 129, 124105,  DOI: 10.1063/1.2978177
  43. 43
    Tan, Z.; Gallicchio, E.; Lapelosa, M.; Levy, R. M. Theory of binless multi-state free energy estimation with applications to protein-ligand binding. J. Chem. Phys. 2012, 136, 144102,  DOI: 10.1063/1.3701175
  44. 44
    Eastman, P.; Swails, J.; Chodera, J. D.; McGibbon, R. T.; Zhao, Y.; Beauchamp, K. A.; Wang, L.-P.; Simmonett, A. C.; Harrigan, M. P.; Stern, C. D.; Wiewiora, R. P.; Brooks, B. R.; Pande, V. S. OpenMM 7: Rapid development of high performance algorithms for molecular dynamics. PLoS Comput. Biol. 2017, 13, e1005659,  DOI: 10.1371/journal.pcbi.1005659
  45. 45
    Azimi, S.; Khuttan, S.; Wu, J. Z.; Deng, N.; Gallicchio, E. Application of the Alchemical Transfer and Potential of Mean Force Methods to the SAMPL8 Cavitand Host-Guest Blinded Challenge. arXiv.org , 2021, 2107.05155.
  46. 46
    Pal, R. K.; Gallicchio, E. Perturbation potentials to overcome order/disorder transitions in alchemical binding free energy calculations. J. Chem. Phys. 2019, 151, 124116,  DOI: 10.1063/1.5123154
  47. 47
    Khuttan, S.; Azimi, S.; Wu, J. Z.; Gallicchio, E. Alchemical Transformations for Concerted Hydration Free Energy Estimation with Explicit Solvation. J. Chem. Phys. 2021, 154, 054103,  DOI: 10.1063/5.0036944
  48. 48
    Gallicchio, E.; Levy, R. M. Recent Theoretical and Computational Advances for Modeling Protein-Ligand Binding Affinities. Adv. Prot. Chem. Struct. Biol. 2011, 85, 2780,  DOI: 10.1016/B978-0-12-386485-7.00002-8
  49. 49
    Gilson, M. K.; Given, J. A.; Bush, B. L.; McCammon, J. A. The Statistical-Thermodynamic Basis for Computation of Binding Affinities: A Critical Review. Biophys. J. 1997, 72, 10471069,  DOI: 10.1016/S0006-3495(97)78756-3
  50. 50
    Roux, B.; Simonson, T. Implicit Solvent Models. Biophys. Chem. 1999, 78, 120,  DOI: 10.1016/S0301-4622(98)00226-9
  51. 51
    Boresch, S.; Tettinger, F.; Leitgeb, M.; Karplus, M. Absolute binding free energies: A quantitative approach for their calculation. J. Phys. Chem. B 2003, 107, 95359551,  DOI: 10.1021/jp0217839
  52. 52
    Chipot; ; Pohorille, Eds. Free Energy Calculations. Theory and Applications in Chemistry and Biology; Springer Series in Chemical Physics; Springer: Berlin Heidelberg, 2007.
  53. 53
    The Single-Decoupling Plugin for OpenMM. https://github.com/Gallicchio-Lab/openmm_sdm_plugin (accessed September 12, 2021).
  54. 54
    SAMPL8 host–guest GDCC challenge. https://github.com/samplchallenges/SAMPL8/tree/master/host_guest/GDCC (accessed September 12, 2021).
  55. 55
    Norman, B. H.; Dodge, J. A.; Richardson, T. I.; Borromeo, P. S.; Lugar, C. W.; Jones, S. A.; Chen, K.; Wang, Y.; Durst, G. L.; Barr, R. J.; Montrose-Rafizadeh, C.; Osborne, H. E.; Amos, R. M.; Guo, S.; Boodhoo, A.; Krishnan, V. Benzopyrans are selective estrogen receptor β agonists with novel activity in models of benign prostatic hyperplasia. J. Med. Chem. 2006, 49, 61556157,  DOI: 10.1021/jm060491j
  56. 56
    SAMPL8 host–guest GDCC challenge submission 37. https://github.com/samplchallenges/SAMPL8/blob/master/host_guest/Analysis/Submissions/GDCC/GDCC-ATM.txt (accessed September 12, 2021).
  57. 57
    Kuntz, K. W.; Campbell, J. E.; Keilhack, H.; Pollock, R. M.; Knutson, S. K.; Porter-Scott, M.; Richon, V. M.; Sneeringer, C. J.; Wigle, T. J.; Allain, C. J.; Majer, C. R.; Moyer, M. P.; Copeland, R. A.; Chesworth, R. The importance of being me: magic methyls, methyltransferase inhibitors, and the discovery of tazemetostat. J. Med. Chem. 2016, 59, 15561564,  DOI: 10.1021/acs.jmedchem.5b01501
  58. 58
    Wang, J.; Wang, W.; Kollman, P. A.; Case, D. A. Automatic atom type and bond type perception in molecular mechanical calculations. J. Mol. Graphics Modell. 2006, 25, 247260,  DOI: 10.1016/j.jmgm.2005.12.005
  59. 59
    He, X.; Liu, S.; Lee, T.-S.; Ji, B.; Man, V. H.; York, D. M.; Wang, J. Fast, Accurate, and Reliable Protocols for Routine Calculations of Protein-Ligand Binding Affinities in Drug Design Projects Using AMBER GPU-TI with ff14SB/GAFF. ACS Omega 2020, 5, 46114619,  DOI: 10.1021/acsomega.9b04233
  60. 60
    Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.; Simmerling, C. ff14SB: improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theory Comput. 2015, 11, 36963713,  DOI: 10.1021/acs.jctc.5b00255
  61. 61
    Relative Binding Free Energy Calculations for Ligands with Diverse Scaffolds with the Alchemical Transfer Method, Simulation Input Files. https://github.com/Gallicchio-Lab/ATM-relative-binding-free-energy-paper (accessed September 12, 2021).
  62. 62
    The Asynchronous Replica Exchange Framework for OpenMM. https://github.com/Gallicchio-Lab/async_re-openmm (accessed September 12, 2021).
  63. 63
    Rizzi, A.; Murkli, S.; McNeill, J. N.; Yao, W.; Sullivan, M.; Gilson, M. K.; Chiu, M. W.; Isaacs, L.; Gibb, B. C.; Mobley, D. L.; Chodera, J. D. Overview of the SAMPL6 host-guest binding affinity prediction challenge. J. Comp.-Aided Mol. Des. 2018, 32, 937963,  DOI: 10.1007/s10822-018-0170-6
  64. 64
    Raman, E. P.; Paul, T. J.; Hayes, R. L.; Brooks, C. L., III Automated, accurate, and scalable relative protein-ligand binding free-energy calculations using lambda dynamics. J. Chem. Theory Comput. 2020, 16, 78957914,  DOI: 10.1021/acs.jctc.0c00830
  65. 65
    Woods, C. J.; Malaisree, M.; Hannongbua, S.; Mulholland, A. J. A water-swap reaction coordinate for the calculation of absolute protein-ligand binding free energies. J. Chem. Phys. 2011, 134, 054114,  DOI: 10.1063/1.3519057
  66. 66
    Procacci, P.; Macchiagodena, M. On the NS-DSSB unidirectional estimates in the SAMPL6 SAMPLing challenge. J. Comput.-Aided Mol. Des. 2021, 35, 1055,  DOI: 10.1007/s10822-021-00419-0
  67. 67
    Öhlknecht, C.; Lier, B.; Petrov, D.; Fuchs, J.; Oostenbrink, C. Correcting electrostatic artifacts due to net-charge changes in the calculation of ligand binding free energies. J. Comput. Chem. 2020, 41, 986999,  DOI: 10.1002/jcc.26143
  68. 68
    Pan, A. C.; Xu, H.; Palpant, T.; Shaw, D. E. Quantitative characterization of the binding and unbinding of millimolar drug fragments with molecular dynamics simulations. J. Chem. Theory Comput. 2017, 13, 33723377,  DOI: 10.1021/acs.jctc.7b00172
  69. 69
    Lin, Y.-L.; Aleksandrov, A.; Simonson, T.; Roux, B. An overview of electrostatic free energy computations for solutions and proteins. J. Chem. Theory Comput. 2014, 10, 26902709,  DOI: 10.1021/ct500195p
  70. 70
    The ATM Meta Force Plugin for OpenMM. https://github.com/Gallicchio-Lab/openmm-atmmetaforce-plugin (accessed September 12, 2021).
  71. 71
    Huang, J.; Lemkul, J. A.; Eastman, P. K.; MacKerell, A. D., Jr. Molecular dynamics simulations using the drude polarizable force field on GPUs with OpenMM: Implementation, validation, and benchmarks. J. Comput. Chem. 2018, 39, 16821689,  DOI: 10.1002/jcc.25339
  72. 72
    Gallicchio, E.; Deng, N.; He, P.; Wickstrom, L.; Perryman, A. L.; Santiago, D. N.; Forli, S.; Olson, A. J.; Levy, R. M. Virtual Screening of Integrase Inhibitors by Large Scale Binding Free Energy Calculations: the SAMPL4 Challenge. J. Comput.-Aided Mol. Des. 2014, 28, 475490,  DOI: 10.1007/s10822-014-9711-9
  73. 73
    Darden, T. A.; York, D. M.; Pedersen, L. G. Particle mesh Ewald: An NlogN method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 1008910092,  DOI: 10.1063/1.464397
  74. 74
    Kilburg, D.; Gallicchio, E. Analytical Model of the Free Energy of Alchemical Molecular Binding. J. Chem. Theory Comput. 2018, 14, 61836196,  DOI: 10.1021/acs.jctc.8b00967
  75. 75
    Gallicchio, E.; Lapelosa, M.; Levy, R. M. Binding Energy Distribution Analysis Method (BEDAM) for Estimation of Protein-Ligand Binding Affinities. J. Chem. Theory Comput. 2010, 6, 29612977,  DOI: 10.1021/ct1002913
  76. 76
    Riniker, S.; Christ, C. D.; Hansen, N.; Mark, A. E.; Nair, P. C.; van Gunsteren, W. F. Comparison of enveloping distribution sampling and thermodynamic integration to calculate binding free energies of phenylethanolamine N-methyltransferase inhibitors. J. Chem. Phys. 2011, 135, 024105,  DOI: 10.1063/1.3604534
  77. 77
    Deng, Y.; Roux, B. Computations of standard binding free energies with molecular dynamics simulations. J. Phys. Chem. B 2009, 113, 22342246,  DOI: 10.1021/jp807701h
  78. 78
    Deng, N.; Cui, D.; Zhang, B. W.; Xia, J.; Cruz, J.; Levy, R. Comparing alchemical and physical pathway methods for computing the absolute binding free energy of charged ligands. Phys. Chem. Chem. Phys. 2018, 20, 1708117092,  DOI: 10.1039/C8CP01524D
  79. 79
    Velez-Vega, C.; Gilson, M. K. Overcoming dissipation in the calculation of standard binding free energies by ligand extraction. J. Comput. Chem. 2013, 34, 23602371,  DOI: 10.1002/jcc.23398
  80. 80
    Hall, R.; Dixon, T.; Dickson, A. On calculating free energy differences using ensembles of transition paths. Frontiers Mol. Biosc. 2020, 7, 106,  DOI: 10.3389/fmolb.2020.00106
  81. 81
    Rizzi, A.; Jensen, T.; Slochower, D. R.; Aldeghi, M.; Gapsys, V.; Ntekoumes, D.; Bosisio, S.; Papadourakis, M.; Henriksen, N. M.; De Groot, B. L.; Cournia, Z.; Dickson, A.; Michel, J.; Gilson, M. K.; Shirts, M. R.; Mobley, D. L.; Chodera, J. D. The SAMPL6 SAMPLing challenge: Assessing the reliability and efficiency of binding free energy calculations. J. Comp. Aid. Mol. Des. 2020, 34, 601,  DOI: 10.1007/s10822-020-00290-5

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 30 publications.

  1. Emilio Gallicchio. Relative Binding Free Energy Estimation of Congeneric Ligands and Macromolecular Mutants with the Alchemical Transfer Method with Coordinate Swapping. Journal of Chemical Information and Modeling 2025, 65 (7) , 3706-3714. https://doi.org/10.1021/acs.jcim.5c00207
  2. Francesc Sabanés Zariquiey, Stephen E. Farr, Stefan Doerr, Gianni De Fabritiis. QuantumBind-RBFE: Accurate Relative Binding Free Energy Calculations Using Neural Network Potentials. Journal of Chemical Information and Modeling 2025, Article ASAP.
  3. Anika J. Friedman, Wei-Tse Hsu, Michael R. Shirts. Multiple Topology Replica Exchange of Expanded Ensembles for Multidimensional Alchemical Calculations. Journal of Chemical Theory and Computation 2025, 21 (1) , 230-240. https://doi.org/10.1021/acs.jctc.4c01268
  4. Solmaz Azimi, Emilio Gallicchio. Binding Selectivity Analysis from Alchemical Receptor Hopping and Swapping Free Energy Calculations. The Journal of Physical Chemistry B 2024, 128 (44) , 10841-10852. https://doi.org/10.1021/acs.jpcb.4c05732
  5. Han Zhang, Wonpil Im. Ligand Binding Affinity Prediction for Membrane Proteins with Alchemical Free Energy Calculation Methods. Journal of Chemical Information and Modeling 2024, 64 (14) , 5671-5679. https://doi.org/10.1021/acs.jcim.4c00764
  6. Shi Zhang, Timothy J. Giese, Tai-Sung Lee, Darrin M. York. Alchemical Enhanced Sampling with Optimized Phase Space Overlap. Journal of Chemical Theory and Computation 2024, 20 (9) , 3935-3953. https://doi.org/10.1021/acs.jctc.4c00251
  7. Zoe Cournia, Christophe Chipot. Applications of Free-Energy Calculations to Biomolecular Processes. A Collection. The Journal of Physical Chemistry B 2024, 128 (14) , 3299-3301. https://doi.org/10.1021/acs.jpcb.4c01283
  8. Zoe Cournia, Christophe Chipot. Applications of Free-Energy Calculations to Biomolecular Processes. A Collection. Journal of Chemical Information and Modeling 2024, 64 (7) , 2129-2131. https://doi.org/10.1021/acs.jcim.4c00349
  9. Gao Tu, Tingting Fu, Guoxun Zheng, Binbin Xu, Rongpei Gou, Ding Luo, Panpan Wang, Weiwei Xue. Computational Chemistry in Structure-Based Solute Carrier Transporter Drug Design: Recent Advances and Future Perspectives. Journal of Chemical Information and Modeling 2024, 64 (5) , 1433-1455. https://doi.org/10.1021/acs.jcim.3c01736
  10. Francesc Sabanés Zariquiey, Raimondas Galvelis, Emilio Gallicchio, John D. Chodera, Thomas E. Markland, Gianni De Fabritiis. Enhancing Protein–Ligand Binding Affinity Predictions Using Neural Network Potentials. Journal of Chemical Information and Modeling 2024, 64 (5) , 1481-1485. https://doi.org/10.1021/acs.jcim.3c02031
  11. Sheenam Khuttan, Emilio Gallicchio. What to Make of Zero: Resolving the Statistical Noise from Conformational Reorganization in Alchemical Binding Free Energy Estimates with Metadynamics Sampling. Journal of Chemical Theory and Computation 2024, 20 (3) , 1489-1501. https://doi.org/10.1021/acs.jctc.3c01250
  12. Peter Eastman, Raimondas Galvelis, Raúl P. Peláez, Charlles R. A. Abreu, Stephen E. Farr, Emilio Gallicchio, Anton Gorenko, Michael M. Henry, Frank Hu, Jing Huang, Andreas Krämer, Julien Michel, Joshua A. Mitchell, Vijay S. Pande, João PGLM Rodrigues, Jaime Rodriguez-Guerra, Andrew C. Simmonett, Sukrit Singh, Jason Swails, Philip Turner, Yuanqing Wang, Ivy Zhang, John D. Chodera, Gianni De Fabritiis, Thomas E. Markland. OpenMM 8: Molecular Dynamics Simulation with Machine Learning Potentials. The Journal of Physical Chemistry B 2024, 128 (1) , 109-116. https://doi.org/10.1021/acs.jpcb.3c06662
  13. Lieyang Chen, Yujie Wu, Chuanjie Wu, Ana Silveira, Woody Sherman, Huafeng Xu, Emilio Gallicchio. Performance and Analysis of the Alchemical Transfer Method for Binding-Free-Energy Predictions of Diverse Ligands. Journal of Chemical Information and Modeling 2024, 64 (1) , 250-264. https://doi.org/10.1021/acs.jcim.3c01705
  14. Candide Champion, René Gall, Benjamin Ries, Salomé R. Rieder, Emilia P. Barros, Sereina Riniker. Accelerating Alchemical Free Energy Prediction Using a Multistate Method: Application to Multiple Kinases. Journal of Chemical Information and Modeling 2023, 63 (22) , 7133-7147. https://doi.org/10.1021/acs.jcim.3c01469
  15. Darrin M. York. Modern Alchemical Free Energy Methods for Drug Discovery Explained. ACS Physical Chemistry Au 2023, 3 (6) , 478-491. https://doi.org/10.1021/acsphyschemau.3c00033
  16. Hannah M. Baumann, Eric Dybeck, Christopher L. McClendon, Frank C. Pickard, IV, Vytautas Gapsys, Laura Pérez-Benito, David F. Hahn, Gary Tresadern, Alan M. Mathiowetz, David L. Mobley. Broadening the Scope of Binding Free Energy Calculations Using a Separated Topologies Approach. Journal of Chemical Theory and Computation 2023, 19 (15) , 5058-5076. https://doi.org/10.1021/acs.jctc.3c00282
  17. Francesc Sabanés Zariquiey, Adrià Pérez, Maciej Majewski, Emilio Gallicchio, Gianni De Fabritiis. Validation of the Alchemical Transfer Method for the Estimation of Relative Binding Affinities of Molecular Series. Journal of Chemical Information and Modeling 2023, 63 (8) , 2438-2444. https://doi.org/10.1021/acs.jcim.3c00178
  18. Gabriela B. Correa, Jéssica C. S. L. Maciel, Frederico W. Tavares, Charlles R. A. Abreu. A New Formulation for the Concerted Alchemical Calculation of van der Waals and Coulomb Components of Solvation Free Energies. Journal of Chemical Theory and Computation 2022, 18 (10) , 5876-5889. https://doi.org/10.1021/acs.jctc.2c00563
  19. Piero Procacci. Relative Binding Free Energy between Chemically Distant Compounds Using a Bidirectional Nonequilibrium Approach. Journal of Chemical Theory and Computation 2022, 18 (6) , 4014-4026. https://doi.org/10.1021/acs.jctc.2c00295
  20. Solmaz Azimi, Emilio Gallicchio. Potential distribution theory of alchemical transfer. The Journal of Chemical Physics 2025, 162 (5) https://doi.org/10.1063/5.0244918
  21. Ryo C Yanagita, Yoshiyuki Suzuki, Yasuhiro Kawanami, Yusuke Hanaki, Kazuhiro Irie. Effect of phenolic-hydroxy-group incorporation on the biological activity of a simplified aplysiatoxin analog with an ( R )-(−)-carvone-based core. Bioscience, Biotechnology, and Biochemistry 2024, 88 (9) , 992-998. https://doi.org/10.1093/bbb/zbae091
  22. Son Tung Ngo, Quynh Mai Thai, Trung Hai Nguyen, Nguyen Ngoc Tuan, T. Ngoc Han Pham, Huong T. T. Phung, Duong Tuan Quang. Alchemical approach performance in calculating the ligand-binding free energy. RSC Advances 2024, 14 (21) , 14875-14885. https://doi.org/10.1039/D4RA00692E
  23. Lei Lu, Xuxu Gou, Sophia K. Tan, Samuel I. Mann, Hyunjun Yang, Xiaofang Zhong, Dimitrios Gazgalis, Jesús Valdiviezo, Hyunil Jo, Yibing Wu, Morgan E. Diolaiti, Alan Ashworth, Nicholas F. Polizzi, William F. DeGrado. De novo design of drug-binding proteins with predictable binding energy and specificity. Science 2024, 384 (6691) , 106-112. https://doi.org/10.1126/science.adl5364
  24. Martin Amezcua, Jeffry Setiadi, David L. Mobley. The SAMPL9 host–guest blind challenge: an overview of binding free energy predictive accuracy. Physical Chemistry Chemical Physics 2024, 26 (12) , 9207-9225. https://doi.org/10.1039/D3CP05111K
  25. Christophe Chipot, Paraskevi Gkeka, Tony Lelièvre, Gabriel Stoltz. Equilibrium and Nonequilibrium Methods for Free-Energy Calculations With Molecular Dynamics. 2024, 384-400. https://doi.org/10.1016/B978-0-12-821978-2.00112-4
  26. Fernando D. Prieto-Martínez, Yelzyn Galván-Ciprés. Free Energy Estimation for Drug Discovery: Background and Perspectives. 2023, 310-345. https://doi.org/10.2174/9789815179934123010011
  27. Sheenam Khuttan, Solmaz Azimi, Joe Z. Wu, Sebastian Dick, Chuanjie Wu, Huafeng Xu, Emilio Gallicchio. Taming multiple binding poses in alchemical binding free energy prediction: the β-cyclodextrin host–guest SAMPL9 blinded challenge. Physical Chemistry Chemical Physics 2023, 25 (36) , 24364-24376. https://doi.org/10.1039/D3CP02125D
  28. Marina Macchiagodena, Marco Pagliai, Piero Procacci. NE‐RDFE : A protocol and toolkit for computing relative dissociation free energies with GROMACS between dissimilar molecules using bidirectional nonequilibrium dual topology schemes. Journal of Computational Chemistry 2023, 44 (12) , 1221-1230. https://doi.org/10.1002/jcc.27077
  29. Davide Bassani, Stefano Moro. Past, Present, and Future Perspectives on Computer-Aided Drug Design Methodologies. Molecules 2023, 28 (9) , 3906. https://doi.org/10.3390/molecules28093906
  30. Felix Potlitz, Andreas Link, Lukas Schulig. Advances in the discovery of new chemotypes through ultra-large library docking. Expert Opinion on Drug Discovery 2023, 18 (3) , 303-313. https://doi.org/10.1080/17460441.2023.2171984

Journal of Chemical Information and Modeling

Cite this: J. Chem. Inf. Model. 2022, 62, 2, 309–323
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jcim.1c01129
Published January 6, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

10k

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Illustration of the ATM-RBFE calculation setup that consists of displacing one ligand from the binding site to the solvent bulk along a translation vector d (in green) while simultaneously translating the second ligand from the solvent bulk to the binding site along the same vector. The protein receptor (ERα) is shown in cartoon representation colored by secondary structure. The ligands are shown in van der Waals representation.

    Figure 2

    Figure 2. Free energy diagram for an ATM-RBFE calculation, which consists of two independent legs that are connected to a single alchemical intermediate state. The alchemical calculation for leg 1 begins at λ = 0, in which ligand A is bound to the binding site of the receptor R and ligand B is dissociated in the solvent bulk, and ends at λ = 1/2 at the alchemical intermediate (denoted by R(AB)1/2 + (BA)1/2), in which A and B are simultaneously present at 50% strength in the binding site and solvent bulk. The alchemical calculation for leg 2 begins with ligand B bound to the binding site and ligand A in the solvent bulk and ends at same alchemical intermediate. ΔG1 and ΔG2 correspond to the free energy changes along each respective ATM leg. The relative binding free energy, ΔΔGb°(B,A), of ligand B with respect to ligand A is the difference between the free energies of legs 1 and 2.

    Figure 3

    Figure 3. Schematic flow diagram of the MD software implementation of the ATM-RBFE method. At each time step, a modified MD integrator routine gives the current coordinates of the system, in which ligand A is bound to the receptor and ligand B is in the solvent bulk, to the MD engine energy/forces calculation routine (RA + B state, left side of the diagram). The coordinates of the system are then transformed to swap the positions of ligands A and B so that B becomes bound to the receptor and A is now present in the solvent bulk (RB + A state, on the right branch of the diagram) and are given to the energy/force calculation routine. The resulting sets of energies and forces are merged according to eq 9 by the energy/force merging routine and fed again to the MD integrator to initiate the subsequent time step. The blue-colored energy/forces calculations routines are used from the MD engine unmodified. The red-colored routines are customized for ATM.

    Figure 4

    Figure 4. ATM binding free energy cycle for the SAMPL8 GDCC benchmark set that includes two hosts, TEMOA (A) and TEETOA (B). A representative complex of each host bound to the guest G1 is shown at the bottom of each panel. Binding free energy estimates in kcal/mol are illustrated alongside arrows connecting each ligand pair transformation (top of each panel). Relative binding free energy estimates are represented in blue, and the difference of the absolute binding free energy for each guest pair are represented in red. These values are also tabulated in Table 1. On the host and guest structures, red corresponds to oxygen atoms and white to hydrogen atoms. Carbon atoms are represented in gray in the host structures and green in the guest structures.

    Figure 5

    Figure 5. Relative binding free energy calculations for the ERα complexes. A representative complex of ERα bound to a ligand is demonstrated in the top left. The alignment frame used to apply a restraining potential to the positions and orientations of each ligand pair is illustrated in the bottom left. Relative free energy calculations for each ligand transformation are presented by arrows connecting each ligand pair (right). Free energy estimates in kcal/mol are color-coordinated according to the method: those computed by ATM-RBFE are in black, those obtained experimentally are in red, and those reported in literature are in blue. The same values are reported in Table 2. In the ligand structures, green represents carbon atoms; red, oxygen; and cyan, fluorine.

    Figure 6

    Figure 6. Relative binding free energy calculations for the EZH2 complexes. A representative complex of EZH2 bound to a ligand is demonstrated in the top left. The alignment frame used to apply a restraining potential to the positions and orientations of each ligand pair is illustrated in the bottom left. Relative free energy calculations for each ligand transformation are presented by arrows connecting each ligand pair (right). Free energy estimates in kcal/mol are color-coordinated according to the method: those computed by ATM-RBFE are in black, those obtained experimentally are in red, and those reported in literature are in blue. The same values are reported in Table 3. In the ligand structures, green represents carbon atoms; red, oxygen; blue, nitrogen; brown, bromine; and white, hydrogen.

    Figure 7

    Figure 7. Comparison of the relative binding free energy estimates against the differences of the corresponding absolute values for the SAMPL8 benchmark set (Table 1). The line represents perfect agreement. The root-mean-square deviation between the two sets is 0.8 kcal/mol within statistical uncertainty.

  • References


    This article references 81 other publications.

    1. 1
      Jorgensen, W. L. The many roles of computation in drug discovery. Science 2004, 303, 18131818,  DOI: 10.1126/science.1096361
    2. 2
      Abel, R.; Wang, L.; Harder, E. D.; Berne, B.; Friesner, R. A. Advancing drug discovery through enhanced free energy calculations. Acc. Chem. Res. 2017, 50, 16251632,  DOI: 10.1021/acs.accounts.7b00083
    3. 3
      Armacost, K. A.; Riniker, S.; Cournia, Z. Novel directions in free energy methods and applications. J. Chem. Inf. Model. 2020, 60, 1,  DOI: 10.1021/acs.jcim.9b01174
    4. 4
      Wang, L.; Wu, Y.; Deng, Y.; Kim, B.; Pierce, L.; Krilov, G.; Lupyan, D.; Robinson, S.; Dahlgren, M. K.; Greenwood, J.; Romero, D. L.; Masse, C.; Knight, J. L.; Steinbrecher, T.; Beuming, T.; Damm, W.; Harder, E.; Sherman, W.; Brewer, M.; Wester, R.; Murcko, M.; Frye, L.; Farid, R.; Lin, T.; Mobley, D. L.; Jorgensen, W. L.; Berne, B. J.; Friesner, R. A.; Abel, R. Accurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force Field. J. Am. Chem. Soc. 2015, 137, 26952703,  DOI: 10.1021/ja512751q
    5. 5
      Cournia, Z.; Allen, B.; Sherman, W. Relative binding free energy calculations in drug discovery: recent advances and practical considerations. J. Chem. Inf. Model. 2017, 57, 29112937,  DOI: 10.1021/acs.jcim.7b00564
    6. 6
      Cournia, Z.; Allen, B. K.; Beuming, T.; Pearlman, D. A.; Radak, B. K.; Sherman, W. Rigorous free energy simulations in virtual screening. J. Chem. Inf. Model. 2020, 60, 41534169,  DOI: 10.1021/acs.jcim.0c00116
    7. 7
      Mey, A. S. J. S.; Allen, B. K.; Macdonald, H. E. B.; Chodera, J. D.; Hahn, D. F.; Kuhn, M.; Michel, J.; Mobley, D. L.; Naden, L. N.; Prasad, S.; Rizzi, A.; Scheen, J.; Shirts, M. R.; Tresadern, G.; Xu, H. Best Practices for Alchemical Free Energy Calculations [Article v1.0]. Living J. Comput. Mol. Sci. 2020, 2, 18378,  DOI: 10.33011/livecoms.2.1.18378
    8. 8
      Lee, T.-S.; Allen, B. K.; Giese, T. J.; Guo, Z.; Li, P.; Lin, C.; McGee, T. D., Jr; Pearlman, D. A.; Radak, B. K.; Tao, Y.; Tsai, H.-C.; Xu, H.; Sherman, W.; York, D. M. Alchemical Binding Free Energy Calculations in AMBER20: Advances and Best Practices for Drug Discovery. J. Chem. Inf. Model. 2020, 60, 55955623,  DOI: 10.1021/acs.jcim.0c00613
    9. 9
      Steinbrecher, T.; Joung, I.; Case, D. A. Soft-core potentials in thermodynamic integration: Comparing one- and two-step transformations. J. Comput. Chem. 2011, 32, 32533263,  DOI: 10.1002/jcc.21909
    10. 10
      Lee, T.-S.; Lin, Z.; Allen, B. K.; Lin, C.; Radak, B. K.; Tao, Y.; Tsai, H.-C.; Sherman, W.; York, D. M. Improved Alchemical Free Energy Calculations with Optimized Smoothstep Softcore Potentials. J. Chem. Theory Comput. 2020, 16, 55125525,  DOI: 10.1021/acs.jctc.0c00237
    11. 11
      Kim, S.; Oshima, H.; Zhang, H.; Kern, N. R.; Re, S.; Lee, J.; Roux, B.; Sugita, Y.; Jiang, W.; Im, W. CHARMM-GUI free energy calculator for absolute and relative ligand solvation and binding free energy simulations. J. Chem. Theory Comput. 2020, 16, 72077218,  DOI: 10.1021/acs.jctc.0c00884
    12. 12
      Zhang, H.; Kim, S.; Giese, T. J.; Lee, T.-S.; Lee, J.; York, D. M.; Im, W. CHARMM-GUI Free Energy Calculator for Practical Ligand Binding Free Energy Simulations with AMBER. J. Chem. Inf. Model. 2021, 61, 41454151,  DOI: 10.1021/acs.jcim.1c00747
    13. 13
      Liu, S.; Wu, Y.; Lin, T.; Abel, R.; Redmann, J. P.; Summa, C. M.; Jaber, V. R.; Lim, N. M.; Mobley, D. L. Lead optimization mapper: automating free energy calculations for lead optimization. J. Comput.-Aided Mol. Des. 2013, 27, 755770,  DOI: 10.1007/s10822-013-9678-y
    14. 14
      Fleck, M.; Wieder, M.; Boresch, S. Dummy Atoms in Alchemical Free Energy Calculations. J. Chem. Theory Comput. 2021, 17, 44034419,  DOI: 10.1021/acs.jctc.0c01328
    15. 15
      Zou, J.; Tian, C.; Simmerling, C. Blinded prediction of protein-ligand binding affinity using Amber thermodynamic integration for the 2018 D3R grand challenge 4. J. Comput.-Aided Mol. Des. 2019, 33, 10211029,  DOI: 10.1007/s10822-019-00223-x
    16. 16
      Gallicchio, E. In Computational Peptide Science: Methods and Protocols; Simonson, T., Ed.; Methods in Molecular Biology; Springer Nature, 2021.
    17. 17
      Jiang, W.; Chipot, C.; Roux, B. Computing relative binding affinity of ligands to receptor: An effective hybrid single-dual-topology free-energy perturbation approach in NAMD. J. Chem. Inf. Model. 2019, 59, 37943802,  DOI: 10.1021/acs.jcim.9b00362
    18. 18
      Rocklin, G. J.; Mobley, D. L.; Dill, K. A. Separated topologiesA method for relative binding free energy calculations using orientational restraints. J. Chem. Phys. 2013, 138, 085104,  DOI: 10.1063/1.4792251
    19. 19
      Chen, W.; Deng, Y.; Russell, E.; Wu, Y.; Abel, R.; Wang, L. Accurate calculation of relative binding free energies between ligands with different net charges. J. Chem. Theory Comput. 2018, 14, 63466358,  DOI: 10.1021/acs.jctc.8b00825
    20. 20
      Rocklin, G. J.; Mobley, D. L.; Dill, K. A.; Hünenberger, P. H. Calculating the binding free energies of charged species based on explicit-solvent simulations employing lattice-sum methods: An accurate correction scheme for electrostatic finite-size effects. J. Chem. Phys. 2013, 139, 184103,  DOI: 10.1063/1.4826261
    21. 21
      Dixit, S. B.; Chipot, C. Can absolute free energies of association be estimated from molecular mechanical simulations? The biotin-streptavidin system revisited. J. Phys. Chem. A 2001, 105, 97959799,  DOI: 10.1021/jp011878v
    22. 22
      Chen, W.; Wallace, J. A.; Yue, Z.; Shen, J. K. Introducing titratable water to all-atom molecular dynamics at constant pH. Biophys. J. 2013, 105, L15L17,  DOI: 10.1016/j.bpj.2013.06.036
    23. 23
      Wang, L.; Deng, Y.; Wu, Y.; Kim, B.; LeBard, D. N.; Wandschneider, D.; Beachy, M.; Friesner, R. A.; Abel, R. Accurate modeling of scaffold hopping transformations in drug discovery. J. Chem. Theory Comput. 2017, 13, 4254,  DOI: 10.1021/acs.jctc.6b00991
    24. 24
      Zou, J.; Li, Z.; Liu, S.; Peng, C.; Fang, D.; Wan, X.; Lin, Z.; Lee, T.-S.; Raleigh, D. P.; Yang, M.; Simmerling, C. Scaffold Hopping Transformations Using Auxiliary Restraints for Calculating Accurate Relative Binding Free Energies. J. Chem. Theory Comput. 2021, 17, 37103726,  DOI: 10.1021/acs.jctc.1c00214
    25. 25
      Loeffler, H. H.; Bosisio, S.; Duarte Ramos Matos, G.; Suh, D.; Roux, B.; Mobley, D. L.; Michel, J. Reproducibility of free energy calculations across different molecular simulation software packages. J. Chem. Theory Comput. 2018, 14, 55675582,  DOI: 10.1021/acs.jctc.8b00544
    26. 26
      Jespers, W.; Esguerra, M.; Åqvist, J.; Gutiérrez-de Terán, H. QligFEP: an automated workflow for small molecule free energy calculations in Q. J. Cheminf 2019, 11, 26,  DOI: 10.1186/s13321-019-0348-5
    27. 27
      Vilseck, J. Z.; Sohail, N.; Hayes, R. L.; Brooks, C. L., III Overcoming challenging substituent perturbations with multisite λ-dynamics: a case study targeting β-secretase 1. J. Phys. Chem. Lett. 2019, 10, 48754880,  DOI: 10.1021/acs.jpclett.9b02004
    28. 28
      Ligandswap. https://siremol.org/pages/apps/ligandswap.html (accessed September 12, 2021).
    29. 29
      Wu, J. Z.; Azimi, S.; Khuttan, S.; Deng, N.; Gallicchio, E. Alchemical Transfer Approach to Absolute Binding Free Energy Estimation. J. Chem. Theory Comput. 2021, 17, 3309,  DOI: 10.1021/acs.jctc.1c00266
    30. 30
      Gapsys, V.; Michielssens, S.; Peters, J. H.; de Groot, B. L.; Leonov, H. Molecular Modeling of Proteins; Springer, 2015; pp 173209.
    31. 31
      Macchiagodena, M.; Pagliai, M.; Karrenbrock, M.; Guarnieri, G.; Iannone, F.; Procacci, P. Virtual Double-System Single-Box: A Nonequilibrium Alchemical Technique for Absolute Binding Free Energy Calculations: Application to Ligands of the SARS-CoV-2 Main Protease. J. Chem. Theory Comput. 2020, 16, 71607172,  DOI: 10.1021/acs.jctc.0c00634
    32. 32
      Harger, M.; Li, D.; Wang, Z.; Dalby, K.; Lagardère, L.; Piquemal, J.-P.; Ponder, J.; Ren, P. Tinker-OpenMM: Absolute and relative alchemical free energies using AMOEBA on GPUs. J. Comput. Chem. 2017, 38, 20472055,  DOI: 10.1002/jcc.24853
    33. 33
      Panel, N.; Villa, F.; Fuentes, E. J.; Simonson, T. Accurate PDZ/peptide binding specificity with additive and polarizable free energy simulations. Biophys. J. 2018, 114, 10911102,  DOI: 10.1016/j.bpj.2018.01.008
    34. 34
      Beierlein, F. R.; Michel, J.; Essex, J. W. A simple QM/MM approach for capturing polarization effects in protein- ligand binding free energy calculations. J. Phys. Chem. B 2011, 115, 49114926,  DOI: 10.1021/jp109054j
    35. 35
      Lodola, A.; De Vivo, M. Adv. Protein Chem. Struct. Biol.; Elsevier, 2012; Vol. 87; pp 337362.
    36. 36
      Hudson, P. S.; Woodcock, H. L.; Boresch, S. Use of interaction energies in QM/MM free energy simulations. J. Chem. Theory Comput. 2019, 15, 46324645,  DOI: 10.1021/acs.jctc.9b00084
    37. 37
      Smith, J. S.; Nebgen, B. T.; Zubatyuk, R.; Lubbers, N.; Devereux, C.; Barros, K.; Tretiak, S.; Isayev, O.; Roitberg, A. E. Approaching coupled cluster accuracy with a general-purpose neural network potential through transfer learning. Nature Commun. 2019, 10, 18,  DOI: 10.1038/s41467-019-10827-4
    38. 38
      Rufa, D. A.; Macdonald, H. E. B.; Fass, J.; Wieder, M.; Grinaway, P. B.; Roitberg, A. E.; Isayev, O.; Chodera, J. D. Towards chemical accuracy for alchemical free energy calculations with hybrid physics-based machine learning/molecular mechanics potentials. bioRxiv , 2020, DOI: 10.1101/2020.07.29.227959
    39. 39
      Zhang, B.; Kilburg, D.; Eastman, P.; Pande, V. S.; Gallicchio, E. Efficient Gaussian Density Formulation of Volume and Surface Areas of Macromolecules on Graphical Processing Units. J. Comput. Chem. 2017, 38, 740752,  DOI: 10.1002/jcc.24745
    40. 40
      Spiriti, J.; Subramanian, S. R.; Palli, R.; Wu, M.; Zuckerman, D. M. Middle-way flexible docking: Pose prediction using mixed-resolution Monte Carlo in estrogen receptor α. PloS One 2019, 14, e0215694,  DOI: 10.1371/journal.pone.0215694
    41. 41
      Gallicchio, E.; Xia, J.; Flynn, W. F.; Zhang, B.; Samlalsingh, S.; Mentes, A.; Levy, R. M. Asynchronous replica exchange software for grid and heterogeneous computing. Comput. Phys. Commun. 2015, 196, 236246,  DOI: 10.1016/j.cpc.2015.06.010
    42. 42
      Shirts, M. R.; Chodera, J. D. Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 2008, 129, 124105,  DOI: 10.1063/1.2978177
    43. 43
      Tan, Z.; Gallicchio, E.; Lapelosa, M.; Levy, R. M. Theory of binless multi-state free energy estimation with applications to protein-ligand binding. J. Chem. Phys. 2012, 136, 144102,  DOI: 10.1063/1.3701175
    44. 44
      Eastman, P.; Swails, J.; Chodera, J. D.; McGibbon, R. T.; Zhao, Y.; Beauchamp, K. A.; Wang, L.-P.; Simmonett, A. C.; Harrigan, M. P.; Stern, C. D.; Wiewiora, R. P.; Brooks, B. R.; Pande, V. S. OpenMM 7: Rapid development of high performance algorithms for molecular dynamics. PLoS Comput. Biol. 2017, 13, e1005659,  DOI: 10.1371/journal.pcbi.1005659
    45. 45
      Azimi, S.; Khuttan, S.; Wu, J. Z.; Deng, N.; Gallicchio, E. Application of the Alchemical Transfer and Potential of Mean Force Methods to the SAMPL8 Cavitand Host-Guest Blinded Challenge. arXiv.org , 2021, 2107.05155.
    46. 46
      Pal, R. K.; Gallicchio, E. Perturbation potentials to overcome order/disorder transitions in alchemical binding free energy calculations. J. Chem. Phys. 2019, 151, 124116,  DOI: 10.1063/1.5123154
    47. 47
      Khuttan, S.; Azimi, S.; Wu, J. Z.; Gallicchio, E. Alchemical Transformations for Concerted Hydration Free Energy Estimation with Explicit Solvation. J. Chem. Phys. 2021, 154, 054103,  DOI: 10.1063/5.0036944
    48. 48
      Gallicchio, E.; Levy, R. M. Recent Theoretical and Computational Advances for Modeling Protein-Ligand Binding Affinities. Adv. Prot. Chem. Struct. Biol. 2011, 85, 2780,  DOI: 10.1016/B978-0-12-386485-7.00002-8
    49. 49
      Gilson, M. K.; Given, J. A.; Bush, B. L.; McCammon, J. A. The Statistical-Thermodynamic Basis for Computation of Binding Affinities: A Critical Review. Biophys. J. 1997, 72, 10471069,  DOI: 10.1016/S0006-3495(97)78756-3
    50. 50
      Roux, B.; Simonson, T. Implicit Solvent Models. Biophys. Chem. 1999, 78, 120,  DOI: 10.1016/S0301-4622(98)00226-9
    51. 51
      Boresch, S.; Tettinger, F.; Leitgeb, M.; Karplus, M. Absolute binding free energies: A quantitative approach for their calculation. J. Phys. Chem. B 2003, 107, 95359551,  DOI: 10.1021/jp0217839
    52. 52
      Chipot; ; Pohorille, Eds. Free Energy Calculations. Theory and Applications in Chemistry and Biology; Springer Series in Chemical Physics; Springer: Berlin Heidelberg, 2007.
    53. 53
      The Single-Decoupling Plugin for OpenMM. https://github.com/Gallicchio-Lab/openmm_sdm_plugin (accessed September 12, 2021).
    54. 54
      SAMPL8 host–guest GDCC challenge. https://github.com/samplchallenges/SAMPL8/tree/master/host_guest/GDCC (accessed September 12, 2021).
    55. 55
      Norman, B. H.; Dodge, J. A.; Richardson, T. I.; Borromeo, P. S.; Lugar, C. W.; Jones, S. A.; Chen, K.; Wang, Y.; Durst, G. L.; Barr, R. J.; Montrose-Rafizadeh, C.; Osborne, H. E.; Amos, R. M.; Guo, S.; Boodhoo, A.; Krishnan, V. Benzopyrans are selective estrogen receptor β agonists with novel activity in models of benign prostatic hyperplasia. J. Med. Chem. 2006, 49, 61556157,  DOI: 10.1021/jm060491j
    56. 56
      SAMPL8 host–guest GDCC challenge submission 37. https://github.com/samplchallenges/SAMPL8/blob/master/host_guest/Analysis/Submissions/GDCC/GDCC-ATM.txt (accessed September 12, 2021).
    57. 57
      Kuntz, K. W.; Campbell, J. E.; Keilhack, H.; Pollock, R. M.; Knutson, S. K.; Porter-Scott, M.; Richon, V. M.; Sneeringer, C. J.; Wigle, T. J.; Allain, C. J.; Majer, C. R.; Moyer, M. P.; Copeland, R. A.; Chesworth, R. The importance of being me: magic methyls, methyltransferase inhibitors, and the discovery of tazemetostat. J. Med. Chem. 2016, 59, 15561564,  DOI: 10.1021/acs.jmedchem.5b01501
    58. 58
      Wang, J.; Wang, W.; Kollman, P. A.; Case, D. A. Automatic atom type and bond type perception in molecular mechanical calculations. J. Mol. Graphics Modell. 2006, 25, 247260,  DOI: 10.1016/j.jmgm.2005.12.005
    59. 59
      He, X.; Liu, S.; Lee, T.-S.; Ji, B.; Man, V. H.; York, D. M.; Wang, J. Fast, Accurate, and Reliable Protocols for Routine Calculations of Protein-Ligand Binding Affinities in Drug Design Projects Using AMBER GPU-TI with ff14SB/GAFF. ACS Omega 2020, 5, 46114619,  DOI: 10.1021/acsomega.9b04233
    60. 60
      Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.; Simmerling, C. ff14SB: improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theory Comput. 2015, 11, 36963713,  DOI: 10.1021/acs.jctc.5b00255
    61. 61
      Relative Binding Free Energy Calculations for Ligands with Diverse Scaffolds with the Alchemical Transfer Method, Simulation Input Files. https://github.com/Gallicchio-Lab/ATM-relative-binding-free-energy-paper (accessed September 12, 2021).
    62. 62
      The Asynchronous Replica Exchange Framework for OpenMM. https://github.com/Gallicchio-Lab/async_re-openmm (accessed September 12, 2021).
    63. 63
      Rizzi, A.; Murkli, S.; McNeill, J. N.; Yao, W.; Sullivan, M.; Gilson, M. K.; Chiu, M. W.; Isaacs, L.; Gibb, B. C.; Mobley, D. L.; Chodera, J. D. Overview of the SAMPL6 host-guest binding affinity prediction challenge. J. Comp.-Aided Mol. Des. 2018, 32, 937963,  DOI: 10.1007/s10822-018-0170-6
    64. 64
      Raman, E. P.; Paul, T. J.; Hayes, R. L.; Brooks, C. L., III Automated, accurate, and scalable relative protein-ligand binding free-energy calculations using lambda dynamics. J. Chem. Theory Comput. 2020, 16, 78957914,  DOI: 10.1021/acs.jctc.0c00830
    65. 65
      Woods, C. J.; Malaisree, M.; Hannongbua, S.; Mulholland, A. J. A water-swap reaction coordinate for the calculation of absolute protein-ligand binding free energies. J. Chem. Phys. 2011, 134, 054114,  DOI: 10.1063/1.3519057
    66. 66
      Procacci, P.; Macchiagodena, M. On the NS-DSSB unidirectional estimates in the SAMPL6 SAMPLing challenge. J. Comput.-Aided Mol. Des. 2021, 35, 1055,  DOI: 10.1007/s10822-021-00419-0
    67. 67
      Öhlknecht, C.; Lier, B.; Petrov, D.; Fuchs, J.; Oostenbrink, C. Correcting electrostatic artifacts due to net-charge changes in the calculation of ligand binding free energies. J. Comput. Chem. 2020, 41, 986999,  DOI: 10.1002/jcc.26143
    68. 68
      Pan, A. C.; Xu, H.; Palpant, T.; Shaw, D. E. Quantitative characterization of the binding and unbinding of millimolar drug fragments with molecular dynamics simulations. J. Chem. Theory Comput. 2017, 13, 33723377,  DOI: 10.1021/acs.jctc.7b00172
    69. 69
      Lin, Y.-L.; Aleksandrov, A.; Simonson, T.; Roux, B. An overview of electrostatic free energy computations for solutions and proteins. J. Chem. Theory Comput. 2014, 10, 26902709,  DOI: 10.1021/ct500195p
    70. 70
      The ATM Meta Force Plugin for OpenMM. https://github.com/Gallicchio-Lab/openmm-atmmetaforce-plugin (accessed September 12, 2021).
    71. 71
      Huang, J.; Lemkul, J. A.; Eastman, P. K.; MacKerell, A. D., Jr. Molecular dynamics simulations using the drude polarizable force field on GPUs with OpenMM: Implementation, validation, and benchmarks. J. Comput. Chem. 2018, 39, 16821689,  DOI: 10.1002/jcc.25339
    72. 72
      Gallicchio, E.; Deng, N.; He, P.; Wickstrom, L.; Perryman, A. L.; Santiago, D. N.; Forli, S.; Olson, A. J.; Levy, R. M. Virtual Screening of Integrase Inhibitors by Large Scale Binding Free Energy Calculations: the SAMPL4 Challenge. J. Comput.-Aided Mol. Des. 2014, 28, 475490,  DOI: 10.1007/s10822-014-9711-9
    73. 73
      Darden, T. A.; York, D. M.; Pedersen, L. G. Particle mesh Ewald: An NlogN method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 1008910092,  DOI: 10.1063/1.464397
    74. 74
      Kilburg, D.; Gallicchio, E. Analytical Model of the Free Energy of Alchemical Molecular Binding. J. Chem. Theory Comput. 2018, 14, 61836196,  DOI: 10.1021/acs.jctc.8b00967
    75. 75
      Gallicchio, E.; Lapelosa, M.; Levy, R. M. Binding Energy Distribution Analysis Method (BEDAM) for Estimation of Protein-Ligand Binding Affinities. J. Chem. Theory Comput. 2010, 6, 29612977,  DOI: 10.1021/ct1002913
    76. 76
      Riniker, S.; Christ, C. D.; Hansen, N.; Mark, A. E.; Nair, P. C.; van Gunsteren, W. F. Comparison of enveloping distribution sampling and thermodynamic integration to calculate binding free energies of phenylethanolamine N-methyltransferase inhibitors. J. Chem. Phys. 2011, 135, 024105,  DOI: 10.1063/1.3604534
    77. 77
      Deng, Y.; Roux, B. Computations of standard binding free energies with molecular dynamics simulations. J. Phys. Chem. B 2009, 113, 22342246,  DOI: 10.1021/jp807701h
    78. 78
      Deng, N.; Cui, D.; Zhang, B. W.; Xia, J.; Cruz, J.; Levy, R. Comparing alchemical and physical pathway methods for computing the absolute binding free energy of charged ligands. Phys. Chem. Chem. Phys. 2018, 20, 1708117092,  DOI: 10.1039/C8CP01524D
    79. 79
      Velez-Vega, C.; Gilson, M. K. Overcoming dissipation in the calculation of standard binding free energies by ligand extraction. J. Comput. Chem. 2013, 34, 23602371,  DOI: 10.1002/jcc.23398
    80. 80
      Hall, R.; Dixon, T.; Dickson, A. On calculating free energy differences using ensembles of transition paths. Frontiers Mol. Biosc. 2020, 7, 106,  DOI: 10.3389/fmolb.2020.00106
    81. 81
      Rizzi, A.; Jensen, T.; Slochower, D. R.; Aldeghi, M.; Gapsys, V.; Ntekoumes, D.; Bosisio, S.; Papadourakis, M.; Henriksen, N. M.; De Groot, B. L.; Cournia, Z.; Dickson, A.; Michel, J.; Gilson, M. K.; Shirts, M. R.; Mobley, D. L.; Chodera, J. D. The SAMPL6 SAMPLing challenge: Assessing the reliability and efficiency of binding free energy calculations. J. Comp. Aid. Mol. Des. 2020, 34, 601,  DOI: 10.1007/s10822-020-00290-5
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jcim.1c01129.

    • Review of theory underpinning the alchemical transfer method for absolute binding free energy estimation and derivation of the statistical mechanical extension for the relative binding free energy estimation with ligand alignment restraints (PDF)


    Terms & Conditions

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