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

Figure 1Loading Img
RETURN TO ISSUEPREVB: Biophysical and B...B: Biophysical and Biochemical Systems and ProcessesNEXT

Exploring the Minimum-Energy Pathways and Free-Energy Profiles of Enzymatic Reactions with QM/MM Calculations

  • Kiyoshi Yagi
    Kiyoshi Yagi
    Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
    More by Kiyoshi Yagi
  • Shingo Ito
    Shingo Ito
    Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
    More by Shingo Ito
  • , and 
  • Yuji Sugita*
    Yuji Sugita
    Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
    Computational Biophysics Research Team, RIKEN Center for Computational Science, 7-1-26 minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, Japan
    Laboratory for Biomolecular Function Simulation, RIKEN Center for Biosystems Dynamics Research, 1-6-5 minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, Japan
    *[email protected]
    More by Yuji Sugita
Cite this: J. Phys. Chem. B 2021, 125, 18, 4701–4713
Publication Date (Web):April 29, 2021
https://doi.org/10.1021/acs.jpcb.1c01862

Copyright © 2021 American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0.
  • Open Access

Article Views

2030

Altmetric

-

Citations

22
LEARN ABOUT THESE METRICS
PDF (4 MB)
Supporting Info (3)»

Abstract

Understanding molecular mechanisms of enzymatic reactions is of vital importance in biochemistry and biophysics. Here, we introduce new functions of hybrid quantum mechanical/molecular mechanical (QM/MM) calculations in the GENESIS program to compute the minimum-energy pathways (MEPs) and free-energy profiles of enzymatic reactions. For this purpose, an interface in GENESIS is developed to utilize a highly parallel electronic structure program, QSimulate-QM (https://qsimulate.com), calling it as a shared library from GENESIS. Second, algorithms to search the MEP are implemented, combining the string method (E et al. J. Chem. Phys. 2007, 126, 164103) with the energy minimization of the buffer MM region. The method implemented in GENESIS is applied to an enzyme, triosephosphate isomerase, which converts dihyroxyacetone phosphate to glyceraldehyde 3-phosphate in four proton-transfer processes. QM/MM-molecular dynamics simulations show performances of greater than 1 ns/day with the density functional tight binding (DFTB), and 10–30 ps/day with the hybrid density functional theory, B3LYP-D3. These performances allow us to compute not only MEP but also the potential of mean force (PMF) of the enzymatic reactions using the QM/MM calculations. The barrier height obtained as 13 kcal mol–1 with B3LYP-D3 in the QM/MM calculation is in agreement with the experimental results. The impact of conformational sampling in PMF calculations and the level of electronic structure calculations (DFTB vs B3LYP-D3) suggests reliable computational protocols for enzymatic reactions without high computational costs.

This publication is licensed under

CC-BY-NC-ND 4.0.
  • cc licence
  • by licence
  • nc licence
  • nd licence

 Special Issue

Published as part of The Journal of Physical Chemistry virtual special issue “Computational Advances in Protein Engineering and Enzyme Design”.

1. Introduction

ARTICLE SECTIONS
Jump To

The role of molecular simulations in life science has been increased rapidly and significantly. (1) The conformational dynamics of membrane proteins, (2) ribosomes, (3) and other biomolecular complexes (4,5) has been investigated in molecular dynamics (MD) simulations based on atomistic (6−9) or coarse-grained (CG) (10−13) molecular models. Biomolecular interactions, such as protein–protein (14,15) or protein–DNA/RNA interactions, (16) are also examined in detail with large-scale biomolecular simulations in the cellular environments (14,17) and viruses. (18,19) MD-special supercomputers like Anton/Anton2 (20) or MD-GRAPE (21) allow us to simulate slow atomic motions of proteins on time scales up to milliseconds. However, most MD simulations are performed on the basis of classical mechanics, relying on the so-called “molecular force fields”, (22−24) which can cause several limitations in the simulations in terms of their accuracy and applicability. The force-field accuracy has long been discussed in simulation communities. (25,26) It is difficult to keep a good balance between the conformational stability of globular proteins and flexible structures of intrinsically disordered proteins/regions (IDP/IDR) even in the latest updates of molecular force fields. (27,28) Divalent cations, namely, Mg2+ or Ca2+, often add another complexity to MD simulations. (29,30) MD simulations using a “molecular force field” cannot describe chemical reactions or enzymatic reactions, since the motion of electrons including bond breaking/formation is totally neglected in the simulations. (31−34)
To overcome those difficulties in MD simulations with “force fields”, a hybrid quantum mechanical/molecular mechanical (QM/MM) calculation was proposed, at first, by Warshal, Levitt, and Karplus more than 40 years ago. (35−37) The electronic structure calculation is applied to a small region where the most interesting chemical reaction happens, while the surrounding buffer regions containing a large number of atoms are treated with molecular force fields. Conventionally, QM/MM calculations were difficult to apply to many interesting biological systems, presumably, due to their high computational costs. The QM region had to be sufficiently small, just containing less than 50 atoms, and only very short-time dynamics (shorter than 100 ps in total) were available if ab initio electronic structure theory was used in the QM calculations. However, recent advances in simulation methodologies and growing computational resources in supercomputers and graphics processing unit (GPU) clusters are opening a new era in the QM/MM calculations. One of the promising future directions is the free-energy calculations using the hybrid QM/MM potential. Yang and co-workers developed methods to compute the free-energy and minimum free-energy path with QM/MM. (38−40) Hayashi and his colleagues decoupled the QM/MM free-energy optimization from atomistic MD simulations to simulate large conformational changes between the reactants, transition states, and products in enzymes or motor proteins. (41,42) Rosta and Hummer applied umbrella sampling (US) based on the QM/MM potentials to describe the free-energy profiles (or potential of mean forces (PMFs)) of adenosine triphosphate (ATP) hydrolysis in solution and in enzymes. (43−45) Hammes-Shiffer and co-workers have also applied the method to various enzymatic reactions. (46−50)
In recent QM/MM simulations, ab initio electronic structure calculations with the QM program are combined with high-performance MD programs, such as AMBER (51,52) and NAMD. (53) We have also developed highly parallel MD software, GENESIS (https://www.r-ccs.riken.jp/labs/cbrt/), for simulating chemical and biological applications. One of the key features in GENESIS is the extensive parallelization for massively parallel supercomputers (54) and GPU-parallel clusters. (55) On the basis of our new computational techniques, such as the midpoint cell method, (56) the volumetric decomposition of three-dimensional (3D) fast Fourier transform (FFT), (57) huge biological systems, such as the cytoplasm of Mycoplasma genetalium, (14) a small bacterial system, or an entire gene locus (GATA4) (16) were simulated on supercomputers like the K computer, Fugaku (both in RIKEN) and Trinity in Los Alamos National Laboratory. The former contains more than 100 million atoms, and the latter includes over 1 billion atoms of biomolecules as well as solvent and ions. The second key feature is the availability of enhanced conformational sampling algorithms for extending the conformational space of biomolecules and accelerating the convergence of free-energy calculations. Replica-exchange MD (REMD), (58,59) replica-exchange US (REUS), (60) generalized replica exchange with solute tempering (gREST), (61) Gaussian accelerated MD (GaMD), (62,63) free-energy perturbation (FEP), (64,65) the string method with mean forces, (66,67) and combinations of these methods are now available. The CHARMM-GUI developed by Im and his colleagues prepares input files of MD simulations and free-energy calculations. (68)
These two key features in GENESIS are also useful in the QM/MM simulation, which has been implemented recently. (69) The well-known QM software, namely, Gaussian, (70) Q-Chem, (71) TeraChem, (72) and DFTB+, (73) is available in GENESIS from the release of 1.5. We did not change the QM programs and just call each binary of the QM software from the QM/MM interface program in GENESIS. This simple approach is sufficient in many QM/MM calculations. However, overhead costs required in the system call and exchanges of information using input/output (I/O) files between GENESIS and QM software are non-negligible when we use DFTB+ in QM simulations. In our previous work, we performed a molecular vibrational analysis using the QM/MM program in GENESIS as well as SINDO, (74) which can introduce the anharmonic effect of molecular vibrations. It was applied to an enzyme, P450 nitric oxide reductase, with the NO molecule bound to a ferric heme. (69)
In this study, we develop two new functions of the QM/MM calculations in GENESIS: (1) the interface with highly parallelized QM software, QSimulate-QM, (75) and (2) a reaction path search algorithm using the string method. (76,77) Note that the original string method in GENESIS (67) can be used with QM/MM calculations as well to compute the minimum free-energy pathway (MFEP). However, the method requires QM/MM-MD simulations to obtain the mean force, which is costly in general. In the present study, we implement an alternative string method (the so-called zero-temperature version), which computes the minimum energy pathway (MEP) on the potential energy. In this method, the buffer MM region is energy minimized with the QM region replaced by atomic charges derived from QM calculation. The QM calculation is required only once in the iteration of string optimization, and thus the computational cost is greatly reduced. The developed method is demonstrated in a conversion reaction of dihyroxyacetone phosphate (DHAP) into glyceraldehyde 3-phosphate (GAP) catalyzed by an enzyme, triosephosphate isomerase (TIM). The mechanism for the reaction has been well-established in previous theoretical works. (38,39,78−82) The whole reaction involves four proton-transfer (PT) processes. The MEP of each process is obtained by the string method, and the free-energy profiles along the PT reactions are calculated using the US method based on the QM/MM MD simulations. The high-performance computations using QSimulate-QM allow us to perform such free-energy calculations of enzymatic reactions in TIM with reasonable computational costs. Different levels of electronic structure calculations suggest promising schemes that are applicable to many other enzymatic reactions based on the QM/MM simulations.

2. Method

ARTICLE SECTIONS
Jump To

2.1. Interface with QSimulate-QM

We have previously implemented a general interface with electronic structure programs for QM/MM calculations in GENESIS. (69) GENESIS sends the atomic coordinates and the MM charges to an electronic structure program and receives the energy and gradient upon return. The interface exchanges the information through external files. It (1) generates an input file of the electronic structure program, (2) invokes the electronic structure program using a system call function, and (3) reads output files of the electronic structure program. The advantage of such a scheme is that any electronic structure program can be readily adapted once the format of input/output files are known. The current version (ver. 1.6) supports DFTB+, Gaussian, Q-Chem, and TeraChem. However, the drawbacks of this scheme are that the file I/O is intensive and that the system call is not suitable for message passing interface (MPI) parallelized programs.
In this study, we developed an interface that directly combines GENESIS with QSimulate-QM through a library. QSimulate-QM has realized high-accuracy electronic structure calculations of large chemical systems via extensive parallelization. QSimulate-QM is first compiled to create a shared library (libqsimulate.so). Then, the QM/MM routine in GENESIS calls an entry function of QSimulate-QM included in the library to perform electronic structure calculations. Note that GENESIS and QSimulate-QM are written in Fortran and C++, respectively. The interface, therefore, implements a Fortran/C++ interlanguage call, which complies with standardized protocols in Fortran 2003 to be interoperable with C++. In this scheme, all the information is sent and received on memory without requiring the file I/O. GENESIS provides the atomic coordinates and MM charges, and QSimulate-QM returns the energy and gradient. The atomic charges of QM atoms derived from the intrinsic atomic orbitals (IAO) (83) are optionally returned. Furthermore, one of the arguments of the entry function is set to an MPI communicator, which specifies MPI processes assigned to QSimulate-QM. This specification makes it feasible to run multiple electronic structure calculations in parallel, for example, over replicas in REMD or REUS, images in the string method, and so on.
The current QM/MM implementation, not only for QSimulate-QM but for any others, assumes that the system is set up as a cluster system, that is, a nonperiodic system. The electronic structure of a QM system is calculated in the presence of all MM charges. In other words, the electrostatic interaction between electrons, nuclei, and MM charges are included in the electronic Hamiltonian. QM programs solve the electronic Schrödinger equation to obtain the QM energy, the nuclear gradients, and the response of MM charges to the electron density. The information is returned to GENESIS and used as forces to propagate the atomic position.

2.2. Reaction-Path Search

We implemented an algorithm to find the MEP. The procedure is outlined in the following.

Step 0. Setup

The system is divided into three layers: active atoms in a path search, buffer atoms, and fixed atoms. One of the natural choices of these layers is to set the active atoms to QM atoms, the buffer atoms to MM atoms in the vicinity of a QM region (e.g., within ∼8.0 Å), and the fixed atoms to other MM atoms in the outer region. The reaction path is represented by atomic coordinates in a number of discretized images. The definition of three layers, the number of images (Nimg), and the coordinates of images along an initial path must be provided in the input by users.

Step 1. Relax the Buffer Atoms

The first step in the calculation is to relax the structure of buffer atoms with active atoms (QM atoms) held fixed. In this step, the QM/MM interaction is calculated by replacing the QM atoms with atomic charges derived from the electronic structure calculations. The gradient correction is applied by adding the difference between the QM/MM gradient and the approximate one, as is commonly done in the micro-iteration scheme. (69,84)

Step 2. QM/MM Calculations

A single-point QM/MM calculation is performed to update the energy, gradient, and atomic charges.

Step 3. Update the Coordinates of Active Atoms

The coordinates of active atoms are updated in two steps based on the simplified string method by E et al. (77) At first, the image moves along the gradient
(1)
where n is an index of iterations, xin is the Cartesian coordinates of active atoms of the ith image in the nth iteration, V is the potential energy, and δt is a constant. Then, the images are reparameterized to be equidistant. Specifically, the path length is calculated as follows
(2)
and the coordinates, {xi*}, are interpolated as a function of path length, {si}, to construct an interpolation function, I(s). Then, the new coordinates are obtained as
(3)

Step 4. Check the Convergence

If the change in the energy profile and path length are both within threshold values, the calculation is terminated. Otherwise, the calculation goes back to the first step and relaxes the buffer atoms in the updated coordinates.
In the above procedure, the computational bottleneck usually comes from electronic structure calculations in step (2). Therefore, step (2) together with step (1) is parallelized over the image. One can calculate either all images or a subset of images in parallel by setting the number of MPI processes to Nimg × M (M is an integer) or a divisor of Nimg, respectively. The number of MPI processes per image is set to M in the former, while it is 1 in the latter.

2.3. Free-Energy Calculations

The free-energy profile of the enzymatic reaction is computed by an umbrella sampling (US) that uses QM/MM MD simulations after the MEP search. A set of interatomic distances, specified by the user, is used for collective variables (CVs). The window is generated at an equal distance along the MEP in the CV space by interpolation. Then, QM/MM-MD simulations are performed for each window with a bias potential, that is, harmonic restraints to CVs.
The analysis of MD trajectories gives the free-energy profile or PMF. The estimator f̂i is obtained by the multistate Bennett acceptance ratio (MBAR) method (85) as
(4)
where K is the number of windows, Uibias denotes the bias potential of the ith window, Nj is the number of snapshots in the jth window, and xjn is a sample from the nth snapshot and the jth window. Equation 4 is solved iteratively to obtain {f̂i}. Then, the PMF is obtained using MBAR
(5)
(6)
where Z is the partition function. In order to explore the free-energy profile along MEP, it is convenient to introduce pathCV, (67,86,87) which characterizes the component of tangent vector along the path
(7)
where rk is the CV at the kth window, and λ is a constant.
(8)

3. Computational Details

ARTICLE SECTIONS
Jump To

3.1. Modeling and Equilibration of a System

The atomistic system of TIM was constructed from an X-ray crystal structure (PDBID: 7TIM, resolution 1.90 Å). (88) The crystal structure is a dimer of TIM cocrystallized with an inhibitor, phosphoglycolohydroxamic acid (C2H6NO6P). The two TIM proteins were both included in the simulation system, because the reaction center was located at the interface of the dimer. The nitrogen atom of the inhibitor was replaced with a carbon atom to generate DHAP. All titratable residues were ionized at a neutral pH condition based on pKa values estimated by PROPKA 3.1. (89) An exception is Glu165 with pKa of 7.75, which was kept ionized because it is an acceptor of a proton from DHAP. The proton of His95 was added at the ε position. CHARMM-GUI (90,91) was used to add hydrogen atoms and to solvate the system in a box of water molecules (110 × 90 × 90 Å3). An initial equilibration was performed in a periodic boundary condition (PBC) for 300 ps by the conventional classical MD with positional restraints to heavy atoms of proteins and DHAP. Then, the system was trimmed to 15 Å around a TIM dimer to create a non-PBC system. The system was equilibrated for 2.6 ns by classical MD gradually reducing the force constants of restraint, followed by a further equilibration for 140 ps by QM/MM-MD using scc-DFTB. (92) The protocol of the equilibration is summarized in Tables S1 and S2. Two spherical boundary potentials with a radius of 35 Å were applied around the center of mass of each TIM. The nonbonding interaction was reduced to zero between 16 and 18 Å employing a switching function. The neighbor list was updated every 20 fs. The force field parameters of TIM and water were CHARMM36 (93) and TIP3P, (94) respectively, and those of DHAP were generated using the force field toolkit. (95) The time integration was done using the velocity-Verlet integrator with a time step of 2.0 fs, and the temperature was controlled at 300 K using the Bussi thermostat. (96) The bond length of covalent bonds involving hydrogen atoms was constrained using the RATTLE (97) and SETTLE (98) algorithms.

3.2. QM/MM Calculations

The QM region was set to DHAP and the side chain of Glu165 and His95, which consists of 35 QM atoms. The boundary between QM and MM regions is set between the Cα and Cβ atoms. The link hydrogen atoms were attached to the Cβ atoms excluding the charges of MM atoms that belong to the same group as the Cα atoms. The electronic structure calculations were mainly performed by density functional theory (DFT) at the level of B3LYP hybrid functionals (99,100) with a D3 version of Grimm’s dispersion (101) (B3LYP-D3) using Dunning’s aug-cc-pVDZ for the basis sets. (102) The PBE functional (103) with def2-SVP (104) basis sets was also employed for benchmark calculations. The number of basis sets was 378 and 623 with def2-SVP and aug-cc-pVDZ, respectively. The scc-DFTB calculations were performed using the 3OB parameter sets. (92) The QM/MM calculations were performed using a development version of GENESIS interfaced with QSimulate-QM.

3.3. MEP Search

The structure of a reactant (DHAP) was obtained with the QM/MM energy minimization from the initial one described in Section 3.1. The residues within 6.0 Å of DHAP were relaxed, whereas others were kept fixed. Then, the structure of a product was obtained in two steps. First, the structure was roughly optimized adding distant restraints to X–H bonds with the reference distance set to 1.7–2.0 Å for a dissociating bond and to 1.0 Å for a newly formed bond. Second, the structure was refined by restarting the energy minimization without restraint. In this step, the residues within 3.0 Å of DHAP were relaxed.
Once the reactant and product structures were obtained, the MEP between them was searched using the string method. The initial path was obtained by a linear interpolation of the reactant and product in Cartesian coordinates. The number of images was set to 16. The active atoms in a path search were the same as QM atoms (DHAP and side chains of His95 and Glu165), and the buffer atoms were set to residues within 6.0 Å of DHAP. The step size δt was set to 0.0005 Å. The relaxation of buffer atoms was performed using a limited memory version of Broyden-Fletcher-Goldfarb-Shanno (L-BFGS-B) (105−107) and a macro/micro-iteration scheme. (84) The IAO charges (83) and the self-consistent charges were used for the micro-iteration in B3LYP-D3 and DFTB3, respectively. Convergence thresholds of the energy profile and the path length were set to 0.01 kcal mol–1 and 0.01 Å, respectively.
The procedure was repeated for all four PT reactions to obtain the structures of DHAP, three intermediates, and GAP and the MEPs connecting them.

3.4. Free-Energy Calculations

The umbrella sampling was performed along the MEP predicted from the string-method calculations. The bond distances r(X-H) involved in the reaction are taken as the CVs. The CVs used for each proton-transfer process are illustrated in Figures S1a–S3a. The force constant was set to 100 kcal/mol Å2, and the windows were set with an interval of δr = 0.1 Å following the convention in the literature. (43−47,108) We checked that the probability distribution of each window has a sufficient overlap, as shown in Figures S1c–S3c. QM/MM-MD simulations were performed for 10 ps by DFTB3 and for 12 ps by B3LYP-D3. The hydrogen atoms (protons) involved in the reaction were unconstrained, and thus the time step of the integration was set to δt = 0.5 fs. The canonical ensemble (NVT) simulation was performed at T = 300 K with τT = 0.5 ps for the Bussi thermostat. The structure was sampled every 5 fs in the last 10 ps of the simulation. The trajectory data were analyzed using the MBAR and path CV analyses to obtain the PMF. The MBAR and PMF analyses were performed using the analysis tools of GENESIS.

4. Results

ARTICLE SECTIONS
Jump To

4.1. Performance

QM/MM calculations were performed for a water droplet (7014 atoms, 2338 water molecules) and TIM in solution (37 QM atoms and 37,182 MM atoms) to study the performance of GENESIS/QSimulate-QM. The timing data were obtained from MD simulations in an NVT condition with a time step of 2 fs for 100 steps (200 fs) and 5000 steps (10 ps) in DFT and DFTB, respectively.
In a water droplet system, water molecules within a radius r of a center were taken as a QM region, which was varied in size by changing r in a range of 3.0–6.0 Å (corresponding to 24–147 atoms). In Figure 1a, QM/MM-MD with DFTB3 shows a performance of more than 1 ns/day up to 114 atoms (or 380 electrons), which is achievable owing to the library interface developed in this work. The DFT calculations are an order of magnitude more expensive than DFTB3. Nonetheless, the performance is found to be more than 10 ps/day with PBE/def2-SVP and B3LYP-D3/aug-cc-pVDZ up to 96 and 45 atoms, respectively, using 2 nodes, and up to 147 and 66 atoms, respectively, using 8 nodes.

Figure 1

Figure 1. Performance of QM/MM-MD simulations combining GENESIS and QSimulate-QM. (a) Visualization of a water droplet and a QM region in the center (left). The timing data for a water droplet with increasing QM size using DFTB3, PBE/def2-SVP, and B3LYP/aug-cc-pVDZ. The total number of atoms is 7014. The computing node has two CPUs of Intel Xeon Gold 6148 (20 core 2.40 GHz), and 10 threads ×4MPI processes were assigned per node. (b) Visualization of TIM and an enlargement of the binding pocket (left). The timing data of TIM using 1 node (7 threads × 8 MPI processes) with various electronic structure methods (middle) and multiple nodes with B3LYP-D3/aug-cc-pVDZ (right). The number of QM and MM atoms are 35 and 37 182, respectively. The number of basis sets is 378 and 623 with def2-SVP and aug-cc-pVDZ, respectively. The computing node has two CPUs of Intel Xeon Platinum 8280 (28 core 2.70 GHz, AVX-512). The timing was measured in NVT condition with a time step of 2 fs.

The results of TIM are displayed in Figure 1b. QM/MM-MD with DFTB3 shows a performance of 1.42 ns/day. The breakdown has found that portions of 63 and 29% of the time are spent for the calculation of DFTB3 and a nonbonded MM interaction, respectively. Therefore, the cost of MM calculations is non-negligible in a system with more than 30 000 atoms.
Among the DFT results, the pure DFT with moderate basis sets (PBE/def2-SVP) exhibits the best performance of 36.9 ps/day. The performance of PBE/aug-cc-pVDZ is decreased to 20.8 ps/day, because the use of diffuse functions (aug-cc-pVDZ) not only increases the number of basis sets but also makes the self-consistent field (SCF) convergence slower. The hybrid DFT, B3LYP-D3/aug-cc-pVDZ, is the most time-consuming due to the need of the Hartree–Fock exchange. Nonetheless, as shown in Figure 1b, the performance of B3LYP-D3 scales well with respect to the number of nodes. The best performance is obtained as 31.0 ps/day using 16 nodes (28 threads ×32 MPI). Interestingly, the result of 7 threads × 8 MPI reaches the maximum at 4 nodes, and a further increase in nodes leads to a drastic decrease of the performance, indicating the importance of a good balance between the number of threads and MPI processes. These results show that QM/MM-MD simulations of picoseconds and nanoseconds order is feasible at the level of DFT and DFTB, respectively, using GENESIS/QSimulate-QM.

4.2. Proton-Transfer Reactions in TIM

The mechanism of the conversion of DHAP to GAP is illustrated in Figure 2. In the first step, the proton H31 is transferred from DHAP to Glu165 of TIM. It is notable that the proton transfer is coupled with an electron transfer, where the negative charge migrates from Glu165 to the carboxylic acid of DHAP, varying the C2–C3 bond from a single to double bond. The negative charge of O2 drives the subsequent PTs of HE2 from His95 and HO3 from the ligand. Finally, the H31 of Glu165 returns to the ligand at C2 to yield GAP.

Figure 2

Figure 2. Schematic illustration of the conversion of DHAP to GAP with Glu165 and His95 of TIM. Protons are indicated in blue, and the electrons as well as CC/CO double bonds are in red.

The five structures of the active site of TIM with the ligand, denoted I–V in the following, are shown in Figure 3. The X–H distances are listed in Table S3 of the Supporting Information. It is seen that the current procedure yields the intermediates and the product starting from the reactant. The energy minimization of each structure was converged in 20–50 iterations. These structures were obtained in 76.8 min using four computing nodes in total.

Figure 3

Figure 3. Structures of DHAP (I), intermediates (II–IV), and GAP (V) with His95 and Glu165 obtained by QM/MM calculations at the B3LYP-D3/aug-cc-pVDZ level of theory. The protons before and after the reaction are indicated in green and pink circles, respectively.

Figure 4 shows the convergence of the string-method calculations in the first PT step (I → II). Both the energy and the pathway rapidly change in the first 10 iterations. The change is less pronounced during 10–50 iterations, and the convergence is achieved at the 93rd iteration. The energy profile of I → II gives a barrier height of 15.5 kcal mol–1 and an endothermic reaction energy of 11.7 kcal mol–1.

Figure 4

Figure 4. Convergence of an MEP search in the first PT step (I → II) with the string-method calculations.

Although the pathway at the 50th iteration overlaps with the converged path around the transition state (TS) in Figure 4, notable changes are observed in the initial stage of the reaction. This stage corresponds to a rotation of a CH2 group, which makes the C3–H31 bond point toward the OE2 of Glu165 (see Movie S1). The path search in Cartesian coordinates enables to describe smooth changes in the structure space, thereby avoiding discontinuities caused by hidden coordinates.
Figures 5 and 6 compare the results of an MEP search in the whole PT processes obtained in the QM/MM calculations based on B3LYP-D3 and DFTB3. A stark difference is that the intermediate III, where His95 is deprotonated, is not a minimum but a TS in DFTB3. The error is ascribed to a known deficiency in DFTB3 to describe a deprotonated nitrogen due to the lack of d-type orbitals in the basis sets. (92) Furthermore, the energetics obtained by DFTB3 is 2–3 times larger than those obtained by B3LYP-D3 as listed in Table 1. For example, the relative energy of II is 11.7 and 28.0 kcal mol–1 in B3LYP-D3 and DFTB3, respectively. The proton affinity (PA) of DHAP at the level of B3LYP-D3 and DFTB3 is calculated to be different by 15.3 kcal mol–1 (Tables S4–S6), indicating that the error in the PA is not the primary cause. As shown in Figure 6, the N–H group of His95 is hydrogen-bonded with an anionic oxygen and a hydroxyl group of the ligand during II to IV. The hydrogen-bond network presumably leads the nitrogen atom toward a deprotonated state and incurs the error in DFTB3. This is mirrored in the product state (V), where GAP is remote from His95. The relative energy is obtained with a reasonable agreement as 8.0 and 12.0 kcal mol–1 using B3LYP-D3 and DFTB3, respectively.

Figure 5

Figure 5. Interatomic distances (left) and energetics (right) along MEPs obtained by QM/MM calculations at the level of (a) B3LYP-D3 and (b) DFTB3. The definition of r1r8 is given in Figure 2.

Figure 6

Figure 6. Structure of the reactive site along the MEP and the relevant intermolecular distances (in Å) obtained by B3LYP-D3/aug-cc-pVDZ. Those of DFTB3 are also given in parentheses. Note that DFTB3 yields III as a transition state, so that TS2 and TS3 are not characterized.

Table 1. Relative Energy of Minimum and TS Structures Relative to the Reactant (in kcal mol–1)
  I TS1 II TS2 III TS3 IV TS4 V
B3LYP-D3              
MEP 0.0 15.5 11.7 18.5 17.2 18.7 11.7 13.1 8.0
PMF 0.0 11.8 7.0 13.2 11.8 13.3 5.9 9.3 3.4
DFTB3              
MEP 0.0 31.1 28.0 51.6 49.1 51.7 32.0 33.3 12.0
PMF 0.0 25.6 20.2 50.7 50.5 51.6 33.5 35.7 15.5
Despite the large errors in the energetics, the structural change along the MEP of DFTB3 is found to be similar to that of B3LYP-D3 as shown in the left panel of Figure 5a,b. The difference in the motion of the MEP obtained by B3LYP-D3 and DFTB3 (Movies S1 and S2, respectively) is hardly discernible. The result suggests the possibility of a multilevel approach where the MEP obtained by DFTB may be used to correct the energetics using DFT.
The umbrella sampling simulations were performed for the PT reactions using the distances r1r8, shown in Figure 2 as CVs. The details of the CVs, windows, and the probability distribution of path CV are given in Figures S1–S3. The PMFs obtained by B3LYP-D3 and DFTB3 are shown as a function of Path CV in Figure 7. DFTB3 gives the free-energy barrier of ∼50 kcal mol–1, nearly 4 times larger compared to that of B3LYP-D3. The comparison of MEP and PMF summarized in Table 1 indicates that the entropic effect on the energy is less than 10 kcal mol–1. The free-energy barrier is obtained as 13 kcal mol–1 by B3LYP-D3 in good agreement with the experimental result. (109)

Figure 7

Figure 7. PMF as a function of Path CV obtained by B3LYP-D3 (left) and DFTB3 (right).

Once the weight of each snapshot [eq 6] is obtained from MBAR, it can be used to derive the PMF in arbitrary coordinates. As an example, two-dimensional (2D) PMFs are plotted in Figure 8 as a function of various bond-forming/dissociating coordinates. The contour plots show that the configurations relevant for the reaction are sampled well. It is notable in Figure 8b that the minima around III is shallow and widely spread. The result indicates that two protons (HE2 and HO3) and a deprotonated His95 share an excess electron and stabilize the reactive motion to form a plateau potential. The PMF forms a channel in a diagonal direction, suggesting a concerted double proton transfer. Interestingly, the corresponding figure obtained by DFTB3 shown in Figure S4 shows a channel parallel to the reaction coordinates suggesting a stepwise mechanism.

Figure 8

Figure 8. 2D PMF of PT reactions obtained by QM/MM calculations at the B3LYP-D3/aug-cc-pVDZ level.

5. Discussion and Perspectives

ARTICLE SECTIONS
Jump To

In this study, the minimum-energy pathways and the free-energy profiles of enzymatic reactions are predicted using the hybrid QM/MM calculations implemented in GENESIS software. The benchmark calculations using four different electronic structure theories, namely, DFTB3, PBE/def2-SVP, PBE/aug-cc-pVDZ, and B3LYP-D3/aug-cc-pVDZ, have revealed efficient computations using QSimulate-QM as the electronic structure calculations. Since the precompiled binary of QSimulate-QM is linked as a shared library from the QM/MM interface program in GENESIS, we can avoid the large overhead that is required to exchange the energy and gradient information through file I/O between GENESIS and externally coupled electronic structure programs. The direct combination between GENESIS and QSimulate-QM in this way is efficient, in particular, for the DFTB3-level calculations, since the computational cost for the electronic structure calculations is rather comparable to the molecular mechanics calculations. The benchmark performance in the combination of QSimulate-QM/GENESIS suggests that the QM/MM-MD simulations become a useful research tool in chemical and biological problems without spending high computational resources. The free-energy calculations based on the QM/MM umbrella sampling or other enhanced sampling methods are important to take into account the entropic effects in the enzymatic reactions.
Comparisons of free-energy profiles obtained from different levels of the electronic structure calculations (DFTB3 vs B3LYP-D3) give us valuable lessons. To obtain sufficiently accurate energetics, ab initio electronic structure theories, such as B3LYP-D3, are inevitable. If we investigate the other chemical reactions with more complicated changes in the electronic structures, for instance, the reactions involving metal irons in a protein or a heme group, higher-level QM theories are necessary. Interestingly, the structures of TIM obtained in the QM/MM US simulations with DFTB3 and B3LYP-D3 are very similar to each other, suggesting that both of them are able to give us reliable local geometry of the chemical reactions. This suggests a possibility to perform QM/MM MD simulations and free-energy calculations using low-cost electronic structure theories, such as DFTB3. After the simulations, we can calibrate the free-energy profiles, computing more reliable potential energies of simulation snapshots using higher-level calculations (B3LYP-D3). Lee et al. (110−112) already examined the reliability and reproducibility of energetics using such schemes and pointed out the importance of overlapped distributions between high- and low-level QM or QM/MM MD simulations. Since the theoretical framework using MBAR as a reweighting method has been already established, it is important to increase our experiences in such calculations for examining the accuracy and reproducibility of the calibration using different levels of electronic calculations.
The QM/MM interface program in the GENESIS software has two unique features. In our previous paper, (69) a high-precision anharmonic vibrational analysis is shown to be possible in combination with the SINDO program. (74) For this application, a large number of the QM/MM single-point energy calculations on the grids near the energy minima are required to describe the potential energy surface. Accurate and high-performance calculations with QSimulate-QM are powerful for this purpose. Second, the QM/MM MD simulations with the speed of 1 ns/day are realized using DFTB3 as shown in this paper. This opens new possibilities to investigate the conformational dynamics of small- or medium-sized proteins on the time scales of 100 ns or longer using the QM/MM-MD simulations. The free-energy profiles of enzymatic reactions are now able to include the effect of conformational dynamics or an entropic effect more directly in the calculations. Enhanced conformational sampling algorithms and/or free-energy calculation methods in GENESIS, which have been originally implemented just for the classical MD simulations, are all available in the QM/MM calculations. Parallel MD simulation algorithms, such as REMD, gREST, and REUS, seem to be useful in the QM/MM-MD simulations with many replicas using massively parallel supercomputers like Fugaku.
The current version of GENESIS includes two different MD engines: ATDYN and SPDYN. ATDYN has been aimed to develop as a simple MD code with easy modifications for testing novel simulation algorithms. Therefore, CG and hybrid QM/MM models were easily implemented in ATDYN. However, if we aim to extend the simulations toward much larger systems, such as membrane proteins, ribosomes, DNA/RNA polymerases, and so on, the cost of molecular mechanics calculations are non-negligible. Also, the periodic boundary condition for QM/MM calculations (113−116) is more important to simulate membrane proteins and/or long DNA chains. In the future, we plan to implement the QM/MM calculations into SPDYN, allowing efficient spatial decomposition schemes for the parallelization of a whole simulation system and making the QM/MM calculations of huge biological systems possible. These approaches should be more important in understanding molecular mechanisms underlying the enzymatic reactions, designing biomacromolecules with new functions, and developing novel drugs that can control the regulation of functions of proteins related to diseases.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcb.1c01862.

  • Protocols of equilibration, interatomic distances, calculated total energies and proton affinities, Cartesian coordinates, details of umbrella sampling, 2D PMF obtained by DFTB3 (PDF)

  • Animation for the visualization of minimum-energy pathways from DHAP to GAP using QM/MM calculations at the level of B3LYP-D3/aug-cc-pVDZ (MP4)

  • Animation for the visualization of minimum-energy pathways from DHAP to GAP using QM/MM calculations at the levels of DFTB3 (MP4)

Terms & Conditions

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

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Yuji Sugita - Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, JapanComputational Biophysics Research Team, RIKEN Center for Computational Science, 7-1-26 minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, JapanLaboratory for Biomolecular Function Simulation, RIKEN Center for Biosystems Dynamics Research, 1-6-5 minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, JapanOrcidhttp://orcid.org/0000-0001-9738-9216 Email: [email protected]
  • Authors
    • Kiyoshi Yagi - Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, JapanOrcidhttp://orcid.org/0000-0003-1120-9355
    • Shingo Ito - Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
  • Notes
    The authors declare no competing financial interest.

    Data Availability. The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

ARTICLE SECTIONS
Jump To

This research is partially supported by RIKEN Pioneering Research Projects (Dynamic Structural Biology/Glycolipidologue Initiative) (to Y.S.), RIKEN Incentive Research Project (to K.Y.), Program for Promoting Research on the Supercomputer Fugaku (Biomolecular dynamics in a living cell/MD-driven Precision Medicine), MEXT/KAKENHI Grant No. JP19H05645 (to Y.S.) and JP20H02701 (to K.Y.). We used the computer system HOKUSAI, provided by the RIKEN Information System Division, and Oakbridge-CX and Octopus, provided by the University of Tokyo and Osaka University, respectively, through the HPCI System Research Project (hp200098). We thank Prof. Y. Matsunaga (Saitama Univ.) for his helpful comments on the free-energy calculations. Dr. T. Shiozaki (QSimulate Inc.) is acknowledged for helping us develop the GENESIS/QSimulate-QM interface program. Dr. Y. Akinaga (VINAS Co., Ltd.) is acknowledged for his help to implement the string-method routines into GENESIS.

References

ARTICLE SECTIONS
Jump To

This article references 116 other publications.

  1. 1
    Brini, E.; Simmerling, C.; Dill, K. Protein storytelling through physics. Science 2020, 370, eaaz3041,  DOI: 10.1126/science.aaz3041
  2. 2
    Muller, M. P.; Jiang, T.; Sun, C.; Lihan, M.; Pant, S.; Mahinthichaichan, P.; Trifan, A.; Tajkhorshid, E. Characterization of Lipid–Protein Interactions and Lipid-Mediated Modulation of Membrane Protein Function through Molecular Simulation. Chem. Rev. 2019, 119, 60866161,  DOI: 10.1021/acs.chemrev.8b00608
  3. 3
    Whitford, P. C.; Geggier, P.; Altman, R. B.; Blanchard, S. C.; Onuchic, J. N.; Sanbonmatsu, K. Y. Accommodation of aminoacyl-tRNA into the ribosome involves reversible excursions along multiple pathways. RNA 2010, 16, 11961204,  DOI: 10.1261/rna.2035410
  4. 4
    Wang, B.; Opron, K.; Burton, Z. F.; Cukier, R. I.; Feig, M. Five checkpoints maintaining the fidelity of transcription by RNA polymerases in structural and energetic details. Nucleic Acids Res. 2015, 43, 11331146,  DOI: 10.1093/nar/gku1370
  5. 5
    Da, L.-T.; Pardo-Avila, F.; Xu, L.; Silva, D.-A.; Zhang, L.; Gao, X.; Wang, D.; Huang, X. Bridge helix bending promotes RNA polymerase II backtracking through a critical and conserved threonine residue. Nat. Commun. 2016, 7, 11244,  DOI: 10.1038/ncomms11244
  6. 6
    Brooks, B. R. CHARMM: The biomolecular simulation program. J. Comput. Chem. 2009, 30, 15451614,  DOI: 10.1002/jcc.21287
  7. 7
    Lee, T.-S.; Cerutti, D. S.; Mermelstein, D.; Lin, C.; LeGrand, S.; Giese, T. J.; Roitberg, A.; Case, D. A.; Walker, R. C.; York, D. M. GPU-Accelerated Molecular Dynamics and Free Energy Methods in Amber18: Performance Enhancements and New Features. J. Chem. Inf. Model. 2018, 58, 20432050,  DOI: 10.1021/acs.jcim.8b00462
  8. 8
    Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 1925,  DOI: 10.1016/j.softx.2015.06.001
  9. 9
    Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kalé, L.; Schulten, K. Scalable Molecular Dynamics with NAMD. J. Comput. Chem. 2005, 26, 17811802,  DOI: 10.1002/jcc.20289
  10. 10
    Marrink, S. J.; Risselada, H. J.; Yefimov, S.; Tieleman, D. P.; de Vries, A. H. The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations. J. Phys. Chem. B 2007, 111, 78127824,  DOI: 10.1021/jp071097f
  11. 11
    Kenzaki, H.; Koga, N.; Hori, N.; Kanada, R.; Li, W.; Okazaki, K.-i.; Yao, X.-Q.; Takada, S. CafeMol: A Coarse-Grained Biomolecular Simulator for Simulating Proteins at Work. J. Chem. Theory Comput. 2011, 7, 19791989,  DOI: 10.1021/ct2001045
  12. 12
    Han, W.; Schulten, K. Further Optimization of a Hybrid United-Atom and Coarse-Grained Force Field for Folding Simulations: Improved Backbone Hydration and Interactions between Charged Side Chains. J. Chem. Theory Comput. 2012, 8, 44134424,  DOI: 10.1021/ct300696c
  13. 13
    Seo, S.; Shinoda, W. SPICA Force Field for Lipid Membranes: Domain Formation Induced by Cholesterol. J. Chem. Theory Comput. 2019, 15, 762774,  DOI: 10.1021/acs.jctc.8b00987
  14. 14
    Yu, I.; Mori, T.; Ando, T.; Harada, R.; Jung, J.; Sugita, Y.; Feig, M. Biomolecular Interactions Modulate Macromolecular Structure and Dynamics in Atomistic Model of a Bacterial Cytoplasm. eLife 2016, 5, 18457,  DOI: 10.7554/eLife.19274
  15. 15
    von Bülow, S.; Siggel, M.; Linke, M.; Hummer, G. Dynamic cluster formation determines viscosity and diffusion in dense protein solutions. Proc. Natl. Acad. Sci. U. S. A. 2019, 116, 98439852,  DOI: 10.1073/pnas.1817564116
  16. 16
    Jung, J.; Nishima, W.; Daniels, M.; Bascom, G.; Kobayashi, C.; Adedoyin, A.; Wall, M.; Lappala, A.; Phillips, D.; Fischer, W.; Tung, C. S.; Schlick, T.; Sugita, Y.; Sanbonmatsu, K. Y. Scaling molecular dynamics beyond 100,000 processor cores for large-scale biophysical simulations. J. Comput. Chem. 2019, 40, 19191930,  DOI: 10.1002/jcc.25840
  17. 17
    Singharoy, A. Atoms to Phenotypes: Molecular Design Principles of Cellular Energy Metabolism. Cell 2019, 179, 10981111,  DOI: 10.1016/j.cell.2019.10.021
  18. 18
    Zhao, G.; Perilla, J. R.; Yufenyuy, E. L.; Meng, X.; Chen, B.; Ning, J.; Ahn, J.; Gronenborn, A. M.; Schulten, K.; Aiken, C.; Zhang, P. Mature HIV-1 capsid structure by cryo-electron microscopy and all-atom molecular dynamics. Nature 2013, 497, 643646,  DOI: 10.1038/nature12162
  19. 19
    Casalino, L.; AI-Driven Multiscale Simulations Illuminate Mechanisms of SARS-CoV-2 Spike Dynamics. bioRxiv 2020,  DOI: 10.1101/2020.11.19.390187 .
  20. 20
    Shaw, D. E.; In Anton 2: Raising the Bar for Performance and Programmability in a Special-Purpose Molecular Dynamics Supercomputer, SC ’14: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis, Nov 16–21, 2014; IEEE, 2014; pp 4153. DOI: 10.1109/SC.2014.9
  21. 21
    Ohmura, I.; Morimoto, G.; Ohno, Y.; Hasegawa, A.; Taiji, M. MDGRAPE-4: a special-purpose computer system for molecular dynamics simulations. Philos. Trans. R. Soc., A 2014, 372, 20130387,  DOI: 10.1098/rsta.2013.0387
  22. 22
    Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J. Y.; Wang, L.; Lupyan, D.; Dahlgren, M. K.; Knight, J. L.; Kaus, J. W.; Cerutti, D. S.; Krilov, G.; Jorgensen, W. L.; Abel, R.; Friesner, R. A. OPLS3: A Force Field Providing Broad Coverage of Drug-like Small Molecules and Proteins. J. Chem. Theory Comput. 2016, 12, 281296,  DOI: 10.1021/acs.jctc.5b00864
  23. 23
    Huang, J.; Rauscher, S.; Nawrocki, G.; Ran, T.; Feig, M.; de Groot, B. L.; Grubmüller, H.; MacKerell, A. D. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat. Methods 2017, 14, 7173,  DOI: 10.1038/nmeth.4067
  24. 24
    Tian, C.; Kasavajhala, K.; Belfon, K. A. A.; Raguette, L.; Huang, H.; Migues, A. N.; Bickel, J.; Wang, Y.; Pincay, J.; Wu, Q.; Simmerling, C. ff19SB: Amino-Acid-Specific Protein Backbone Parameters Trained against Quantum Mechanics Energy Surfaces in Solution. J. Chem. Theory Comput. 2020, 16, 528552,  DOI: 10.1021/acs.jctc.9b00591
  25. 25
    Best, R. B.; Zheng, W.; Mittal, J. Balanced Protein–Water Interactions Improve Properties of Disordered Proteins and Non-Specific Protein Association. J. Chem. Theory Comput. 2014, 10, 51135124,  DOI: 10.1021/ct500569b
  26. 26
    van der Spoel, D. Systematic design of biomolecular force fields. Curr. Opin. Struct. Biol. 2021, 67, 1824,  DOI: 10.1016/j.sbi.2020.08.006
  27. 27
    Robustelli, P.; Piana, S.; Shaw, D. E. Developing a molecular dynamics force field for both folded and disordered protein states. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, E4758E4766,  DOI: 10.1073/pnas.1800690115
  28. 28
    Piana, S.; Robustelli, P.; Tan, D.; Chen, S.; Shaw, D. E. Development of a Force Field for the Simulation of Single-Chain Proteins and Protein–Protein Complexes. J. Chem. Theory Comput. 2020, 16, 24942507,  DOI: 10.1021/acs.jctc.9b00251
  29. 29
    Li, P.; Merz, K. M. Metal Ion Modeling Using Classical Mechanics. Chem. Rev. 2017, 117, 15641686,  DOI: 10.1021/acs.chemrev.6b00440
  30. 30
    Zhang, A.; Yu, H.; Liu, C.; Song, C. The Ca2+ permeation mechanism of the ryanodine receptor revealed by a multi-site ion model. Nat. Commun. 2020, 11, 922,  DOI: 10.1038/s41467-020-14573-w
  31. 31
    Senn, H. M.; Thiel, W. QM/MM Methods for Biomolecular Systems. Angew. Chem., Int. Ed. 2009, 48, 11981229,  DOI: 10.1002/anie.200802019
  32. 32
    van der Kamp, M. W.; Mulholland, A. J. Combined Quantum Mechanics/Molecular Mechanics (QM/MM) Methods in Computational Enzymology. Biochemistry 2013, 52, 27082728,  DOI: 10.1021/bi400215w
  33. 33
    Cui, Q. Perspective: Quantum Mechanical Methods in Biochemistry and Biophysics. J. Chem. Phys. 2016, 145, 140901,  DOI: 10.1063/1.4964410
  34. 34
    Cui, Q.; Pal, T.; Xie, L. Biomolecular QM/MM Simulations: What Are Some of the ″burning Issues″?. J. Phys. Chem. B 2021, 125, 689702,  DOI: 10.1021/acs.jpcb.0c09898
  35. 35
    Warshel, A.; Karplus, M. Calculation of Ground and Excited State Potential Surfaces of Conjugated Molecules. I. Formulation and Parametrization. J. Am. Chem. Soc. 1972, 94, 56125625,  DOI: 10.1021/ja00771a014
  36. 36
    Warshel, A.; Levitt, M. Theoretical Studies of Enzymic Reactions: Dielectric, Electrostatic and Steric Stabilization of the Carbonium Ion in the Reaction of Lysozyme. J. Mol. Biol. 1976, 103, 227249,  DOI: 10.1016/0022-2836(76)90311-9
  37. 37
    Field, M. J.; Bash, P. A.; Karplus, M. A Combined Quantum Mechanical and Molecular Mechanical Potential for Molecular Dynamics Simulations. J. Comput. Chem. 1990, 11, 700733,  DOI: 10.1002/jcc.540110605
  38. 38
    Zhang, Y.; Liu, H.; Yang, W. Free energy calculation on enzyme reactions with an efficient iterative procedure to determine minimum energy paths on a combined ab initio QM/MM potential energy surface. J. Chem. Phys. 2000, 112, 3483,  DOI: 10.1063/1.480503
  39. 39
    Hu, H.; Lu, Z.; Yang, W. QM/MM Minimum Free-Energy Path: Methodology and Application to Triosephosphate Isomerase. J. Chem. Theory Comput. 2007, 3, 390406,  DOI: 10.1021/ct600240y
  40. 40
    Hu, H.; Yang, W. Free Energies of Chemical Reactions in Solution and in Enzymes with Ab Initio Quantum Mechanics/Molecular Mechanics Methods. Annu. Rev. Phys. Chem. 2008, 59, 573601,  DOI: 10.1146/annurev.physchem.59.032607.093618
  41. 41
    Kosugi, T.; Hayashi, S. QM/MM Reweighting Free Energy SCF for Geometry Optimization on Extensive Free Energy Surface of Enzymatic Reaction. J. Chem. Theory Comput. 2012, 8, 322334,  DOI: 10.1021/ct2005837
  42. 42
    Hayashi, S.; Uchida, Y.; Hasegawa, T.; Higashi, M.; Kosugi, T.; Kamiya, M. QM/MM Geometry Optimization on Extensive Free-Energy Surfaces for Examination of Enzymatic Reactions and Design of Novel Functional Properties of Proteins. Annu. Rev. Phys. Chem. 2017, 68, 135154,  DOI: 10.1146/annurev-physchem-052516-050827
  43. 43
    Rosta, E.; Woodcock, H. L.; Brooks, B. R.; Hummer, G. Artificial reaction coordinate “tunneling” in free-energy calculations: The catalytic reaction of RNase H. J. Comput. Chem. 2009, 30, 16341641,  DOI: 10.1002/jcc.21312
  44. 44
    Rosta, E.; Nowotny, M.; Yang, W.; Hummer, G. Catalytic Mechanism of RNA Backbone Cleavage by Ribonuclease H from Quantum Mechanics/Molecular Mechanics Simulations. J. Am. Chem. Soc. 2011, 133, 89348941,  DOI: 10.1021/ja200173a
  45. 45
    Rosta, E.; Yang, W.; Hummer, G. Calcium Inhibition of Ribonuclease H1 Two-Metal Ion Catalysis. J. Am. Chem. Soc. 2014, 136, 31373144,  DOI: 10.1021/ja411408x
  46. 46
    Ganguly, A.; Thaplyal, P.; Rosta, E.; Bevilacqua, P. C.; Hammes-Schiffer, S. Quantum mechanical/molecular mechanical free energy simulations of the self-cleavage reaction in the hepatitis delta virus ribozyme. J. Am. Chem. Soc. 2014, 136, 14831496,  DOI: 10.1021/ja4104217
  47. 47
    Zhang, S.; Ganguly, A.; Goyal, P.; Bingaman, J. L.; Bevilacqua, P. C.; Hammes-Schiffer, S. Role of the active site guanine in the glmS ribozyme self-cleavage mechanism: Quantum mechanical/molecular mechanical free energy simulations. J. Am. Chem. Soc. 2015, 137, 784798,  DOI: 10.1021/ja510387y
  48. 48
    Li, P.; Soudackov, A. V.; Hammes-Schiffer, S. Fundamental Insights into Proton-Coupled Electron Transfer in Soybean Lipoxygenase from Quantum Mechanical/Molecular Mechanical Free Energy Simulations. J. Am. Chem. Soc. 2018, 140, 30683076,  DOI: 10.1021/jacs.7b13642
  49. 49
    Stevens, D. R.; Hammes-Schiffer, S. Exploring the Role of the Third Active Site Metal Ion in DNA Polymerase η with QM/MM Free Energy Simulations. J. Am. Chem. Soc. 2018, 140, 89658969,  DOI: 10.1021/jacs.8b05177
  50. 50
    Li, P.; Rangadurai, A.; Al-Hashimi, H. M.; Hammes-Schiffer, S. Environmental Effects on Guanine-Thymine Mispair Tautomerization Explored with Quantum Mechanical/Molecular Mechanical Free Energy Simulations. J. Am. Chem. Soc. 2020, 142, 1118311191,  DOI: 10.1021/jacs.0c03774
  51. 51
    Walker, R. C.; Crowley, M. F.; Case, D. A. The Implementation of a Fast and Accurate QM/MM Potential Method in Amber. J. Comput. Chem. 2008, 29, 10191031,  DOI: 10.1002/jcc.20857
  52. 52
    Götz, A. W.; Clark, M. A.; Walker, R. C. An Extensible Interface for QM/MM Molecular Dynamics Simulations with AMBER. J. Comput. Chem. 2014, 35, 95108,  DOI: 10.1002/jcc.23444
  53. 53
    Melo, M. C. R.; Bernardi, R. C.; Rudack, T.; Scheurer, M.; Riplinger, C.; Phillips, J. C.; Maia, J. D. C.; Rocha, G. B.; Ribeiro, J. V.; Stone, J. E.; Neese, F.; Schulten, K.; Luthey-Schulten, Z. NAMD goes quantum: an integrative suite for hybrid simulations. Nat. Methods 2018, 15, 351354,  DOI: 10.1038/nmeth.4638
  54. 54
    Jung, J.; Kobayashi, C.; Kasahara, K.; Tan, C.; Kuroda, A.; Minami, K.; Ishiduki, S.; Nishiki, T.; Inoue, H.; Ishikawa, Y.; Feig, M.; Sugita, Y. New parallel computing algorithm of molecular dynamics for extremely huge scale biological systems. J. Comput. Chem. 2021, 42, 231241,  DOI: 10.1002/jcc.26450
  55. 55
    Jung, J.; Naurse, A.; Kobayashi, C.; Sugita, Y. Graphics Processing Unit Acceleration and Parallelization of GENESIS for Large-Scale Molecular Dynamics Simulations. J. Chem. Theory Comput. 2016, 12, 49474958,  DOI: 10.1021/acs.jctc.6b00241
  56. 56
    Jung, J.; Mori, T.; Sugita, Y. Midpoint cell method for hybrid (MPI+OpenMP) parallelization of molecular dynamics simulations. J. Comput. Chem. 2014, 35, 10641072,  DOI: 10.1002/jcc.23591
  57. 57
    Jung, J.; Kobayashi, C.; Imamura, T.; Sugita, Y. Parallel implementation of 3D FFT with volumetric decomposition schemes for efficient molecular dynamics simulations. Comput. Phys. Commun. 2016, 200, 5765,  DOI: 10.1016/j.cpc.2015.10.024
  58. 58
    Sugita, Y.; Okamoto, Y. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett. 1999, 314, 141151,  DOI: 10.1016/S0009-2614(99)01123-9
  59. 59
    Sugita, Y.; Okamoto, Y. Replica-exchange multicanonical algorithm and multicanonical replica-exchange method for simulating systems with rough energy landscape. Chem. Phys. Lett. 2000, 329, 261270,  DOI: 10.1016/S0009-2614(00)00999-4
  60. 60
    Sugita, Y.; Kitao, A.; Okamoto, Y. Multidimensional replica-exchange method for free-energy calculations. J. Chem. Phys. 2000, 113, 60426051,  DOI: 10.1063/1.1308516
  61. 61
    Kamiya, M.; Sugita, Y. Flexible selection of the solute region in replica exchange with solute tempering: Application to protein-folding simulations. J. Chem. Phys. 2018, 149, 72304,  DOI: 10.1063/1.5016222
  62. 62
    Miao, Y.; Feher, V. A.; McCammon, J. A. Gaussian Accelerated Molecular Dynamics: Unconstrained Enhanced Sampling and Free Energy Calculation. J. Chem. Theory Comput. 2015, 11, 35843595,  DOI: 10.1021/acs.jctc.5b00436
  63. 63
    Oshima, H.; Re, S.; Sugita, Y. Replica-Exchange Umbrella Sampling Combined with Gaussian Accelerated Molecular Dynamics for Free-Energy Calculation of Biomolecules. J. Chem. Theory Comput. 2019, 15, 51995208,  DOI: 10.1021/acs.jctc.9b00761
  64. 64
    Tembre, B. L.; Mc Cammon, J. A. Ligand-receptor interactions. Comput. Chem. 1984, 8, 281283,  DOI: 10.1016/0097-8485(84)85020-2
  65. 65
    Oshima, H.; Re, S.; Sugita, Y. Prediction of Protein-Ligand Binding Pose and Affinity Using the gREST+FEP Method. J. Chem. Inf. Model. 2020, 60, 53825394,  DOI: 10.1021/acs.jcim.0c00338
  66. 66
    Maragliano, L.; Fischer, A.; Vanden-Eijnden, E.; Ciccotti, G. String method in collective variables: Minimum free energy paths and isocommittor surfaces. J. Chem. Phys. 2006, 125, 24106,  DOI: 10.1063/1.2212942
  67. 67
    Matsunaga, Y.; Komuro, Y.; Kobayashi, C.; Jung, J.; Mori, T.; Sugita, Y. Dimensionality of Collective Variables for Describing Conformational Changes of a Multi-Domain Protein. J. Phys. Chem. Lett. 2016, 7, 14461451,  DOI: 10.1021/acs.jpclett.6b00317
  68. 68
    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
  69. 69
    Yagi, K.; Yamada, K.; Kobayashi, C.; Sugita, Y. Anharmonic Vibrational Analysis of Biomolecules and Solvated Molecules Using Hybrid QM/MM Computations. J. Chem. Theory Comput. 2019, 15, 19241938,  DOI: 10.1021/acs.jctc.8b01193
  70. 70
    Frisch, M. J.; Gaussian 16, rev. C.01; Gaussian, Inc.: Wallingford, CT, 2016.
  71. 71
    Shao, Y. Advances in molecular quantum chemistry contained in the Q-Chem 4 program package. Mol. Phys. 2015, 113, 184215,  DOI: 10.1080/00268976.2014.952696
  72. 72
    Seritan, S.; Bannwarth, C.; Fales, B. S.; Hohenstein, E. G.; Isborn, C. M.; Kokkila-Schumacher, S. I. L.; Li, X.; Liu, F.; Luehr, N.; Snyder, J. W., Jr; Song, C.; Titov, A. V.; Ufimtsev, I. S.; Wang, L.-P.; Martínez, T. J. TeraChem: A graphical processing unit-accelerated electronic structure package for large-scale ab initio molecular dynamics. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2021, 11, e1494,  DOI: 10.1002/wcms.1494
  73. 73
    Hourahine, B. DFTB+, a software package for efficient approximate density functional theory based atomistic simulations. J. Chem. Phys. 2020, 152, 124101,  DOI: 10.1063/1.5143190
  74. 74
    Yagi, K. SINDO 4.0 beta. 2020, https://tms.riken.jp/en/research/software/sindo/.
  75. 75
    QSimulate-QM. Quantum Simulation Technologies, Inc. 2020, https://qsimulate.com/.
  76. 76
    E, W.; Ren, W.; Vanden-Eijnden, E. String method for the study of rare events. Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 66, 52301,  DOI: 10.1103/PhysRevB.66.052301
  77. 77
    E, W.; Ren, W.; Vanden-Eijnden, E. Simplified and improved string method for computing the minimum energy paths in barrier-crossing events. J. Chem. Phys. 2007, 126, 164103,  DOI: 10.1063/1.2720838
  78. 78
    Cui, Q.; Karplus, M. Quantum mechanics/molecular mechanics studies of triosephosphate isomerase-catalyzed reactions: Effect of geometry and tunneling on proton-transfer rate constants. J. Am. Chem. Soc. 2002, 124, 30933124,  DOI: 10.1021/ja0118439
  79. 79
    Zhang, X.; Harrison, D. H. T.; Cui, Q. Functional specificities of methylglyoxal synthase and triosephosphate isomerase: A combined QM/MM analysis. J. Am. Chem. Soc. 2002, 124, 1487114878,  DOI: 10.1021/ja027063x
  80. 80
    Cui, Q.; Karplus, M. Quantum mechanical/molecular mechanical studies of the triosephosphate isomerase-catalyzed reaction: Verification of methodology and analysis of reaction mechanisms. J. Phys. Chem. B 2002, 106, 17681798,  DOI: 10.1021/jp012659c
  81. 81
    Lennartz, C.; Schäfer, A.; Terstegen, F.; Thiel, W. Enzymatic reactions of triosephosphate isomerase: A theoretical calibration study. J. Phys. Chem. B 2002, 106, 17581767,  DOI: 10.1021/jp012658k
  82. 82
    Mendieta-Moreno, J. I.; Walker, R. C.; Lewis, J. P.; Gómez-Puertas, P.; Mendieta, J.; Ortega, J. FIREBALL/AMBER: An Efficient Local-Orbital DFT QM/MM Method for Biomolecular Systems. J. Chem. Theory Comput. 2014, 10, 21852193,  DOI: 10.1021/ct500033w
  83. 83
    Knizia, G. Intrinsic atomic orbitals: An unbiased bridge between quantum theory and chemical concepts. J. Chem. Theory Comput. 2013, 9, 48344843,  DOI: 10.1021/ct400687b
  84. 84
    Kästner, J.; Thiel, S.; Senn, H. M.; Sherwood, P.; Thiel, W. Exploiting QM/MM Capabilities in Geometry Optimization: A Microiterative Approach Using Electrostatic Embedding. J. Chem. Theory Comput. 2007, 3, 10641072,  DOI: 10.1021/ct600346p
  85. 85
    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
  86. 86
    Branduardi, D.; Gervasio, F. L.; Parrinello, M. From A to B in free energy space. J. Chem. Phys. 2007, 126, 054103,  DOI: 10.1063/1.2432340
  87. 87
    Bonomi, M.; Branduardi, D.; Bussi, G.; Camilloni, C.; Provasi, D.; Raiteri, P.; Donadio, D.; Marinelli, F.; Pietrucci, F.; Broglia, R. A.; Parrinello, M. PLUMED: A portable plugin for free-energy calculations with molecular dynamics. Comput. Phys. Commun. 2009, 180, 19611972,  DOI: 10.1016/j.cpc.2009.05.011
  88. 88
    Davenport, R. C.; Bash, P. A.; Seaton, B. A.; Karplus, M.; Petsko, G. A.; Ringe, D. Structure of the Triosephosphate Isomerase-Phosphoglycolohydroxamate Complex: An Analogue of the Intermediate on the Reaction Pathway. Biochemistry 1991, 30, 58215826,  DOI: 10.1021/bi00238a002
  89. 89
    Olsson, M. H. M.; Søndergaard, C. R.; Rostkowski, M.; Jensen, J. H. PROPKA3: Consistent Treatment of Internal and Surface Residues in Empirical pKa predictions. J. Chem. Theory Comput. 2011, 7, 525537,  DOI: 10.1021/ct100578z
  90. 90
    Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: A web-based graphical user interface for CHARMM. J. Comput. Chem. 2008, 29, 18591865,  DOI: 10.1002/jcc.20945
  91. 91
    Lee, J.; Cheng, X.; Swails, J. M.; Yeom, M. S.; Eastman, P. K.; Lemkul, J. A.; Wei, S.; Buckner, J.; Jeong, J. C.; Qi, Y.; Jo, S.; Pande, V. S.; Case, D. A.; Brooks, C. L.; MacKerell, A. D.; Klauda, J. B.; Im, W. CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field. J. Chem. Theory Comput. 2016, 12, 405413,  DOI: 10.1021/acs.jctc.5b00935
  92. 92
    Gaus, M.; Cui, Q.; Elstner, M. DFTB3: Extension of the Self-Consistent-Charge Density-Functional Tight-Binding Method (SCC-DFTB). J. Chem. Theory Comput. 2011, 7, 931948,  DOI: 10.1021/ct100684s
  93. 93
    Best, R. B.; Zhu, X.; Shim, J.; Lopes, P. E. M.; Mittal, J.; Feig, M.; Mackerell, A. D. Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone φ, ψ and Side-Chain χ1 and χ2 Dihedral Angles. J. Chem. Theory Comput. 2012, 8, 32573273,  DOI: 10.1021/ct300400x
  94. 94
    Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926935,  DOI: 10.1063/1.445869
  95. 95
    Mayne, C. G.; Saam, J.; Schulten, K.; Tajkhorshid, E.; Gumbart, J. C. Rapid parameterization of small molecules using the force field toolkit. J. Comput. Chem. 2013, 34, 27572770,  DOI: 10.1002/jcc.23422
  96. 96
    Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101,  DOI: 10.1063/1.2408420
  97. 97
    Andersen, H. C. Rattle: A “velocity” version of the shake algorithm for molecular dynamics calculations. J. Comput. Phys. 1983, 52, 2434,  DOI: 10.1016/0021-9991(83)90014-1
  98. 98
    Miyamoto, S.; Kollman, P. A. Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models. J. Comput. Chem. 1992, 13, 952962,  DOI: 10.1002/jcc.540130805
  99. 99
    Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785789,  DOI: 10.1103/PhysRevB.37.785
  100. 100
    Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 56485652,  DOI: 10.1063/1.464913
  101. 101
    Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104,  DOI: 10.1063/1.3382344
  102. 102
    Dunning, T. H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 10071023,  DOI: 10.1063/1.456153
  103. 103
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  104. 104
    Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 32973305,  DOI: 10.1039/b508541a
  105. 105
    Byrd, R. H.; Lu, P.; Nocedal, J.; Zhu, C. A Limited Memory Algorithm for Bound Constrained Optimization. SIAM J. Sci. Stat. Comp. 1995, 16, 11901208,  DOI: 10.1137/0916069
  106. 106
    Zhu, C.; Byrd, R.; Nocedal, J.; Morales, J. L. L-BFGS-B (ver. 3.0), http://users.iems.northwestern.edu/~nocedal/lbfgsb.html.
  107. 107
    Zhu, C.; Byrd, R. H.; Lu, P.; Nocedal, J. L-BFGS-B: Algorithm 778: L-BFGS-B, FORTRAN routines for large scale bound constrained optimization. ACM Transactions on Mathematical Software 1997, 23, 550560,  DOI: 10.1145/279232.279236
  108. 108
    Lence, E.; van der Kamp, M. W.; González-Bello, C.; Mulholland, A. J. QM/MM simulations identify the determinants of catalytic activity differences between type II dehydroquinase enzymes. Org. Biomol. Chem. 2018, 16, 44434455,  DOI: 10.1039/C8OB00066B
  109. 109
    Albery, W. J.; Knowles, J. R. Free-Energy Profile for the Reaction Catalyzed by Triosephosphate Isomerase. Biochemistry 1976, 15, 56275631,  DOI: 10.1021/bi00670a031
  110. 110
    Hudson, P. S.; Woodcock, H. L.; Boresch, S. Use of Nonequilibrium Work Methods to Compute Free Energy Differences Between Molecular Mechanical and Quantum Mechanical Representations of Molecular Systems. J. Phys. Chem. Lett. 2015, 6, 48504856,  DOI: 10.1021/acs.jpclett.5b02164
  111. 111
    Kearns, F. L.; Hudson, P. S.; Woodcock, H. L.; Boresch, S. Computing Converged Free Energy Differences between Levels of Theory via Nonequilibrium Work Methods: Challenges and Opportunities. J. Comput. Chem. 2017, 38, 13761388,  DOI: 10.1002/jcc.24706
  112. 112
    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
  113. 113
    Holden, Z. C.; Richard, R. M.; Herbert, J. M. Periodic boundary conditions for QM/MM calculations: Ewald summation for extended Gaussian basis sets. J. Chem. Phys. 2013, 139, 244108,  DOI: 10.1063/1.4850655
  114. 114
    Giese, T. J.; York, D. M. Ambient-Potential Composite Ewald Method for ab Initio Quantum Mechanical/Molecular Mechanical Molecular Dynamics Simulation. J. Chem. Theory Comput. 2016, 12, 26112632,  DOI: 10.1021/acs.jctc.6b00198
  115. 115
    Vasilevskaya, T.; Thiel, W. Periodic Boundary Conditions in QM/MM Calculations: Implementation and Tests. J. Chem. Theory Comput. 2016, 12, 35613570,  DOI: 10.1021/acs.jctc.6b00269
  116. 116
    Kawashima, Y.; Ishimura, K.; Shiga, M. Ab initio quantum mechanics/molecular mechanics method with periodic boundaries employing Ewald summation technique to electron-charge interaction: Treatment of the surface-dipole term. J. Chem. Phys. 2019, 150, 124103,  DOI: 10.1063/1.5048451

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 22 publications.

  1. Maximilian C. Pöverlein, Andreas Hulm, Johannes C. B. Dietschreit, Jörg Kussmann, Christian Ochsenfeld, Ville R. I. Kaila. QM/MM Free Energy Calculations of Long-Range Biological Protonation Dynamics by Adaptive and Focused Sampling. Journal of Chemical Theory and Computation 2024, Article ASAP.
  2. 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
  3. 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
  4. Kwangho Nam, Yihan Shao, Dan T. Major, Magnus Wolf-Watz. Perspectives on Computational Enzyme Modeling: From Mechanisms to Design and Drug Development. ACS Omega 2024, 9 (7) , 7393-7412. https://doi.org/10.1021/acsomega.3c09084
  5. Rajwinder Kaur, Dylan J. Nikkel, Mohamed M. Aboelnga, Stacey D. Wetmore. The Impact of DFT Functional, Cluster Model Size, and Implicit Solvation on the Structural Description of Single-Metal-Mediated DNA Phosphodiester Bond Cleavage: The Case Study of APE1. The Journal of Physical Chemistry B 2022, 126 (50) , 10672-10683. https://doi.org/10.1021/acs.jpcb.2c06756
  6. Kohei Sato, Ryo Sasaki, Ryoto Matsuda, Mayuko Nakagawa, Toru Ekimoto, Tsutomu Yamane, Mitsunori Ikeguchi, Kazuhito V. Tabata, Hiroyuki Noji, Kazushi Kinbara. Correction to “Supramolecular Mechanosensitive Potassium Channel Formed by Fluorinated Amphiphilic Cyclophane”. Journal of the American Chemical Society 2022, 144 (30) , 13983-13984. https://doi.org/10.1021/jacs.2c07339
  7. Zilin Song, Francesco Trozzi, Hao Tian, Chao Yin, Peng Tao. Mechanistic Insights into Enzyme Catalysis from Explaining Machine-Learned Quantum Mechanical and Molecular Mechanical Minimum Energy Pathways. ACS Physical Chemistry Au 2022, 2 (4) , 316-330. https://doi.org/10.1021/acsphyschemau.2c00005
  8. Sekhar Talluri. Engineering and Design of Programmable Genome Editors. The Journal of Physical Chemistry B 2022, 126 (28) , 5140-5150. https://doi.org/10.1021/acs.jpcb.2c03761
  9. Shingo Ito, Kiyoshi Yagi, Yuji Sugita. Computational Analysis on the Allostery of Tryptophan Synthase: Relationship between α/β-Ligand Binding and Distal Domain Closure. The Journal of Physical Chemistry B 2022, 126 (17) , 3300-3308. https://doi.org/10.1021/acs.jpcb.2c01556
  10. Darren Demapan, Jörg Kussmann, Christian Ochsenfeld, Qiang Cui. Factors That Determine the Variation of Equilibrium and Kinetic Properties of QM/MM Enzyme Simulations: QM Region, Conformation, and Boundary Condition. Journal of Chemical Theory and Computation 2022, 18 (4) , 2530-2542. https://doi.org/10.1021/acs.jctc.1c00714
  11. Etienne Derat, Shina Caroline Lynn Kamerlin. Computational Advances in Protein Engineering and Enzyme Design. The Journal of Physical Chemistry B 2022, 126 (13) , 2449-2451. https://doi.org/10.1021/acs.jpcb.2c01198
  12. Qiang Shao, Muya Xiong, Jiameng Li, Hangchen Hu, Haixia Su, Yechun Xu. Unraveling the catalytic mechanism of SARS-CoV-2 papain-like protease with allosteric modulation of C270 mutation using multiscale computational approaches. Chemical Science 2023, 14 (18) , 4681-4696. https://doi.org/10.1039/D3SC00166K
  13. T. Kubař, M. Elstner, Q. Cui. Hybrid Quantum Mechanical/Molecular Mechanical Methods For Studying Energy Transduction in Biomolecular Machines. Annual Review of Biophysics 2023, 52 (1) , 525-551. https://doi.org/10.1146/annurev-biophys-111622-091140
  14. Shingo Ito, Kiyoshi Yagi, Yuji Sugita. Allosteric regulation of β-reaction stage I in tryptophan synthase upon the α-ligand binding. The Journal of Chemical Physics 2023, 158 (11) https://doi.org/10.1063/5.0134117
  15. Yasuhiro Matsunaga, Motoshi Kamiya, Hiraku Oshima, Jaewoon Jung, Shingo Ito, Yuji Sugita. Use of multistate Bennett acceptance ratio method for free-energy calculations from enhanced sampling and free-energy perturbation. Biophysical Reviews 2022, 14 (6) , 1503-1512. https://doi.org/10.1007/s12551-022-01030-9
  16. Avijeet Kulshrestha, Sudeep N. Punnathanam, K. Ganapathy Ayappa. Finite temperature string method with umbrella sampling using path collective variables: application to secondary structure change in a protein. Soft Matter 2022, 18 (39) , 7593-7603. https://doi.org/10.1039/D2SM00888B
  17. Zilin Song, Peng Tao. Graph-learning guided mechanistic insights into imipenem hydrolysis in GES carbapenemases. Electronic Structure 2022, 4 (3) , 034001. https://doi.org/10.1088/2516-1075/ac7993
  18. Madushanka Manathunga, Andreas W. Götz, Kenneth M. Merz. Computer-aided drug design, quantum-mechanical methods for biological problems. Current Opinion in Structural Biology 2022, 75 , 102417. https://doi.org/10.1016/j.sbi.2022.102417
  19. Heidi Klem, Martin McCullagh, Robert S. Paton. Modeling Catalysis in Allosteric Enzymes: Capturing Conformational Consequences. Topics in Catalysis 2022, 65 (1-4) , 165-186. https://doi.org/10.1007/s11244-021-01521-1
  20. Vyshnavi Vennelakanti, Azadeh Nazemi, Rimsha Mehmood, Adam H. Steeves, Heather J. Kulik. Harder, better, faster, stronger: Large-scale QM and QM/MM for predictive modeling in enzymes and proteins. Current Opinion in Structural Biology 2022, 72 , 9-17. https://doi.org/10.1016/j.sbi.2021.07.004
  21. Kiyoshi Yagi, Suyong Re, Takaharu Mori, Yuji Sugita. Weight average approaches for predicting dynamical properties of biomolecules. Current Opinion in Structural Biology 2022, 72 , 88-94. https://doi.org/10.1016/j.sbi.2021.08.008
  22. Victor A. Lorenz-Fonfria, Kiyoshi Yagi, Shota Ito, Hideki Kandori. Retinal Vibrations in Bacteriorhodopsin are Mechanically Harmonic but Electrically Anharmonic: Evidence From Overtone and Combination Bands. Frontiers in Molecular Biosciences 2021, 8 https://doi.org/10.3389/fmolb.2021.749261
  • Abstract

    Figure 1

    Figure 1. Performance of QM/MM-MD simulations combining GENESIS and QSimulate-QM. (a) Visualization of a water droplet and a QM region in the center (left). The timing data for a water droplet with increasing QM size using DFTB3, PBE/def2-SVP, and B3LYP/aug-cc-pVDZ. The total number of atoms is 7014. The computing node has two CPUs of Intel Xeon Gold 6148 (20 core 2.40 GHz), and 10 threads ×4MPI processes were assigned per node. (b) Visualization of TIM and an enlargement of the binding pocket (left). The timing data of TIM using 1 node (7 threads × 8 MPI processes) with various electronic structure methods (middle) and multiple nodes with B3LYP-D3/aug-cc-pVDZ (right). The number of QM and MM atoms are 35 and 37 182, respectively. The number of basis sets is 378 and 623 with def2-SVP and aug-cc-pVDZ, respectively. The computing node has two CPUs of Intel Xeon Platinum 8280 (28 core 2.70 GHz, AVX-512). The timing was measured in NVT condition with a time step of 2 fs.

    Figure 2

    Figure 2. Schematic illustration of the conversion of DHAP to GAP with Glu165 and His95 of TIM. Protons are indicated in blue, and the electrons as well as CC/CO double bonds are in red.

    Figure 3

    Figure 3. Structures of DHAP (I), intermediates (II–IV), and GAP (V) with His95 and Glu165 obtained by QM/MM calculations at the B3LYP-D3/aug-cc-pVDZ level of theory. The protons before and after the reaction are indicated in green and pink circles, respectively.

    Figure 4

    Figure 4. Convergence of an MEP search in the first PT step (I → II) with the string-method calculations.

    Figure 5

    Figure 5. Interatomic distances (left) and energetics (right) along MEPs obtained by QM/MM calculations at the level of (a) B3LYP-D3 and (b) DFTB3. The definition of r1r8 is given in Figure 2.

    Figure 6

    Figure 6. Structure of the reactive site along the MEP and the relevant intermolecular distances (in Å) obtained by B3LYP-D3/aug-cc-pVDZ. Those of DFTB3 are also given in parentheses. Note that DFTB3 yields III as a transition state, so that TS2 and TS3 are not characterized.

    Figure 7

    Figure 7. PMF as a function of Path CV obtained by B3LYP-D3 (left) and DFTB3 (right).

    Figure 8

    Figure 8. 2D PMF of PT reactions obtained by QM/MM calculations at the B3LYP-D3/aug-cc-pVDZ level.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 116 other publications.

    1. 1
      Brini, E.; Simmerling, C.; Dill, K. Protein storytelling through physics. Science 2020, 370, eaaz3041,  DOI: 10.1126/science.aaz3041
    2. 2
      Muller, M. P.; Jiang, T.; Sun, C.; Lihan, M.; Pant, S.; Mahinthichaichan, P.; Trifan, A.; Tajkhorshid, E. Characterization of Lipid–Protein Interactions and Lipid-Mediated Modulation of Membrane Protein Function through Molecular Simulation. Chem. Rev. 2019, 119, 60866161,  DOI: 10.1021/acs.chemrev.8b00608
    3. 3
      Whitford, P. C.; Geggier, P.; Altman, R. B.; Blanchard, S. C.; Onuchic, J. N.; Sanbonmatsu, K. Y. Accommodation of aminoacyl-tRNA into the ribosome involves reversible excursions along multiple pathways. RNA 2010, 16, 11961204,  DOI: 10.1261/rna.2035410
    4. 4
      Wang, B.; Opron, K.; Burton, Z. F.; Cukier, R. I.; Feig, M. Five checkpoints maintaining the fidelity of transcription by RNA polymerases in structural and energetic details. Nucleic Acids Res. 2015, 43, 11331146,  DOI: 10.1093/nar/gku1370
    5. 5
      Da, L.-T.; Pardo-Avila, F.; Xu, L.; Silva, D.-A.; Zhang, L.; Gao, X.; Wang, D.; Huang, X. Bridge helix bending promotes RNA polymerase II backtracking through a critical and conserved threonine residue. Nat. Commun. 2016, 7, 11244,  DOI: 10.1038/ncomms11244
    6. 6
      Brooks, B. R. CHARMM: The biomolecular simulation program. J. Comput. Chem. 2009, 30, 15451614,  DOI: 10.1002/jcc.21287
    7. 7
      Lee, T.-S.; Cerutti, D. S.; Mermelstein, D.; Lin, C.; LeGrand, S.; Giese, T. J.; Roitberg, A.; Case, D. A.; Walker, R. C.; York, D. M. GPU-Accelerated Molecular Dynamics and Free Energy Methods in Amber18: Performance Enhancements and New Features. J. Chem. Inf. Model. 2018, 58, 20432050,  DOI: 10.1021/acs.jcim.8b00462
    8. 8
      Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 1925,  DOI: 10.1016/j.softx.2015.06.001
    9. 9
      Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kalé, L.; Schulten, K. Scalable Molecular Dynamics with NAMD. J. Comput. Chem. 2005, 26, 17811802,  DOI: 10.1002/jcc.20289
    10. 10
      Marrink, S. J.; Risselada, H. J.; Yefimov, S.; Tieleman, D. P.; de Vries, A. H. The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations. J. Phys. Chem. B 2007, 111, 78127824,  DOI: 10.1021/jp071097f
    11. 11
      Kenzaki, H.; Koga, N.; Hori, N.; Kanada, R.; Li, W.; Okazaki, K.-i.; Yao, X.-Q.; Takada, S. CafeMol: A Coarse-Grained Biomolecular Simulator for Simulating Proteins at Work. J. Chem. Theory Comput. 2011, 7, 19791989,  DOI: 10.1021/ct2001045
    12. 12
      Han, W.; Schulten, K. Further Optimization of a Hybrid United-Atom and Coarse-Grained Force Field for Folding Simulations: Improved Backbone Hydration and Interactions between Charged Side Chains. J. Chem. Theory Comput. 2012, 8, 44134424,  DOI: 10.1021/ct300696c
    13. 13
      Seo, S.; Shinoda, W. SPICA Force Field for Lipid Membranes: Domain Formation Induced by Cholesterol. J. Chem. Theory Comput. 2019, 15, 762774,  DOI: 10.1021/acs.jctc.8b00987
    14. 14
      Yu, I.; Mori, T.; Ando, T.; Harada, R.; Jung, J.; Sugita, Y.; Feig, M. Biomolecular Interactions Modulate Macromolecular Structure and Dynamics in Atomistic Model of a Bacterial Cytoplasm. eLife 2016, 5, 18457,  DOI: 10.7554/eLife.19274
    15. 15
      von Bülow, S.; Siggel, M.; Linke, M.; Hummer, G. Dynamic cluster formation determines viscosity and diffusion in dense protein solutions. Proc. Natl. Acad. Sci. U. S. A. 2019, 116, 98439852,  DOI: 10.1073/pnas.1817564116
    16. 16
      Jung, J.; Nishima, W.; Daniels, M.; Bascom, G.; Kobayashi, C.; Adedoyin, A.; Wall, M.; Lappala, A.; Phillips, D.; Fischer, W.; Tung, C. S.; Schlick, T.; Sugita, Y.; Sanbonmatsu, K. Y. Scaling molecular dynamics beyond 100,000 processor cores for large-scale biophysical simulations. J. Comput. Chem. 2019, 40, 19191930,  DOI: 10.1002/jcc.25840
    17. 17
      Singharoy, A. Atoms to Phenotypes: Molecular Design Principles of Cellular Energy Metabolism. Cell 2019, 179, 10981111,  DOI: 10.1016/j.cell.2019.10.021
    18. 18
      Zhao, G.; Perilla, J. R.; Yufenyuy, E. L.; Meng, X.; Chen, B.; Ning, J.; Ahn, J.; Gronenborn, A. M.; Schulten, K.; Aiken, C.; Zhang, P. Mature HIV-1 capsid structure by cryo-electron microscopy and all-atom molecular dynamics. Nature 2013, 497, 643646,  DOI: 10.1038/nature12162
    19. 19
      Casalino, L.; AI-Driven Multiscale Simulations Illuminate Mechanisms of SARS-CoV-2 Spike Dynamics. bioRxiv 2020,  DOI: 10.1101/2020.11.19.390187 .
    20. 20
      Shaw, D. E.; In Anton 2: Raising the Bar for Performance and Programmability in a Special-Purpose Molecular Dynamics Supercomputer, SC ’14: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis, Nov 16–21, 2014; IEEE, 2014; pp 4153. DOI: 10.1109/SC.2014.9
    21. 21
      Ohmura, I.; Morimoto, G.; Ohno, Y.; Hasegawa, A.; Taiji, M. MDGRAPE-4: a special-purpose computer system for molecular dynamics simulations. Philos. Trans. R. Soc., A 2014, 372, 20130387,  DOI: 10.1098/rsta.2013.0387
    22. 22
      Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J. Y.; Wang, L.; Lupyan, D.; Dahlgren, M. K.; Knight, J. L.; Kaus, J. W.; Cerutti, D. S.; Krilov, G.; Jorgensen, W. L.; Abel, R.; Friesner, R. A. OPLS3: A Force Field Providing Broad Coverage of Drug-like Small Molecules and Proteins. J. Chem. Theory Comput. 2016, 12, 281296,  DOI: 10.1021/acs.jctc.5b00864
    23. 23
      Huang, J.; Rauscher, S.; Nawrocki, G.; Ran, T.; Feig, M.; de Groot, B. L.; Grubmüller, H.; MacKerell, A. D. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat. Methods 2017, 14, 7173,  DOI: 10.1038/nmeth.4067
    24. 24
      Tian, C.; Kasavajhala, K.; Belfon, K. A. A.; Raguette, L.; Huang, H.; Migues, A. N.; Bickel, J.; Wang, Y.; Pincay, J.; Wu, Q.; Simmerling, C. ff19SB: Amino-Acid-Specific Protein Backbone Parameters Trained against Quantum Mechanics Energy Surfaces in Solution. J. Chem. Theory Comput. 2020, 16, 528552,  DOI: 10.1021/acs.jctc.9b00591
    25. 25
      Best, R. B.; Zheng, W.; Mittal, J. Balanced Protein–Water Interactions Improve Properties of Disordered Proteins and Non-Specific Protein Association. J. Chem. Theory Comput. 2014, 10, 51135124,  DOI: 10.1021/ct500569b
    26. 26
      van der Spoel, D. Systematic design of biomolecular force fields. Curr. Opin. Struct. Biol. 2021, 67, 1824,  DOI: 10.1016/j.sbi.2020.08.006
    27. 27
      Robustelli, P.; Piana, S.; Shaw, D. E. Developing a molecular dynamics force field for both folded and disordered protein states. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, E4758E4766,  DOI: 10.1073/pnas.1800690115
    28. 28
      Piana, S.; Robustelli, P.; Tan, D.; Chen, S.; Shaw, D. E. Development of a Force Field for the Simulation of Single-Chain Proteins and Protein–Protein Complexes. J. Chem. Theory Comput. 2020, 16, 24942507,  DOI: 10.1021/acs.jctc.9b00251
    29. 29
      Li, P.; Merz, K. M. Metal Ion Modeling Using Classical Mechanics. Chem. Rev. 2017, 117, 15641686,  DOI: 10.1021/acs.chemrev.6b00440
    30. 30
      Zhang, A.; Yu, H.; Liu, C.; Song, C. The Ca2+ permeation mechanism of the ryanodine receptor revealed by a multi-site ion model. Nat. Commun. 2020, 11, 922,  DOI: 10.1038/s41467-020-14573-w
    31. 31
      Senn, H. M.; Thiel, W. QM/MM Methods for Biomolecular Systems. Angew. Chem., Int. Ed. 2009, 48, 11981229,  DOI: 10.1002/anie.200802019
    32. 32
      van der Kamp, M. W.; Mulholland, A. J. Combined Quantum Mechanics/Molecular Mechanics (QM/MM) Methods in Computational Enzymology. Biochemistry 2013, 52, 27082728,  DOI: 10.1021/bi400215w
    33. 33
      Cui, Q. Perspective: Quantum Mechanical Methods in Biochemistry and Biophysics. J. Chem. Phys. 2016, 145, 140901,  DOI: 10.1063/1.4964410
    34. 34
      Cui, Q.; Pal, T.; Xie, L. Biomolecular QM/MM Simulations: What Are Some of the ″burning Issues″?. J. Phys. Chem. B 2021, 125, 689702,  DOI: 10.1021/acs.jpcb.0c09898
    35. 35
      Warshel, A.; Karplus, M. Calculation of Ground and Excited State Potential Surfaces of Conjugated Molecules. I. Formulation and Parametrization. J. Am. Chem. Soc. 1972, 94, 56125625,  DOI: 10.1021/ja00771a014
    36. 36
      Warshel, A.; Levitt, M. Theoretical Studies of Enzymic Reactions: Dielectric, Electrostatic and Steric Stabilization of the Carbonium Ion in the Reaction of Lysozyme. J. Mol. Biol. 1976, 103, 227249,  DOI: 10.1016/0022-2836(76)90311-9
    37. 37
      Field, M. J.; Bash, P. A.; Karplus, M. A Combined Quantum Mechanical and Molecular Mechanical Potential for Molecular Dynamics Simulations. J. Comput. Chem. 1990, 11, 700733,  DOI: 10.1002/jcc.540110605
    38. 38
      Zhang, Y.; Liu, H.; Yang, W. Free energy calculation on enzyme reactions with an efficient iterative procedure to determine minimum energy paths on a combined ab initio QM/MM potential energy surface. J. Chem. Phys. 2000, 112, 3483,  DOI: 10.1063/1.480503
    39. 39
      Hu, H.; Lu, Z.; Yang, W. QM/MM Minimum Free-Energy Path: Methodology and Application to Triosephosphate Isomerase. J. Chem. Theory Comput. 2007, 3, 390406,  DOI: 10.1021/ct600240y
    40. 40
      Hu, H.; Yang, W. Free Energies of Chemical Reactions in Solution and in Enzymes with Ab Initio Quantum Mechanics/Molecular Mechanics Methods. Annu. Rev. Phys. Chem. 2008, 59, 573601,  DOI: 10.1146/annurev.physchem.59.032607.093618
    41. 41
      Kosugi, T.; Hayashi, S. QM/MM Reweighting Free Energy SCF for Geometry Optimization on Extensive Free Energy Surface of Enzymatic Reaction. J. Chem. Theory Comput. 2012, 8, 322334,  DOI: 10.1021/ct2005837
    42. 42
      Hayashi, S.; Uchida, Y.; Hasegawa, T.; Higashi, M.; Kosugi, T.; Kamiya, M. QM/MM Geometry Optimization on Extensive Free-Energy Surfaces for Examination of Enzymatic Reactions and Design of Novel Functional Properties of Proteins. Annu. Rev. Phys. Chem. 2017, 68, 135154,  DOI: 10.1146/annurev-physchem-052516-050827
    43. 43
      Rosta, E.; Woodcock, H. L.; Brooks, B. R.; Hummer, G. Artificial reaction coordinate “tunneling” in free-energy calculations: The catalytic reaction of RNase H. J. Comput. Chem. 2009, 30, 16341641,  DOI: 10.1002/jcc.21312
    44. 44
      Rosta, E.; Nowotny, M.; Yang, W.; Hummer, G. Catalytic Mechanism of RNA Backbone Cleavage by Ribonuclease H from Quantum Mechanics/Molecular Mechanics Simulations. J. Am. Chem. Soc. 2011, 133, 89348941,  DOI: 10.1021/ja200173a
    45. 45
      Rosta, E.; Yang, W.; Hummer, G. Calcium Inhibition of Ribonuclease H1 Two-Metal Ion Catalysis. J. Am. Chem. Soc. 2014, 136, 31373144,  DOI: 10.1021/ja411408x
    46. 46
      Ganguly, A.; Thaplyal, P.; Rosta, E.; Bevilacqua, P. C.; Hammes-Schiffer, S. Quantum mechanical/molecular mechanical free energy simulations of the self-cleavage reaction in the hepatitis delta virus ribozyme. J. Am. Chem. Soc. 2014, 136, 14831496,  DOI: 10.1021/ja4104217
    47. 47
      Zhang, S.; Ganguly, A.; Goyal, P.; Bingaman, J. L.; Bevilacqua, P. C.; Hammes-Schiffer, S. Role of the active site guanine in the glmS ribozyme self-cleavage mechanism: Quantum mechanical/molecular mechanical free energy simulations. J. Am. Chem. Soc. 2015, 137, 784798,  DOI: 10.1021/ja510387y
    48. 48
      Li, P.; Soudackov, A. V.; Hammes-Schiffer, S. Fundamental Insights into Proton-Coupled Electron Transfer in Soybean Lipoxygenase from Quantum Mechanical/Molecular Mechanical Free Energy Simulations. J. Am. Chem. Soc. 2018, 140, 30683076,  DOI: 10.1021/jacs.7b13642
    49. 49
      Stevens, D. R.; Hammes-Schiffer, S. Exploring the Role of the Third Active Site Metal Ion in DNA Polymerase η with QM/MM Free Energy Simulations. J. Am. Chem. Soc. 2018, 140, 89658969,  DOI: 10.1021/jacs.8b05177
    50. 50
      Li, P.; Rangadurai, A.; Al-Hashimi, H. M.; Hammes-Schiffer, S. Environmental Effects on Guanine-Thymine Mispair Tautomerization Explored with Quantum Mechanical/Molecular Mechanical Free Energy Simulations. J. Am. Chem. Soc. 2020, 142, 1118311191,  DOI: 10.1021/jacs.0c03774
    51. 51
      Walker, R. C.; Crowley, M. F.; Case, D. A. The Implementation of a Fast and Accurate QM/MM Potential Method in Amber. J. Comput. Chem. 2008, 29, 10191031,  DOI: 10.1002/jcc.20857
    52. 52
      Götz, A. W.; Clark, M. A.; Walker, R. C. An Extensible Interface for QM/MM Molecular Dynamics Simulations with AMBER. J. Comput. Chem. 2014, 35, 95108,  DOI: 10.1002/jcc.23444
    53. 53
      Melo, M. C. R.; Bernardi, R. C.; Rudack, T.; Scheurer, M.; Riplinger, C.; Phillips, J. C.; Maia, J. D. C.; Rocha, G. B.; Ribeiro, J. V.; Stone, J. E.; Neese, F.; Schulten, K.; Luthey-Schulten, Z. NAMD goes quantum: an integrative suite for hybrid simulations. Nat. Methods 2018, 15, 351354,  DOI: 10.1038/nmeth.4638
    54. 54
      Jung, J.; Kobayashi, C.; Kasahara, K.; Tan, C.; Kuroda, A.; Minami, K.; Ishiduki, S.; Nishiki, T.; Inoue, H.; Ishikawa, Y.; Feig, M.; Sugita, Y. New parallel computing algorithm of molecular dynamics for extremely huge scale biological systems. J. Comput. Chem. 2021, 42, 231241,  DOI: 10.1002/jcc.26450
    55. 55
      Jung, J.; Naurse, A.; Kobayashi, C.; Sugita, Y. Graphics Processing Unit Acceleration and Parallelization of GENESIS for Large-Scale Molecular Dynamics Simulations. J. Chem. Theory Comput. 2016, 12, 49474958,  DOI: 10.1021/acs.jctc.6b00241
    56. 56
      Jung, J.; Mori, T.; Sugita, Y. Midpoint cell method for hybrid (MPI+OpenMP) parallelization of molecular dynamics simulations. J. Comput. Chem. 2014, 35, 10641072,  DOI: 10.1002/jcc.23591
    57. 57
      Jung, J.; Kobayashi, C.; Imamura, T.; Sugita, Y. Parallel implementation of 3D FFT with volumetric decomposition schemes for efficient molecular dynamics simulations. Comput. Phys. Commun. 2016, 200, 5765,  DOI: 10.1016/j.cpc.2015.10.024
    58. 58
      Sugita, Y.; Okamoto, Y. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett. 1999, 314, 141151,  DOI: 10.1016/S0009-2614(99)01123-9
    59. 59
      Sugita, Y.; Okamoto, Y. Replica-exchange multicanonical algorithm and multicanonical replica-exchange method for simulating systems with rough energy landscape. Chem. Phys. Lett. 2000, 329, 261270,  DOI: 10.1016/S0009-2614(00)00999-4
    60. 60
      Sugita, Y.; Kitao, A.; Okamoto, Y. Multidimensional replica-exchange method for free-energy calculations. J. Chem. Phys. 2000, 113, 60426051,  DOI: 10.1063/1.1308516
    61. 61
      Kamiya, M.; Sugita, Y. Flexible selection of the solute region in replica exchange with solute tempering: Application to protein-folding simulations. J. Chem. Phys. 2018, 149, 72304,  DOI: 10.1063/1.5016222
    62. 62
      Miao, Y.; Feher, V. A.; McCammon, J. A. Gaussian Accelerated Molecular Dynamics: Unconstrained Enhanced Sampling and Free Energy Calculation. J. Chem. Theory Comput. 2015, 11, 35843595,  DOI: 10.1021/acs.jctc.5b00436
    63. 63
      Oshima, H.; Re, S.; Sugita, Y. Replica-Exchange Umbrella Sampling Combined with Gaussian Accelerated Molecular Dynamics for Free-Energy Calculation of Biomolecules. J. Chem. Theory Comput. 2019, 15, 51995208,  DOI: 10.1021/acs.jctc.9b00761
    64. 64
      Tembre, B. L.; Mc Cammon, J. A. Ligand-receptor interactions. Comput. Chem. 1984, 8, 281283,  DOI: 10.1016/0097-8485(84)85020-2
    65. 65
      Oshima, H.; Re, S.; Sugita, Y. Prediction of Protein-Ligand Binding Pose and Affinity Using the gREST+FEP Method. J. Chem. Inf. Model. 2020, 60, 53825394,  DOI: 10.1021/acs.jcim.0c00338
    66. 66
      Maragliano, L.; Fischer, A.; Vanden-Eijnden, E.; Ciccotti, G. String method in collective variables: Minimum free energy paths and isocommittor surfaces. J. Chem. Phys. 2006, 125, 24106,  DOI: 10.1063/1.2212942
    67. 67
      Matsunaga, Y.; Komuro, Y.; Kobayashi, C.; Jung, J.; Mori, T.; Sugita, Y. Dimensionality of Collective Variables for Describing Conformational Changes of a Multi-Domain Protein. J. Phys. Chem. Lett. 2016, 7, 14461451,  DOI: 10.1021/acs.jpclett.6b00317
    68. 68
      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
    69. 69
      Yagi, K.; Yamada, K.; Kobayashi, C.; Sugita, Y. Anharmonic Vibrational Analysis of Biomolecules and Solvated Molecules Using Hybrid QM/MM Computations. J. Chem. Theory Comput. 2019, 15, 19241938,  DOI: 10.1021/acs.jctc.8b01193
    70. 70
      Frisch, M. J.; Gaussian 16, rev. C.01; Gaussian, Inc.: Wallingford, CT, 2016.
    71. 71
      Shao, Y. Advances in molecular quantum chemistry contained in the Q-Chem 4 program package. Mol. Phys. 2015, 113, 184215,  DOI: 10.1080/00268976.2014.952696
    72. 72
      Seritan, S.; Bannwarth, C.; Fales, B. S.; Hohenstein, E. G.; Isborn, C. M.; Kokkila-Schumacher, S. I. L.; Li, X.; Liu, F.; Luehr, N.; Snyder, J. W., Jr; Song, C.; Titov, A. V.; Ufimtsev, I. S.; Wang, L.-P.; Martínez, T. J. TeraChem: A graphical processing unit-accelerated electronic structure package for large-scale ab initio molecular dynamics. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2021, 11, e1494,  DOI: 10.1002/wcms.1494
    73. 73
      Hourahine, B. DFTB+, a software package for efficient approximate density functional theory based atomistic simulations. J. Chem. Phys. 2020, 152, 124101,  DOI: 10.1063/1.5143190
    74. 74
      Yagi, K. SINDO 4.0 beta. 2020, https://tms.riken.jp/en/research/software/sindo/.
    75. 75
      QSimulate-QM. Quantum Simulation Technologies, Inc. 2020, https://qsimulate.com/.
    76. 76
      E, W.; Ren, W.; Vanden-Eijnden, E. String method for the study of rare events. Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 66, 52301,  DOI: 10.1103/PhysRevB.66.052301
    77. 77
      E, W.; Ren, W.; Vanden-Eijnden, E. Simplified and improved string method for computing the minimum energy paths in barrier-crossing events. J. Chem. Phys. 2007, 126, 164103,  DOI: 10.1063/1.2720838
    78. 78
      Cui, Q.; Karplus, M. Quantum mechanics/molecular mechanics studies of triosephosphate isomerase-catalyzed reactions: Effect of geometry and tunneling on proton-transfer rate constants. J. Am. Chem. Soc. 2002, 124, 30933124,  DOI: 10.1021/ja0118439
    79. 79
      Zhang, X.; Harrison, D. H. T.; Cui, Q. Functional specificities of methylglyoxal synthase and triosephosphate isomerase: A combined QM/MM analysis. J. Am. Chem. Soc. 2002, 124, 1487114878,  DOI: 10.1021/ja027063x
    80. 80
      Cui, Q.; Karplus, M. Quantum mechanical/molecular mechanical studies of the triosephosphate isomerase-catalyzed reaction: Verification of methodology and analysis of reaction mechanisms. J. Phys. Chem. B 2002, 106, 17681798,  DOI: 10.1021/jp012659c
    81. 81
      Lennartz, C.; Schäfer, A.; Terstegen, F.; Thiel, W. Enzymatic reactions of triosephosphate isomerase: A theoretical calibration study. J. Phys. Chem. B 2002, 106, 17581767,  DOI: 10.1021/jp012658k
    82. 82
      Mendieta-Moreno, J. I.; Walker, R. C.; Lewis, J. P.; Gómez-Puertas, P.; Mendieta, J.; Ortega, J. FIREBALL/AMBER: An Efficient Local-Orbital DFT QM/MM Method for Biomolecular Systems. J. Chem. Theory Comput. 2014, 10, 21852193,  DOI: 10.1021/ct500033w
    83. 83
      Knizia, G. Intrinsic atomic orbitals: An unbiased bridge between quantum theory and chemical concepts. J. Chem. Theory Comput. 2013, 9, 48344843,  DOI: 10.1021/ct400687b
    84. 84
      Kästner, J.; Thiel, S.; Senn, H. M.; Sherwood, P.; Thiel, W. Exploiting QM/MM Capabilities in Geometry Optimization: A Microiterative Approach Using Electrostatic Embedding. J. Chem. Theory Comput. 2007, 3, 10641072,  DOI: 10.1021/ct600346p
    85. 85
      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
    86. 86
      Branduardi, D.; Gervasio, F. L.; Parrinello, M. From A to B in free energy space. J. Chem. Phys. 2007, 126, 054103,  DOI: 10.1063/1.2432340
    87. 87
      Bonomi, M.; Branduardi, D.; Bussi, G.; Camilloni, C.; Provasi, D.; Raiteri, P.; Donadio, D.; Marinelli, F.; Pietrucci, F.; Broglia, R. A.; Parrinello, M. PLUMED: A portable plugin for free-energy calculations with molecular dynamics. Comput. Phys. Commun. 2009, 180, 19611972,  DOI: 10.1016/j.cpc.2009.05.011
    88. 88
      Davenport, R. C.; Bash, P. A.; Seaton, B. A.; Karplus, M.; Petsko, G. A.; Ringe, D. Structure of the Triosephosphate Isomerase-Phosphoglycolohydroxamate Complex: An Analogue of the Intermediate on the Reaction Pathway. Biochemistry 1991, 30, 58215826,  DOI: 10.1021/bi00238a002
    89. 89
      Olsson, M. H. M.; Søndergaard, C. R.; Rostkowski, M.; Jensen, J. H. PROPKA3: Consistent Treatment of Internal and Surface Residues in Empirical pKa predictions. J. Chem. Theory Comput. 2011, 7, 525537,  DOI: 10.1021/ct100578z
    90. 90
      Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: A web-based graphical user interface for CHARMM. J. Comput. Chem. 2008, 29, 18591865,  DOI: 10.1002/jcc.20945
    91. 91
      Lee, J.; Cheng, X.; Swails, J. M.; Yeom, M. S.; Eastman, P. K.; Lemkul, J. A.; Wei, S.; Buckner, J.; Jeong, J. C.; Qi, Y.; Jo, S.; Pande, V. S.; Case, D. A.; Brooks, C. L.; MacKerell, A. D.; Klauda, J. B.; Im, W. CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field. J. Chem. Theory Comput. 2016, 12, 405413,  DOI: 10.1021/acs.jctc.5b00935
    92. 92
      Gaus, M.; Cui, Q.; Elstner, M. DFTB3: Extension of the Self-Consistent-Charge Density-Functional Tight-Binding Method (SCC-DFTB). J. Chem. Theory Comput. 2011, 7, 931948,  DOI: 10.1021/ct100684s
    93. 93
      Best, R. B.; Zhu, X.; Shim, J.; Lopes, P. E. M.; Mittal, J.; Feig, M.; Mackerell, A. D. Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone φ, ψ and Side-Chain χ1 and χ2 Dihedral Angles. J. Chem. Theory Comput. 2012, 8, 32573273,  DOI: 10.1021/ct300400x
    94. 94
      Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926935,  DOI: 10.1063/1.445869
    95. 95
      Mayne, C. G.; Saam, J.; Schulten, K.; Tajkhorshid, E.; Gumbart, J. C. Rapid parameterization of small molecules using the force field toolkit. J. Comput. Chem. 2013, 34, 27572770,  DOI: 10.1002/jcc.23422
    96. 96
      Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101,  DOI: 10.1063/1.2408420
    97. 97
      Andersen, H. C. Rattle: A “velocity” version of the shake algorithm for molecular dynamics calculations. J. Comput. Phys. 1983, 52, 2434,  DOI: 10.1016/0021-9991(83)90014-1
    98. 98
      Miyamoto, S.; Kollman, P. A. Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models. J. Comput. Chem. 1992, 13, 952962,  DOI: 10.1002/jcc.540130805
    99. 99
      Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785789,  DOI: 10.1103/PhysRevB.37.785
    100. 100
      Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 56485652,  DOI: 10.1063/1.464913
    101. 101
      Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104,  DOI: 10.1063/1.3382344
    102. 102
      Dunning, T. H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 10071023,  DOI: 10.1063/1.456153
    103. 103
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    104. 104
      Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 32973305,  DOI: 10.1039/b508541a
    105. 105
      Byrd, R. H.; Lu, P.; Nocedal, J.; Zhu, C. A Limited Memory Algorithm for Bound Constrained Optimization. SIAM J. Sci. Stat. Comp. 1995, 16, 11901208,  DOI: 10.1137/0916069
    106. 106
      Zhu, C.; Byrd, R.; Nocedal, J.; Morales, J. L. L-BFGS-B (ver. 3.0), http://users.iems.northwestern.edu/~nocedal/lbfgsb.html.
    107. 107
      Zhu, C.; Byrd, R. H.; Lu, P.; Nocedal, J. L-BFGS-B: Algorithm 778: L-BFGS-B, FORTRAN routines for large scale bound constrained optimization. ACM Transactions on Mathematical Software 1997, 23, 550560,  DOI: 10.1145/279232.279236
    108. 108
      Lence, E.; van der Kamp, M. W.; González-Bello, C.; Mulholland, A. J. QM/MM simulations identify the determinants of catalytic activity differences between type II dehydroquinase enzymes. Org. Biomol. Chem. 2018, 16, 44434455,  DOI: 10.1039/C8OB00066B
    109. 109
      Albery, W. J.; Knowles, J. R. Free-Energy Profile for the Reaction Catalyzed by Triosephosphate Isomerase. Biochemistry 1976, 15, 56275631,  DOI: 10.1021/bi00670a031
    110. 110
      Hudson, P. S.; Woodcock, H. L.; Boresch, S. Use of Nonequilibrium Work Methods to Compute Free Energy Differences Between Molecular Mechanical and Quantum Mechanical Representations of Molecular Systems. J. Phys. Chem. Lett. 2015, 6, 48504856,  DOI: 10.1021/acs.jpclett.5b02164
    111. 111
      Kearns, F. L.; Hudson, P. S.; Woodcock, H. L.; Boresch, S. Computing Converged Free Energy Differences between Levels of Theory via Nonequilibrium Work Methods: Challenges and Opportunities. J. Comput. Chem. 2017, 38, 13761388,  DOI: 10.1002/jcc.24706
    112. 112
      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
    113. 113
      Holden, Z. C.; Richard, R. M.; Herbert, J. M. Periodic boundary conditions for QM/MM calculations: Ewald summation for extended Gaussian basis sets. J. Chem. Phys. 2013, 139, 244108,  DOI: 10.1063/1.4850655
    114. 114
      Giese, T. J.; York, D. M. Ambient-Potential Composite Ewald Method for ab Initio Quantum Mechanical/Molecular Mechanical Molecular Dynamics Simulation. J. Chem. Theory Comput. 2016, 12, 26112632,  DOI: 10.1021/acs.jctc.6b00198
    115. 115
      Vasilevskaya, T.; Thiel, W. Periodic Boundary Conditions in QM/MM Calculations: Implementation and Tests. J. Chem. Theory Comput. 2016, 12, 35613570,  DOI: 10.1021/acs.jctc.6b00269
    116. 116
      Kawashima, Y.; Ishimura, K.; Shiga, M. Ab initio quantum mechanics/molecular mechanics method with periodic boundaries employing Ewald summation technique to electron-charge interaction: Treatment of the surface-dipole term. J. Chem. Phys. 2019, 150, 124103,  DOI: 10.1063/1.5048451
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcb.1c01862.

    • Protocols of equilibration, interatomic distances, calculated total energies and proton affinities, Cartesian coordinates, details of umbrella sampling, 2D PMF obtained by DFTB3 (PDF)

    • Animation for the visualization of minimum-energy pathways from DHAP to GAP using QM/MM calculations at the level of B3LYP-D3/aug-cc-pVDZ (MP4)

    • Animation for the visualization of minimum-energy pathways from DHAP to GAP using QM/MM calculations at the levels of DFTB3 (MP4)


    Terms & Conditions

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

Pair your accounts.

Export articles to Mendeley

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

Pair your accounts.

Export articles to Mendeley

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

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

STEP 1:
Click to create an ACS ID

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

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

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

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect