Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Dec 23.
Published in final edited form as: J Phys Chem A. 2021 Dec 12;125(50):10677–10685. doi: 10.1021/acs.jpca.1c07727

Affordable Ab Initio Path Integral for Thermodynamic Properties via Molecular Dynamics Simulations Using Semiempirical Reference Potential

Yuanfei Xue , Jia-Ning Wang , Wenxin Hu , Jun Zheng , Yongle Li , Xiaoliang Pan §, Yan Mo †,∥,, Yihan Shao §, Lu Wang #, Ye Mei †,∥,
PMCID: PMC9108008  NIHMSID: NIHMS1803100  PMID: 34894680

Abstract

Path integral molecular dynamics (PIMD) is becoming a routinely applied method for incorporating the nuclear quantum effect in computer simulations. However, direct PIMD simulations at an ab initio level of theory are formidably expensive. Using the protonated 1,8-bis(dimethylamino)naphthalene molecule as an example, we show in this work that the computational expense for the intramolecular proton transfer between the two nitrogen atoms can be remarkably reduced by implementing the idea of reference-potential methods. The simulation time can be easily extended to a scale of nanoseconds while maintaining the accuracy on an ab initio level of theory for thermodynamic properties. In addition, post-processing can be carried out in parallel on massive computer nodes. A 545-fold reduction in the total CPU time can be achieved in this way as compared to a direct PIMD simulation at the same ab initio level of theory.

Graphical Abstract

graphic file with name nihms-1803100-f0001.jpg

Introduction

Hybrid QM/MM is now the method of choice for studying enzymatic reactions and chemical reactions in the condensed phase.113 However, the application of QM/MM methods is often plagued with extremely poor computational scaling of the ab initio QM methods. In addition, the long time scales of molecular dynamics propagation is usually required before any essential dynamic processes can be observed, especially for the degrees-of-freedom (DoF) orthogonal to the boosted one.14 Furthermore, exploring the nuclear quantum effect (NQE), such as tunneling, nuclear delocalization phenomena, and zero-point energy of molecules, has gained in popularity over the years.1522 One of the most appealing approaches to incorporating NQE is path integral molecular dynamics (PIMD).23 PIMD is based on the isomorphism to approximately transform the representation of quantum effect in single-particle systems into a classical system with a finite number of pseudo-particles (or beads) along a cyclic path.24 Being one of the most commonly used simulation methods, PIMD can be used to obtain thermostatistical equilibrium properties and dynamic properties, with free energy profiles of reactions and reaction rate coefficients as two representative examples.2528

However, applying PIMD simulations in the studies of chemical reactions of complex systems is extremely demanding. On the one hand, PIMD requires P times more computational time than the classical molecular dynamics (MD), where P is the number of beads onto which each quantum particle is mapped. The value of P determines the convergence of PIMD simulations. P > ħωmax/kBT must be satisfied, where ωmax is the maximal vibrational frequency of the system. P surges as ωmax increases or T decreases. It is thus expensive to perform high-level PIMD with a convergence-ensured number of beads for a sufficiently long timescale. To reduce the computational cost, some progress has been achieved. Tuckerman et al. combined a noncanonical transformation on the quadratic part of the action with multiple time scale integration techniques and proposed a very efficient algorithm.29 Markland and Manolopoulos proposed an approach to reduce the computational effort by separating the short-range interaction from the long-range one.30 Cheng et al. proposed the multiple-timestep molecular dynamics (MTS-MD) algorithm to accelerate the propagation by dividing the force into a slowly oscillating part and a quickly varying component.31 However, the former factor is still the efficiency bottleneck. A large integration time step can also be made possible by applying proper transformations, so the analytic integration over the high-frequency modes is allowed.32 After the transformation, the number of the beads can be reduced via ring-polymer contraction, in which some highest normal modes can be ignored.3335 Marsalek and Markland proposed a ring polymer contraction approach that can dramatically reduce the computational cost of PIMD at an ab initio level.36 On the other hand, the accessible simulation time is further limited when using high-level electronic structure methods such as post-Hartree-Fock theories and density functional theory (DFT). Approximated levels of theory, such as semi-empirical (SE) methods, can improve computational efficiency and allow sufficient exploration in the phase space for systems in the condensed phase. However, these methods are crude in treating electronic structures, leading to inexact results for specific physical properties.

In addition, ergodicity in the configurational space is critical for the convergence of the simulated macrostate properties. Enhanced sampling methods, such as umbrella sampling (US),37 adaptively biased molecular dynamics (ABMD),38 and metadynamics,39,40 are now frequently employed to boost one or two DoF and accelerate the reaction. For the orthogonal DoF, the exploration can still be hindered by hidden barriers, and a ns-timescale simulation can be inevitable. Besides, the length of the simulations determines the effective number of samples, because independent and identically distributed (i.i.d.) samples are required for the calculations of ensemble averages. For simulations in aqueous solutions, the energy correlation time is on a scale of around one ps. If the samples are saved too frequently, sampling inefficiency must be considered.41 Affordable studies are limited to relatively simple systems,42 low levels of theory,43 or with the aid of artificial intelligence.44

Therefore, it is desirable to develop a practical method for the study of thermodynamics properties based on PIMD that can substantially reduce the computational expense while maintaining accuracy. The multistate thermodynamic perturbation (MsTP) is the method of choice, of which the efficiency and reliability have been evaluated by Li et al. in an early study of some model reactions.45 MsTP is a variant of the reference-potential method, in which the ensemble averages of any time-independent physical properties at a high level of theory are obtained indirectly from simulations at a lower level of theory. With further improvement,4648 this method is becoming more robust for the studies of chemical processes in the condensed phase. The idea of the reference-potential method has also been implemented by other groups.4965

In the current work, we combine PIMD with the reference potential method for the first time and demonstrate its application in the study of the intramolecular proton transfer in the protonated 1,8-bis(dimethylamino)naphthalene (DMANH) molecule. The free energy profile at the PIMD/BLYP-D3/6–31G(d)/MM level of theory along the proton transfer pathway was calculated leveraging a more efficient sampling at the PM6/MM level. A 545-fold enhancement in the computational efficiency was achieved compared to a direct PIMD simulation at the same ab initio level of theory.

Theory

The partition function of a system with N distinguishable particles is defined as,66

ZP(β)=NdNPpdNPxeβH^p(x,p), (1)

where β = 1/kBT, N is the total normalization function, x represents the position of particles, and p the momenta conjugate to x. The isomorphic primitive Hamiltonian operator H^p(x,p) of this N-atom P-bead cyclic model is

H^p(x,p)=i=1P{n=1N[pni22m˜ni+12mnωp2(xni+1xni)2]+1PU(x1i,,xNi)}xnP+1=xn1, (2)

under the constraint xnP+1=xn1. This quantum Hamiltonian operator is made up of P group of beads (also known as replicas or time-slices), and each forms a classical representation of the original system. Beads of each atom are connected by harmonic potentials. In this way, a P-membered classical “necklace” model is constructed, and the simulation of quantum systems is transformed into an expanded classical calculation. mn is the mass of the particle, U is the potential energy, m˜ni and pni are the fictitious mass of the ith bead of the nth particle and its corresponding momentum. The ωp=P/βħ is the harmonic constraint connecting the two most adjacent beads. For the ideal quantum problem, the full quantum Hamiltonian H^ is established when P → ∞. The introduction of this isomorphic relationship also means the quantum mechanical sampling can be mapped onto classical simulations by either Monte Carlo (MC)67,68 or MD69 methods. The potential of mean force with a collection of PIMD samples can thus be written as

F(ξPI)=1βln1Pi=1Pδ(ξPIξ(xi)). (3)

To accelerate computation, we incorporate of the PIMD and the reference-potential method (RPM) for the first time to calculate the statistical quantum mechanical properties at a level-of-interest Hamiltonian while simulating at a lower level Hamiltonian state. Additionally, we adopt the US method to enhance the sampling. From K simulations, we can collect Nk configurations from the kth simulation. Each simulation is characterized by its specific potential function Uk, for k ∈ 1 …,K. In the US method,

Uk(r)=U0(r)+Wk(η(r)), (4)

where U0(r) is the unbiased potential and Wk(r) is the biasing potential acting on some chosen collective variables (CV) η(r). The unnormalized weight of each configuration under another Hamiltonian Ut can be calculated via extrapolation as

ωt(rn)=eβUt(rn)k=1KNkeβ[Uk(rn)fk]=eβΔUt(rn)k=1KNkeβ[Wk(η(rn))fk], (5)

in which fk is the free energy of state k and is calculated by iteratively solving the Multistate Bennett Acceptance Ratio (MBAR)70 equations

fi=β1lnn=1NeβUi(rn)k=1KNkeβ[Uk(rn)fk]    i=1,,K. (6)

Here, N=k=1KNk is total number of snapshots collected in all the simulations, and ΔUt(r) = Ut(r) − U0(r). ωt (rn) can be understood as the weight factor for each configuration in a thermodynamic perturbation from the mixed reference potentials (Uk) to the target Hamiltonian (Ut). Therefore, this method is referred to as the multistate thermodynamic perturbation (MsTP). Thermodynamic properties that depend only on atomic coordinates under Ut can thus be obtained as

A^t=n=1Nωt(rn)A^(rn)n=1Nωt(rn). (7)

If A is an indication function δ of some chosen CV ξ(r)

δ(ξmξ(r))={1,if Δξ/2<ξmξ(r)<Δξ/20,otherwise , (8)

the potential of mean force (PMF) under Ut(r) can be written as

Ft(ξm)=β1lnn=1Nωt(rn)δ(ξmξ(rn)) (9)

defined up to an additive constant. Similarly, the PMF under U0(r) can be written as

F0(ξm)=β1lnn=1Nω0(rn)δ(ξmξ(rn)), (10)

where

ω0(rn)=eβU0(rn)k=1KNkeβ[Uk(rn)fk]=1k=1KNkeβ[Wk(η(rn))fk] (11)

is the unnormalized weight for configuration rn under U0(r). To ameliorate the numerical instability originated from incomplete sampling, the weights ωt(rn) under the target Hamiltonian are scaled by Gaussian-smoothing on the density-of-states.47

In this work, we take U0 as the PM6/MM71 Hamiltonian, and Ut as the BLYP-D3/6–31G(d)/MM7274 level of theory. For simplicity, they will be labeled as the SQM and DFT levels of theory. Please also note that η is unnecessarily ξ, but we set them equal in this work.

Simulations

One DMANH molecule, shown in Fig. 1,75,76 was solvated in a TIP3P water77 sphere with a radius of 25 Å using the LEaP module in the AmberTools19 package.78 One chlorine ion was added for neutralization. The general AMBER force field (GAFF)79 and the AM1-bcc charges were assigned to the DMANH molecule. The QM region contained only the DMANH molecule, and the MM region comprised all the solvent molecules and the counter ion. In the MM region, the SHAKE algorithm80 was applied to constrain all the bonds involving hydrogen atoms. The system was energy-minimized for 2000 steps and heated up to 298 K in 100 ps. Then a 100-ps classical MD simulation was conducted to further relax the system. Finally, the quantum dynamics simulations were performed with the primitive approximation PIMD (PRIMPIMD)67 module in the AmberTools19 package.78 The temperature was regulated using the “middle scheme” Langevin thermostat.81,82 The non-periodic boundary condition was used, and a soft cap restraining potential with a force constant of 10 kcal · mol−1 · Å−2 was added to prevent the solvent from evaporation. The nonbonded interaction was fully counted without any truncation. All the simulations were conducted using the SANDER module in the AmberTools 19 package.78

Figure 1:

Figure 1:

Molecular structure of DMANH

The CV was defined as the difference in the N–H bond lengths, i.e., dNaHdNdH, where dNaH is the distance between the proton H and the acceptor nitrogen atom Na, and dNdH the distance between the proton H and the donor nitrogen atom Nd. The CV, ranged from −0.15 Å to 1.10 Å, was discretized into 26 windows in total. Harmonic potentials Wk(x)=12kξ(ξξk)2 were used with kξ = 100kcal · mol−1 · Å−2 for each group of the beads of the system. We first performed the primitive approximation PIMD (PRIMPIMD)67 with 16 beads for each QM atom for 1-ns per window, and then an additional 1-ns simulation was performed for each of the first 16 windows (with RC between −0.15 Å and 0.60 Å) to improve the convergence in this region. The temperature was regulated using the “middle scheme” Langevin thermostat.81,82 The step size was set to 0.5 fs. Configurations were saved every 1000 steps (500 fs) for single-point energy calculations at the DFT level, as well as the SQM level. Due to the symmetry of the molecule, we only calculated the PMF along the CV with positive values. The potential of mean force at the SQM level was generated via the MBAR analysis, and that at the DFT level was calculated with the MsTP and was smoothed using the Gaussian process regression.83 For comparison, classical QM/MM MD (CMD) simulations at the same level of theory enhanced by Hamiltonian replica exchange molecular dynamics (H-REMD) of umbrella sampling were also performed.

Results and Discussion

Reweighting entropy (RE) is a measure of the reliability of the RPM. We have previously shown that when RE is greater than 0.3 in the MsTP calculations, the extrapolation process is reliable.45 When it is below 0.3, large biases may reside in the free energy profile. As shown in Fig. 2, the free energy profile under the DFT level is reliable when the CV is smaller than 0.8 Å. When the CV is greater than 0.8 Å, RE plunges. This is the consequence of the significant difference in the preferred structures under these two Hamiltonians, i.e., SQM and DFT. Shown in Fig. 3 are the means and fluctuations of the local structures for the proton transfer reaction. With the CV increases from 0.00 Å to 0.80 Å, dNdH decreases from around 1.34 Å to about 1.05 Å, and dNaH increases from 1.34 Å to 1.85 Å. These distance variations are very similar to those under the SQM Hamiltonian, as shown in Fig. S1. However, dNdH shows a sudden drop when the CV is greater than 0.8 Å and deviates from the mean bond length under the SQM Hamiltonian, indicating an insufficient overlap in the sub-phase space under these two Hamiltonians. The difference is more evident for the Nd–H–Na angle. Therefore, thermodynamic properties with the CV greater than 0.8 Å have not reached convergence, and more samples are required in this region. For the PMF analysis at the DFT level in the following, we can only trust the results for the region with the CV below 0.8 Å.

Figure 2:

Figure 2:

Reweighting entropy

Figure 3:

Figure 3:

Structural alternations during the proton transfer process at the DFT/MM level from MsTP. (A) Distance between the proton (H) and the Nd atom. (B) Distance between the proton (H) and the Na. (C) Distance between the Nd and the Na atoms. (D) Angle of Nd–H–Na. The shaded areas are the standard error.

The free energy profiles under the SQM and DFT levels are shown in Fig. 4. Under the SQM level, the uncertainties are much smaller than their counterparts under the extrapolated DFT level. The large magnitude in the uncertainties under the DFT level is the consequence of the sampling inefficiency when the MBAR weights are assigned to the samples. As shown in Eq. 5, the weight decays exponentially as the energy difference between the two Hamiltonians increases. Precisely, ΔUt is the difference in the deviations of the potential energies from their respective means, since the difference in the means can be canceled in normalization. Under the SQM level, the minimum in the free energy profile locates at CV 0.6Å, which is about 0.1 Å larger than the minimum under the DFT level. This indicates that in the optimal structure under the SQM level, the proton prefers to be closer to one of the nitrogen atoms, while the proton is more diffusive under the DFT level. This can be explained by the difference in the basis sets used in the SQM and DFT calculations. PM6 used the minimum basis set, while a larger basis set 6–31G(d) was used in the DFT calculations. Both of the profiles show a reaction free energy barrier of about 1.2 kcal/mol. This picture is different from the results in a recent study by Zhou and Wang, in which the reaction is barrierless.84 The disagreement comes from the difference in the Hamiltonians. All the atoms were quantum particles in their studies, while in the present study, only the DMANH molecule was treated quantum mechanically. Quantum-mechanically described water molecules can stabilize the delocalized proton and thereby lower the reaction barrier. In the current work, we are not aiming at reproducing the results of a full-QM PIMD model. However, the central idea developed in the current work can be applied to a full-QM PIMD model, if a proper reference Hamiltonian can be found. This is beyond the interest of the current work. For the CV to the right side of the minimum, the potential of mean force increases very fast w.r.t the CV under the DFT level than under the SQM level, which indicates a higher probability of a shared proton between the nitrogen atoms under the DFT level.

Figure 4:

Figure 4:

Free energy profiles of the proton transfer under different levels of theory

The estimated CPU times of obtaining PMF at the BLYP-D3 level with direct QM/MM and RPM are listed in Table 1. The CPU time of the RPM includes two parts, i.e., the CPU time for collecting the trajectories at the PM6/MM level and the CPU time for the single-point energy calculations at both the PM6/MM and BYLP-D3/MM levels. The direct QM/MM method at the BLYP-D3/MM level only requires the time for the simulations. Since the PM6/MM level is several orders of magnitude more efficient than the BLYP-D3/MM level and the configurations were saved once every 1000 steps of MD propagation for the single-point energy calculations, a 545-fold enhancement in the efficiency was observed for the RPM method as compared to a direct PIMD simulation at the same level of theory.

Table 1:

Estimated CPU Time in a unit of Days. Assuming one node with 28 cores of Intel Xeon Gold 6132 CPU 2.60 GHz was used.

RP method Direct DFT a
Sampling Energy Evaluation Total
1.04 K 1.25 K 2.29 K 1250 K
a

Estimated CPU time assuming the number of windows and simulation length for each window

In principle, this method is valid no matter how large the QM region is. However, the convergence rate depends on the difference between the two Hamiltonians.85,86 The larger the QM region is, the more violent the magnitude of the fluctuation in energy difference is. Therefore, extending the QM size slows down the convergence, and the number of samples required increases fast practically. If a larger QM region is necessary, for instance, when the solvent plays a more critical role than pure background charges polarizing the reaction center, more accurate reference potential is indispensable. Some recent studies have looked into this issue, and some practical solutions have been proposed, including optimizing the semi-empirical methods and fitting the (delta) energy using machine learning techniques.46,57,63,8793 It is worth emphasizing that with the uncertainties in the free energy profiles in the present study, it is difficult to quantitatively predict the magnitude of nuclear quantum effect (See Fig. S4), and more extended simulations are required.

Conclusion

The calculations of thermodynamic properties from computer simulations require independent and identically distributed samples. The long correlation time of chemical processes in condensed phases plagues the simulations with slow convergence in thermodynamic properties. It poses a challenge for the computational studies at ab initio levels of theory. This difficulty is more severe for PIMD, in which the representation of quantum effect in single-particle systems is transformed into a classical problem of discrete cyclic model systems with a number of beads for each atom. The reference-potential methods can remedy the large expense by leveraging an inexpensive sampling method and a post-processing process on the harvested samples. This work studied the proton transfer within a protonated 1,8-bis(dimethylamino)naphthalene molecule, and the free energy profile at the BLYP-D3/6–31G(d)/MM level of theory was calculated using a more efficient sampling at the PM6/MM level. With a finite number of samples, the reliability of this reference-potential method heavily depends on the overlap in the phase space between the target (BLYP-D3/6–31G(d)/MM) Hamiltonian and the reference (PM6/MM) Hamiltonian. The results show that when the CV, defined as the difference in the distances from the proton to the acceptor and to the donor, is below 0.8 Å, the potential of mean force and the structural properties can be reliably produced with a 545-fold enhancement in efficiency.

Supplementary Material

supporting information

Acknowledgement

Y. Mei owes Dr. Jian Liu many thanks for the helpful discussions. Y. Mei is supported by the Ministry of Science and Technology of China (Grant No. 2016YFA0501700) and the National Natural Science Foundation of China (Grant No. 22073030). W.H. is supported by the Fundamental Research Funds for the Central Universities. Y. Mo is supported by the National Natural Science Foundation of China (Grant No. 21973030). Y.L. is supported by the National Natural Science Foundation of China (Grant No. 11674212). Y.S. is supported by the National Institutes of Health (Grant No. R01GM135392). CPU time was supported by the Supercomputer Center of East China Normal University (ECNU Public Platform for Innovation No. 001).

Footnotes

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpca.XXXXX
  • Structural alternations during the reaction at the PM6/MM level
  • Convergence check of the free energy profiles with Gaussian process regression at the BLYP-D3 level with respect to numbers of samples
  • Distribution of ΔU for all the samples
  • Comparison of the free energy profiles between the classical QM/MM MD (CMD) and PIMD

This material is available free of charge via the Internet at http://pubs.acs.org/.

References

  • (1).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, 227–249. [DOI] [PubMed] [Google Scholar]
  • (2).Field MJ; Bash PA; Karplus M A Combined Quantum Mechanical and Molecular Mechanical Potential for Molecular Dynamics Simulations. J. Comput. Chem 1990, 11, 700–733. [Google Scholar]
  • (3).Monard G; Merz KM Combined Quantum Mechanical/Molecular Mechanical Methodologies Applied to Biomolecular Systems. Acc. Chem. Res 1999, 32, 904–911. [Google Scholar]
  • (4).Tresadern G; Nunez S; Faulder PF; Wang H; Hillier IH; Burton NA Direct Dynamics Calculations of Reaction Rate and Kinetic Isotope Effects in Enzyme Catalysed Reactions. Faraday Discuss 2003, 122, 223–242. [DOI] [PubMed] [Google Scholar]
  • (5).Friesner RA; Guallar V Ab Initio Quantum Chemical and Mixed Quantum Mechanics/Molecular Mechanics (QM/MM) Methods for Studying Enzymatic Catalysis. Annu. Rev. Phys. Chem 2005, 56, 389–427. [DOI] [PubMed] [Google Scholar]
  • (6).Gao J; Ma S; Major DT; Nam K; Pu J; Truhlar DG Mechanisms and Free Energies of Enzymatic Reactions. Chem. Rev 2006, 106, 3188–3209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).Lin H; Truhlar DG QM/MM: What Have We Learned, Where Are We, and Where Do We Go from Here? Theor. Chem. Acc 2006, 117, 185. [Google Scholar]
  • (8).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, 573–601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).Senn HM; Thiel W QM/MM Methods for Biomolecular Systems. Angew. Chem., Int. Ed 2009, 48, 1198–1229. [DOI] [PubMed] [Google Scholar]
  • (10).Brunk E; Rothlisberger U Mixed Quantum Mechanical/Molecular Mechanical Molecular Dynamics Simulations of Biological Systems in Ground and Electronically Excited States. Chem. Rev 2015, 115, 6217–6263. [DOI] [PubMed] [Google Scholar]
  • (11).Chung LW; Sameera WMC; Ramozzi R; Page AJ; Hatanaka M; Petrova GP; Harris TV; Li X; Ke Z; Liu F et al. The ONIOM Method and Its Applications. Chem. Rev 2015, 115, 5678–5796. [DOI] [PubMed] [Google Scholar]
  • (12).Garcia-Viloca M; Gao J; Karplus M; Truhlar DG How Enzymes Work: Analysis by Modern Rate Theory and Computer Simulations. Science 2004, 303, 186–195. [DOI] [PubMed] [Google Scholar]
  • (13).Fernanda D; Amrein BA; David BN; Kamerlin SCL Recent Advances in QM/MM Free Energy Calculations using Reference Potentials. Biochim. Biophys. Acta 2015, 1850, 954–965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Cui Q; Pal T; Xie L Biomolecular QM/MM Simulations: What Are Some of the âĂIJBurning IssuesâĂİ? J. Phys. Chem. B 2021, 125, 689–702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Tuckerman ME Ab Initio Molecular Dynamics: Basic Concepts, Current Trends and Novel Applications. J. Phys. Condens. Matter 2002, 14, R1297–R1355. [Google Scholar]
  • (16).Marx D; Chandra A; Tuckerman ME Aqueous Basic Solutions: Hydroxide Solvation, Structural Diffusion, and Comparison to the Hydrated Proton. Chem. Rev 2010, 110, 2174–2216. [DOI] [PubMed] [Google Scholar]
  • (17).Sumner I; Iyengar SS Combining Quantum Wavepacket Ab Initio Molecular Dynamics with QM/MM and QM/QM Techniques: Implementation Blending ONIOM and Empirical Valence Bond Theory. J. Chem. Phys 2008, 129, 054109. [DOI] [PubMed] [Google Scholar]
  • (18).Marx D Proton Transfer 200 Years after von Grotthuss: Insights from Ab Initio Simulations. Chem. Phys. Chem 2010, 8, 1848–1870. [DOI] [PubMed] [Google Scholar]
  • (19).Ceriotti M; Fang W; Kusalik PG; McKenzie RH; Michaelides A; Morales MA; Markland TE Nuclear Quantum Effects in Water and Aqueous Systems: Experiment, Theory, and Current Challenges. Chem. Rev 2016, 116, 7529–7550. [DOI] [PubMed] [Google Scholar]
  • (20).Markland TE; Ceriotti M Nuclear Quantum Effects Enter the Mainstream. Nat. Rev. Chem 2018, 2, 0109. [Google Scholar]
  • (21).Dammak H; Hayoun M; Brieuc F; Geneste G Nuclear Quantum Effects in Molecular Dynamics Simulations. J. Phys. Conf. Ser 2018, 1136, 012014. [Google Scholar]
  • (22).Berger A; Ciardi G; Sidler D; Hamm P; Shalit A Impact of Nuclear Quantum Effects on the Structural Inhomogeneity of Liquid Water. Proc. Natl. Acad. Sci. U.S.A 2019, 116, 2458–2463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Feynman RP Atomic Theory of the λ Transition in Helium. Phys. Rev 1953, 91, 1291–1301. [Google Scholar]
  • (24).Chandler D; Wolynes PG Exploiting the Isomorphism between Quantum Theory and Classical Statistical Mechanics of Polyatomic Fluids. J. Chem. Phys 1981, 74, 4078–4095. [Google Scholar]
  • (25).Craig IR; Manolopoulos DE Quantum Statistics and Classical Mechanics: Real Time Correlation Functions from Ring Polymer Molecular Dynamics. J. Chem. Phys 2004, 121, 3368–3373. [DOI] [PubMed] [Google Scholar]
  • (26).Braams BJ; Manolopoulos DE On the Short-Time Limit of Ring Polymer Molecular Dynamics. J. Chem. Phys 2006, 125, 124105. [DOI] [PubMed] [Google Scholar]
  • (27).Cao J; Voth GA The Formulation of Quantum Statistical Mechanics Based on the Feynman Path Centroid Density. I. Equilibrium Properties. J. Chem. Phys 1994, 100, 5093–5105. [Google Scholar]
  • (28).Jang S; Voth GA A Derivation of Centroid Molecular Dynamics and Other Approximate Time Evolution Methods for Path Integral Centroid Variables. J. Chem. Phys 1999, 111, 2371–2384. [Google Scholar]
  • (29).Tuckerman ME; Berne BJ; Martyna GJ; Klein ML Efficient molecular dynamics and hybrid Monte Carlo algorithms for path integrals. J. Chem. Phys 1993, 99, 2796–2808. [Google Scholar]
  • (30).Markland TE; Manolopoulos DE An Efficient Ring Polymer Contraction Scheme for Imaginary Time Path Integral Simulations. J. Chem. Phys 2008, 129, 024105. [DOI] [PubMed] [Google Scholar]
  • (31).Cheng X; Herr JD; Steele RP Accelerating Ab Initio Path Integral Simulations via Imaginary Multiple-Timestepping. J. Chem. Theory Comput 2016, 12, 1627–1638. [DOI] [PubMed] [Google Scholar]
  • (32).Ceriotti M; Parrinello M; Markland TE; Manolopoulos DE Efficient Stochastic Thermostatting of Path Integral Molecular Dynamics. J. Chem. Phys 2010, 133, 124104. [DOI] [PubMed] [Google Scholar]
  • (33).Markland TE; Manolopoulos DE A Refined Ring Polymer Contraction Scheme for Systems with Electrostatic Interactions. Chem. Phys. Lett 2008, 464, 256–261. [Google Scholar]
  • (34).Kapil V; VandeVondele J; Ceriotti M Accurate Molecular Dynamics and Nuclear Quantum Effects at Low Cost by Multiple Steps in Real and Imaginary Time: Using Density Functional Theory to Accelerate Wavefunction Methods. J. Chem. Phys 2016, 144, 054111. [DOI] [PubMed] [Google Scholar]
  • (35).John C; Spura T; Habershon S; Kühne TD Quantum Ring-polymer Contraction Method: Including Nuclear Quantum Effects at No Additional Computational Cost in Comparison to Ab Initio Molecular Dynamics. Phys. Rev. E 2016, 93, 043305. [DOI] [PubMed] [Google Scholar]
  • (36).Marsalek O; Markland TE Ab Initio Molecular Dynamics with Nuclear Quantum Effects at Classical Cost: Ring Polymer Contraction for Density Functional Theory. J. Chem. Phys 2016, 144, 054112. [DOI] [PubMed] [Google Scholar]
  • (37).Torrie G; Valleau J Nonphysical Sampling Distributions in Monte Carlo Free-energy Estimation: Umbrella Sampling. J. Comput. Phys 1977, 23, 187–199. [Google Scholar]
  • (38).Babin V; Roland C; Sagui C Adaptively Biased Molecular Dynamics for Free Energy Calculations. J. Chem. Phys 2008, 128, 134101. [DOI] [PubMed] [Google Scholar]
  • (39).Laio A; Parrinello M Escaping Free-energy Minima. Proc. Natl. Acad. Sci. U.S.A 2002, 99, 12562–12566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (40).Quhe R; Nava M; Tiwary P; Parrinello M Path Integral Metadynamics. J. Chem. Theory Comput 2015, 11, 1383–1388. [DOI] [PubMed] [Google Scholar]
  • (41).Chodera JD; Swope WC; Pitera JW; Seok C; Dill KA Use of the Weighted Histogram Analysis Method for the Analysis of Simulated and Parallel Tempering Simulations. J. Chem. Theory Comput 2007, 3, 26–41. [DOI] [PubMed] [Google Scholar]
  • (42).Marx D; Tuckerman M; Hutter J; Parrinello M The Nature of the Hydrated Excess Proton in Water. Nature 1999, 397, 601–604. [Google Scholar]
  • (43).Kosugi K; Nakano H; Sato H SCC-DFTB-PIMD Method to Evaluate a Multidimensional Quantum Free-Energy Surface for a Proton-Transfer Reaction. J. Chem. Theory Comput 2019, 15, 4965–4973. [DOI] [PubMed] [Google Scholar]
  • (44).Cendagorta JR; Shen H; Bačić Z; Tuckerman ME Enhanced Sampling Path Integral Methods Using Neural Network Potential Energy Surfaces with Application to Diffusion in Hydrogen Hydrates. Adv. Theory Simul 2021, 4, 2000258. [Google Scholar]
  • (45).Li P; Jia X; Pan X; Shao Y; Mei Y Accelerated Computation of Free Energy Profile at Ab Initio Quantum Mechanical/Molecular Mechanics Accuracy via a SemiEmpirical Reference Potential. I. Weighted Thermodynamics Perturbation. J. Chem. Theory Comput 2018, 14, 5583–5596. [DOI] [PubMed] [Google Scholar]
  • (46).Pan X; Li P; Ho J; Pu J; Mei Y; Shao Y Accelerated Computation of Free Energy Profile at Ab Initio Quantum Mechanical/Molecular Mechanical Accuracy via a Semi-empirical Reference Potential. II. Recalibrating Semi-empirical Parameters with Force Matching. Phys. Chem. Chem. Phys 2019, 21, 20595–20605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (47).Hu W; Li P; Wang J-N; Xue Y; Mo Y; Zheng J; Pan X; Shao Y; Mei Y Accelerated Computation of Free Energy Profile at Ab Initio Quantum Mechanical/Molecular Mechanics Accuracy via a Semiempirical Reference Potential. 3. Gaussian Smoothing on Density-of-States. J. Chem. Theory Comput 2020, 16, 6814–6822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (48).Wang J-N; Liu W; Li P; Mo Y; Hu W; Zheng J; Pan X; Shao Y; Mei Y Accelerated Computation of Free Energy Profile at Ab Initio Quantum Mechanical/Molecular Mechanics Accuracy via a Semiempirical Reference Potential. 4. Adaptive QM/MM. J. Chem. Theory Comput 2021, 17, 1318–1325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (49).Rod TH; Ryde U Quantum Mechanical Free Energy Barrier for an Enzymatic Reaction. Phys. Rev. Lett 2005, 94, 138302. [DOI] [PubMed] [Google Scholar]
  • (50).Beierlein FR; Michel J; Essex JW A Simple QM/MM Approach for Capturing Polarization Effects in Protein-Ligand Binding Free Energy Calculations. J. Phys. Chem. B 2011, 115, 4911–4926. [DOI] [PubMed] [Google Scholar]
  • (51).König G; Boresch S Non-Boltzmann Sampling and Bennett’s Acceptance Ratio Method: How to Profit from Bending the Rules. J. Comput. Chem 2011, 32, 1082–1090. [DOI] [PubMed] [Google Scholar]
  • (52).Heimdal J; Ryde U Convergence of QM/MM Free-Energy Perturbations Based on Molecular-Mechanics or Semiempirical Simulations. Phys. Chem. Chem. Phys 2012, 14, 12592–12604. [DOI] [PubMed] [Google Scholar]
  • (53).Polyak I; Benighaus T; Boulanger E; Thiel W Quantum Mechanics/Molecular Mechanics Dual Hamiltonian Free Energy Perturbation. J. Chem. Phys 2013, 139, 064105. [DOI] [PubMed] [Google Scholar]
  • (54).König G; Hudson PS; Boresch S; Woodcock HL Multiscale Free Energy Simulations: An Efficient Method for Connecting Classical MD Simulations to QM or QM/MM Free Energies Using Non-Boltzmann Bennett Reweighting Schemes. J. Chem. Theory Comput 2014, 10, 1406–1419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (55).Hudson PS; Woodcock HL; 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, 4850–4856. [DOI] [PubMed] [Google Scholar]
  • (56).Hudson PS; White JK; Kearns FL; Hodoscek M; Boresch S; Woodcock HL Efficiently Computing Pathway Free Energies: New Approaches Based on Chain-of-Replica and Non-Boltzmann Bennett Reweighting Schemes. Biochim. Biophys. Acta, Gen. Subj 2015, 1850, 944–953. [DOI] [PubMed] [Google Scholar]
  • (57).Shen L; Wu J; Yang W Multiscale Quantum Mechanics/Molecular Mechanics Simulations with Neural Networks. J. Chem. Theory Comput 2016, 12, 4934–4946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Dybeck EC; König G; Brooks BR; Shirts MR Comparison of Methods To Reweight from Classical Molecular Simulations to QM/MM Potentials. J. Chem. Theory Comput 2016, 12, 1466–1480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (59).Kearns FL; Hudson PS; Woodcock HL; Boresch S Computing Converged Free Energy Differences between Levels of Theory via Nonequilibrium Work Methods: Challenges and Opportunities. J. Comput. Chem 2017, 38, 1376–1388. [DOI] [PubMed] [Google Scholar]
  • (60).Hudson PS; Boresch S; Rogers DM; Woodcock HL Accelerating QM/MM Free Energy Computations via Intramolecular Force Matching. J. Chem. Theory Comput 2018, 14, 6327–6335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (61).König G; Brooks BR; Thiel W; York DM On the Convergence of Multi-Scale Free Energy Simulations. Mol. Simulat 2018, 44, 1062–1081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (62).Hudson PS; Woodcock HL; Boresch S Use of Interaction Energies in QM/MM Free Energy Simulations. J. Chem. Theory Comput 2019, 15, 4632–4645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (63).Giese TJ; York DM Development of a Robust Indirect Approach for MM → QM Free Energy Calculations That Combines Force-Matched Reference Potential and BennettâĂŹs Acceptance Ratio Methods. J. Chem. Theory Comput 2019, 15, 5543–5562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (64).Piccini G; Parrinello M Accurate Quantum Chemical Free Energies at Affordable Cost. J. Phys. Chem. Lett 2019, 10, 3727–3731. [DOI] [PubMed] [Google Scholar]
  • (65).Chung S; Choi SM; Lee W; Cho KH; Rhee YM Free Energy Level Correction by Monte Carlo Resampling with Weighted Histogram Analysis Method. Chin. J. Chem. Phys 2020, 33, 183–195. [Google Scholar]
  • (66).Feynman RP; Hibbs AR Quantum Mechanics and Path Integrals; McGraw-Hill: New York, 1965. [Google Scholar]
  • (67).Ceperley DM Path Integrals in the Theory of Condensed Helium. Rev. Mod. Phys 1995, 67, 279–355. [Google Scholar]
  • (68).Barker JA A Quantum-statistical Monte Carlo Method: Path Integrals with Boundary Conditions. J. Chem. Phys 1979, 70, 2914–2918. [Google Scholar]
  • (69).Parrinello M; Rahman A Study of an F Center in Molten KCl. J. Chem. Phys 1984, 80, 860–867. [Google Scholar]
  • (70).Shirts MR; Chodera JD Statistically Optimal Analysis of Samples from Multiple Equilibrium States. J. Chem. Phys 2008, 129, 124105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (71).Stewart JJP Optimization of Parameters for Semi-Empirical Methods V: Modification of NDDO Approximations and Application to 70 Elements. J. Mol. Model 2007, 13, 1173–1213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (72).Becke AD Density-functional Exchange-energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098–3100. [DOI] [PubMed] [Google Scholar]
  • (73).Lee C; Yang W; Parr RG Development of the Colle-Salvetti Correlation-energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]
  • (74).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] [PubMed] [Google Scholar]
  • (75).White PB; Hong M 15N and 1H Solid-State NMR Investigation of a Canonical Low-Barrier Hydrogen-Bond Compound: 1,8-Bis(dimethylamino)naphthalene. J. Phys. Chem. B 2015, 119, 11581–11589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (76).Wozniak K; He H; Klinowski J; Barr TL; Hardcastle SE ESCA, Solid-State NMR, and X-Ray Diffraction Studies of Perisubstituted Naphthalene Derivatives. J. Phys. Chem. B 1996, 100, 11408–11419. [Google Scholar]
  • (77).Jorgensen WL; Chandresekhar J; Madura JD; Impey RW; Klein ML Application of the Multimolecule and Multiconformational RESP Methodology to Biopolymers: Charge Derivation for DNA, RNA, and Proteins. J. Chem. Phys 1983, 79, 926–935. [Google Scholar]
  • (78).Case DA; Ben-Shalom IY; Brozell SR; Cerutti DS; Cheatham TE III; Cruzeiro VWD; Darden TA; Duke RE; Ghoreishi D; Gilson MK et al. AMBER 19, University of California, San Francisco. 2019. [Google Scholar]
  • (79).Wang J; Wolf RM; Caldwell JW; Kollman PA; Case DA Development and Testing of a General Amber Force Field. J. Comput. Chem 2004, 25, 1157–1174. [DOI] [PubMed] [Google Scholar]
  • (80).Ryckaert J-P; Ciccotti G; Berendsen HJ Numerical Integration of the Cartesian Equations of Motion of a System with Constraints: Molecular Dynamics of N-Alkanes. J. Comput. Phys 1977, 23, 327 – 341. [Google Scholar]
  • (81).Zhang Z; Liu X; Chen Z; Zheng H; Yan K; Liu J A Unified Thermostat Scheme for Efficient Configurational Sampling for Classical/Quantum Canonical Ensembles via Molecular Dynamics. J. Chem. Phys 2017, 147, 034109. [DOI] [PubMed] [Google Scholar]
  • (82).Liu J; Li D; Liu X A Simple and Accurate Algorithm for Path Integral Molecular Dynamics with the Langevin Thermostat. J. Chem. Phys 2016, 145, 024103. [DOI] [PubMed] [Google Scholar]
  • (83).Rasmussen C; Williams C Gaussian Processes for Machine Learning; MIT Press, 2006; pp 7–32. [Google Scholar]
  • (84).Zhou S; Wang L Symmetry and 1H NMR Chemical Shifts of Short Hydrogen Bonds: Impact of Electronic and Nuclear Quantum Effects. Phys. Chem. Chem. Phys 2020, 22, 4884–4895. [DOI] [PubMed] [Google Scholar]
  • (85).Ceriotti M; Brain GAR; Riordan O; Manolopoulos DE The Inefficiency of Re-weighted Sampling and the Curse of System Size in High-order Path Integration. Proc. R. Soc. London, Ser. A 2012, 468, 2–17. [Google Scholar]
  • (86).Boresch S; Woodcock HL Convergence of Single-step Free Energy Perturbation. Mol. Phys 2017, 115, 1200–1213. [Google Scholar]
  • (87).Pan X; Yang J; Van R; Epifanovsky E; Ho J; Huang J; Pu J; Mei Y; Nam K; Shao Y Machine-Learning-Assisted Free Energy Simulation of Solution-Phase and Enzyme Reactions. J. Chem. Theory Comput 2021, 17, 5745–5758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (88).Zhou Y; Pu J Reaction Path Force Matching: A New Strategy of Fitting Specific Reaction Parameters for Semiempirical Methods in Combined QM/MM Simulations. J. Chem. Theory Comput 2014, 10, 3038–3054. [DOI] [PubMed] [Google Scholar]
  • (89).Wu J; Shen L; Yang W Internal Force Corrections with Machine Learning for Quantum Mechanics/Molecular Mechanics Simulations. J. Chem. Phys 2017, 147, 161732. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (90).Shen L; Yang W Molecular Dynamics Simulations with Quantum Mechanics/Molecular Mechanics and Adaptive Neural Networks. J. Chem. Theory Comput 2018, 14, 1442–1455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (91).Kim B; Shao Y; Pu J Doubly Polarized QM/MM with Machine Learning Chaperone Polarizability. J. Chem. Theory Comput DOI: 10.1021/acs.jctc.1c00567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (92).Böselt L; Thürlemann M; Riniker S Machine Learning in QM/MM Molecular Dynamics Simulations of Condensed-Phase Systems. J. Chem. Theory Comput 2021, 17, 2641–2658. [DOI] [PubMed] [Google Scholar]
  • (93).Zeng J; Giese TJ; Ekesan S; York DM Development of Range-Corrected Deep Learning Potentials for Fast, Accurate Quantum Mechanical/Molecular Mechanical Simulations of Chemical Reactions in Solution. J. Chem. Theory Comput 2021, 17, 6993–7009. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

supporting information

RESOURCES