
Citation: Takumi Washio, Akihiro Fujii, Toshiaki Hisada. On random force correction for large time steps in semi-implicitly discretized overdamped Langevin equations[J]. AIMS Mathematics, 2024, 9(8): 20793-20810. doi: 10.3934/math.20241011
[1] | Xiao Qi, Mejdi Azaiez, Can Huang, Chuanju Xu . An efficient numerical approach for stochastic evolution PDEs driven by random diffusion coefficients and multiplicative noise. AIMS Mathematics, 2022, 7(12): 20684-20710. doi: 10.3934/math.20221134 |
[2] | Lanyin Sun, Kunkun Pang . Numerical solution of unsteady elastic equations with C-Bézier basis functions. AIMS Mathematics, 2024, 9(1): 702-722. doi: 10.3934/math.2024036 |
[3] | Hyam Abboud, Clara Al Kosseifi, Jean-Paul Chehab . Stabilized bi-grid projection methods in finite elements for the 2D incompressible Navier-Stokes equations. AIMS Mathematics, 2018, 3(4): 485-513. doi: 10.3934/Math.2018.4.485 |
[4] | Tingting Ma, Yuehua He . An efficient linearly-implicit energy-preserving scheme with fast solver for the fractional nonlinear wave equation. AIMS Mathematics, 2023, 8(11): 26574-26589. doi: 10.3934/math.20231358 |
[5] | Chaeyoung Lee, Seokjun Ham, Youngjin Hwang, Soobin Kwak, Junseok Kim . An explicit fourth-order accurate compact method for the Allen-Cahn equation. AIMS Mathematics, 2024, 9(1): 735-762. doi: 10.3934/math.2024038 |
[6] | Zeliha Korpinar, Mustafa Inc, Dumitru Baleanu . On the fractional model of Fokker-Planck equations with two different operator. AIMS Mathematics, 2020, 5(1): 236-248. doi: 10.3934/math.2020015 |
[7] | Paola F. Antonietti, Benoît Merlet, Morgan Pierre, Marco Verani . Convergence to equilibrium for a second-order time semi-discretization ofthe Cahn-Hilliard equation. AIMS Mathematics, 2016, 1(3): 178-194. doi: 10.3934/Math.2016.3.178 |
[8] | Zheng Kou, Saeed Kosari . On a generalization of fractional Langevin equation with boundary conditions. AIMS Mathematics, 2022, 7(1): 1333-1345. doi: 10.3934/math.2022079 |
[9] | Jing Li, Linlin Dai, Kamran, Waqas Nazeer . Numerical solution of multi-term time fractional wave diffusion equation using transform based local meshless method and quadrature. AIMS Mathematics, 2020, 5(6): 5813-5838. doi: 10.3934/math.2020373 |
[10] | Xiaoming Wang, Rizwan Rizwan, Jung Rey Lee, Akbar Zada, Syed Omar Shah . Existence, uniqueness and Ulam's stabilities for a class of implicit impulsive Langevin equation with Hilfer fractional derivatives. AIMS Mathematics, 2021, 6(5): 4915-4929. doi: 10.3934/math.2021288 |
The limited time step size in molecular dynamics (MD) simulations is a major barrier to simulating long-time behaviors of various biological processes, even if coarse-grained molecular models are applied. The limitation originates from the large stiffnesses of strong bonds. In most cases, these strong interactions are necessary to maintain the basic structure of molecules, but the associated small fluctuations do not seem to be essential for molecular deformations of interest. In continuum mechanics, we can overcome the barrier with the use of implicit time integration schemes [1], where the Hessian matrix of the potential function is used to prevent the amplification of oscillations associated with its large eigenvalues. However, the complicated landscapes of MD potential functions hamper the use of implicit schemes because Newton iteration is unstable [2]. In our previous work [3], we introduced a semi-implicit Hessian correction scheme using only the positive parts of Hessian matrices of elemental potentials to avoid instabilities. This scheme allowed us to use time steps 50–800 times larger than those in the explicit Euler-Maruyama (EM) time integration scheme [4]. However, further increases in the time step created considerable errors in the conformational distributions even though the molecular structures were not destroyed.
In this paper, we focus on the damping effect of the Hessian matrix on the fluctuations caused by random forces. In the overdamped Langevin (OL) equation, the increment of the random variable X is given by
GdX=F(X)dt+Ξdt, | (1.1) |
where G is a positive-definite matrix representing the friction, F(X) is the force as a function of X, and Ξ is the random force satisfying ⟨Ξdt⟩= 0. We use differential forms and their discrete approximations according to the Itô calculus, as in (1.1), throughout this article. The increment of the probability density function is determined by the averages:
{⟨dX⟩=G−1F(X)dt,⟨dX⊗dX⟩=G−1⟨Ξdt⊗Ξdt⟩G−1. | (1.2) |
Here, the first equation represents advection by the force F(X). The second equation represents the diffusion of the probability density function. Therefore, the random force is the source of diffusion, which is damped by the friction. In the explicit EM scheme, (1.1) is directly discretized for a finite time interval has
GΔX=F(X(t))h+Ξhh | (1.3) |
to obtain the increment ΔX=X(t+h)−X(h). Therefore, the impact of the friction coefficient G on the fluctuation of the numerical solution is estimated by replacing dt with h in (1.2). If F(X+ΔX) can be approximated with a positive-semidefinite matrix K(X) as
F(X+ΔX) F(X)−K(X)ΔX, | (1.4) |
the following semi-implicit scheme may be applied to enable the use of a large time interval h:
(G+hK(X(t)))ΔX=F(X(t))h+Ξhh. | (1.5) |
In this case, the variance is given by
⟨ΔX⊗ΔX⟩=(G+hK)−1⟨Ξhh⊗Ξhh⟩(G+hK)−1. | (1.6) |
If there are eigenvalues of hK that are much greater than those of G, the damping by hK may underestimate the diffusion. In this study, we pursue a correction scheme for the random force in the semi-implicit EM scheme to shift the diffusion toward the desired direction and magnitude. Comparing (1.3) and (1.5), the coefficient G+hK(X(t)) on the left-hand side of (1.5) can be regarded as a modified friction coefficient. Therefore, a natural strategy would appear to involve correcting the random force Ξh on the right-hand side to be consistent with the modified friction.
We begin our analysis with a linear overdamped Langevin (OL) equation with a positive semidefinite matrix K. Through comparison of the numerical solution obtained by the implicit scheme with the analytical solution, we find an appropriate correction to the random force of the form hK, where h is the time interval of the temporal discretization. Next, we extend the idea of random force correction to nonlinear OL equations using the Hessian matrix instead of K in the linear case. We derive a discrete Fokker-Planck (FP) equation from the semi-implicitly discretized OL equation with a random force correction. Here, we see that the correction of the random force corresponds to the addition of artificial friction associated with the Hessian matrix. Finally, we validate the proposed semi-implicit scheme on the entropic elasticity problem of the freely jointed chain model.
Grønbech-jensen and Doniach [5] proposed a semi-implicit scheme for the OL equation in which strong bonded interactions are constrained. Though their approach is limited to infinitely stiff bonds, our approach can analyze the effects of the magnitude of strong bonded interactions on the stochastic behavior of molecules, as shown in our numerical experiments. Sweet et al. [6] applied a friction matrix constructed by normal mode partitioning of a mass reweighted Hessian matrix for the underdamped Langevin (UL) equation. Their scheme requires the computation of all eigenvectors of the mass reweighted Hessian matrix to perform the projections on the two subspaces spanned by the low- and high-frequency modes, whereas our scheme requires only the inversion in solving (1.5) and the construction of the additional random forces constructed from the elemental Hessian matrices of the individual strong bonds. This configuration enables faster computations.
To determine the essence of the discrepancy of implicit scheme for large times, we start our discussion with a linear OL equation:
γdX(t)dt=−KX(t)+Ξ(t), | (2.1) |
where γ is the friction coefficient, K is a positive semidefinite matrix, and Ξ(t) is the random force vector satisfying
⟨Ξ(s)ds⊗Ξ(u)du⟩=δ(s−u)2kBTγIds | (2.2) |
with the Boltzmann constant kB and the temperature T. Here, I is the identity matrix, and the tensor product a⊗b of two vectors a and b can be identified with the square matrix abT. For a given initial solution X(t) at t, the analytical solution X(t+h) at t+h is as follows:
X(t+h)=exp(hγK)−1{γ−1∫h0exp(sγK)Ξ(t+s)ds+X(t)}. | (2.3) |
Here, the variance of the integrated random force on the right-hand side is calculated using (2.2) to be
⟨∫h0exp(sγK)Ξ(t+s)ds⊗∫h0exp(uγK)Ξ(t+u)du⟩=2kBTγ∫h0exp(2sγK)ds= kBTγ2K−1(exp(2hγK)−I) =2kBTγh(I+hγK+∑l≥22l(l+1)!(hγK)l). | (2.4) |
Here, we assume x−1(expx−1) = 1 at x = 0 for the null space of K. Let us analyze the relationship between the exact solution at t+h in (2.3) and the temporally discretized solution of the implicit scheme defined as
γˆX(t+h)−ˆX(t)h=−KˆX(t+h)+ˆΞh(t), | (2.5) |
where ˆΞh(t) is the discrete random force at t given later. The above equation can be rewritten as
ˆX(t+h)=(I+hγK)−1{γ−1ˆΞh(t)h+ˆX(t)}. | (2.6) |
Here, the matrix (I+γ−1hK) on the right-hand side can be regarded as the first order Taylor expansion of exp(γ−1hK) in (2.3). Furthermore, if we additionally take the first order Taylor expansion of γ−1hK in the last term of (2.4), the following condition to be imposed on the random force ˆΞh(t) is derived:
⟨ˆΞh(t)h⊗ˆΞh(t)h⟩=2kBTh(γI+hK). | (2.7) |
Namely, the addition of hK to the identity matrix γI, which is used in the explicit EM scheme, may improve the accuracy of implicit schemes. This suggests that the random force should be corrected using the Hessian matrix in semi-implicit schemes for nonlinear problems as well. In other words, the proposed implicit scheme can be regarded as the EM scheme with the random force satisfies (2.7), where hK plays the role of the artificial friction.
Note that the accuracy of the approximation of (2.4) to (2.7) is ensured only for the low frequency components of small eigenvalues of γ−1hK. However, we are interested in using time steps much larger than that of the stability condition for explicit schemes, i.e.,
hγρ(K)≫1, | (2.8) |
where ρ(K) is the spectral radius of K. According to (2.4), we cannot expect high accuracy for the high-frequency components for a single time step. However, if the high frequency components are damped over k steps (k≫1), the temporal change of the stochastic distribution may be correctly obtained. To see this damping effect, we compare the numerical solution after the k steps to the analytical solution. First, we split the analytical solution at t+kh into the stochastic part XS and the deterministic part XD, where X=XS+XD, and the two parts are given by
{XS(t+kh)=exp(khγK)−1γ−1∫kh0exp(sγK)Ξ(t+s)ds,XD(t+kh)=exp(khγK)−1X(t). | (2.9) |
Here, the variance of XS(t+kh) is calculated as
⟨XS(t+kh)⊗XS(t+kh)⟩=kBTK−1(I−exp(−2khγK))=2kBTkhγ(2khγK)−1(I−exp(−2khγK)). | (2.10) |
Here, we assume x−1(1−exp(−x)) = 1 at x = 0 for the null space of K. Similarly, the numerical solution after k steps is split as ˆXk=ˆXS,k+ˆXD,k, where the stochastic part ˆXS,k and deterministic part ˆXD,k are given by
{ˆXS,k=γ−1k−1∑l=0(I+hγK)−(k−l)ˆΞh,lh,ˆXD,k=(I+hγK)−kˆX(t). | (2.11) |
Here, ˆΞh,l=ˆΞh(t+lh) is the random force applied at the (l+1)-th step. From the decorrelation of these random forces and (2.7), the variance of ˆXS,k is
⟨ˆXS,k⊗ˆXS,k⟩=γ−2k−1∑l=0(1+hγK)−(k−l)⟨ˆΞh,lh⊗ˆΞh,lh⟩(I+hγK)−(k−l) =2kBTkhγ1kk−1∑l=0(I+hγK)−2(k−l)+1. | (2.12) |
Thus, the difference between (2.10) and (2.12) can be evaluated by taking the difference of the following two functions, where x=γ−1hK:
{vk(x)=12kx(1−exp(−2kx)),ˆvk(x)=1kk−1∑l=0(1+x)−2(k−l)+1. | (2.13) |
The two functions and their differences are depicted in Figure 1A and B for k=1, 10, and 100. The damping effects for x≫o(1) are clearly shown for both the functions and the differences. Similarly, the deterministic parts in (2.9) and (2.11) can be evaluated using the functions defined by
{dk(x)=exp(−kx),ˆdk(x)=(1+x)−k. | (2.14) |
In Figure 1C and D, the same trends as in the stochastic parts are observed.
A comparison with the EM scheme
{γˇX(t+h)−ˇX(t)h=−KˇX(t)+Ξh(t),⟨Ξh(t)h⊗Ξh(t)h⟩=2kBTγhI, | (2.15) |
will further elucidate the effects of the implicit scheme. In this case, the stochastic and deterministic parts of the numerical solution are represented as
ˇXS,k=γ−1k−1∑l=0(I−hγK)k−lΞh,lh,ˇXD,k=(I−hγK)kˇX(t). | (2.16) |
The variance of the stochastic part is calculated as
⟨ˇXS,k⊗ˇXS,k⟩=2kBTkhγ1kk−1∑l=0(I−hγK)2(k−l). | (2.17) |
In Figure 2, the following functions representing the evolution of solutions are compared with their analytical counterparts for 0<x<2, where convergence is ensured:
{ˇvk=γ−1k−1∑l=0(1−x)2(k−l),ˇdk=(1−x)k. | (2.18) |
As in the implicit case, the numerical error around x=O(1) is damped as k increases.
We consider numerical time integration schemes of the nonlinear OL equation for n particles in three-dimensional space:
GdX(t)dt=−dUdX(X)+Ξ(t), | (3.1) |
where G is a 3n×3npositive diagonal matrix representing the friction coefficients imposed on the n particles, U(X) is a potential function, and Ξ(t) is a random vector satisfying
⟨Ξ(t)dt⊗Ξ(t)dt⟩=2kBTGdt. | (3.2) |
The potential function consists of bonded and nonbonded interactions, and the thermal effects of the implicit solvent are approximated by the random force. The limitation on the time step size comes from the large eigenvalues of the Hessian matrix H=∂2U/∂X2. These large eigenvalues originate from the strong bonded interactions. Because the inversion of a banded matrix consisting of elemental Hessian matrices of the strong bonded interactions is computationally efficient, the use of a semi-implicit method is natural. In the following, we start with the EM scheme, and modify the friction coefficient and the random force to enable a large time interval.
In accordance with the Ito integral for a constant time interval h as in the EM scheme, we must update the particle coordinates Xh from t to t+h as follows:
GXh(t+h)−Xh(t)h=−dUdX(Xh(t))+Ξh(t), | (3.3) |
where the right-hand side involves the force −dU/dX at t and the discrete random force satisfies
⟨Ξh(t)h⊗Ξh(t)h⟩=2kBTGh. | (3.4) |
When applying the above scheme to MD simulations, we encounter a severe restriction of the time interval h owing to the stability condition:
ρ(I−hG−1∂2U∂X2)≤1. | (3.5) |
In our previous work [3], we introduced a semi-implicit Hessian correction scheme (SimHec) with the intention of replacing −dUdX(Xh(t)) in (3.3) by a linear approximation of −dUdX(Xh(t+h)) as follows:
(G+h˜H(˜Xh(t)))˜Xh(t+h)−˜Xh(t)h=−dUdX(˜Xh(t))+Ξh(t). | (3.6) |
Here, ˜H(X) is a 3n×3n symmetric matrix designed to approximate the positive part of the Hessian matrix of U at X. Based on the derivation of (2.7), we regard the coefficient h˜H(˜Xh(t)) on the left-hand side of (3.6) as the artificial friction and apply the corrected random force as follows:
(G+h˜H(ˆXh(t)))ˆXh(t+h)−ˆXh(t)h=−dUdX(ˆXh(t))+Ξh(t)+˜Ξh(t), | (3.7) |
where the additional random force ˜Ξh(t) satisfies the conditions
{⟨Ξh(t)h⊗˜Ξh(t)h⟩=0,⟨˜Ξh(t)h⊗˜Ξh(t)h⟩=2kBTh˜H(ˆXh(t))h. | (3.8) |
Here, the first condition indicates that the additional random force ˜Ξh must be uncorrelated with the original random force Ξh(t), and the second condition corresponds to the consistency of the additional random force and the artificial friction h˜H. From these conditions, we can easily derive the consistency condition of the total friction ˆGh(ˆXh(t))=G+h˜H(ˆXh(t)) and the total random force ˆΞh(t)=Ξh(t)+˜Ξh(t), which is
⟨ˆΞh(t)h⊗ˆΞh(t)h⟩=2kBTˆGh(ˆXh(t))h. | (3.9) |
In the following, we discuss the meaning of this consistency condition in the context of the probability density. Hereafter, we refer to this method as the "semi-implicit Hessian correction scheme with random force correction" (SimHec-RC).
As shown in the literature [7,8,9], the property (3.2) of the random force is key to deriving the FP equation when applying the Ito integral to the OL Eq (3.1). Here, we trace the process from the OL equation to the FP equation based on the discrete temporal integration of SimHec-RC.
In SimHec-RC, the update ΔˆXh=ˆXh(t+h)−ˆXh(t) is performed as follows:
ΔˆXh=−ˆGh(ˆXh(t))−1dUdX(ˆXh(t))h+ˆGh(ˆXh(t))−1ˆΞh(t)h. | (4.1) |
Therefore, we obtain
⟨ΔˆXh⟩=−ˆGh(ˆXh(t))−1dUdX(ˆXh(t))h,⟨ΔˆXh⊗ΔˆXh⟩=2kBTˆGh(ˆXh(t))−1h+O(h2). | (4.2) |
Here, we have derived the second equation assuming (3.9) as follows:
⟨ˆGh(ˆXh)−1ˆΞhh⊗ˆGh(ˆXh)−1ˆΞhh⟩=ˆGh(ˆXh)−1⟨ˆΞhh⊗ˆΞhh⟩ˆGh(ˆXh)−1=2kBThˆGh(ˆXh)−1. | (4.3) |
Note that the higher order products of ˆGh(ˆXh)−1ˆΞhh are O(h2). If we replace the modified random force ˆΞh(t) with the original random force Ξh(t) as in the original SimHec in (3.6), the second term is replaced by
⟨Δ˜Xh⊗Δ˜Xh⟩=2kBThˆGh(˜Xh(t))−1GˆGh(˜Xh(t))−1+O(h2). | (4.4) |
Here, ⟨Δ˜Xh⊗Δ˜Xh⟩ differs by the factor GˆGh(˜Xh(t))−1from ⟨ΔˆXh⊗ΔˆXh⟩ in (4.2). Below, we evaluate the impact of this difference on the probability density functions produced by these discrete stochastic processes.
From (4.2), we can derive the discrete version of Ito's lemma [10] for any smooth function f using Taylor expansion:
⟨f(ˆXh(t)+ΔˆXh)⟩−f(ˆXh(t))=⟨f(ˆXh(t)+ΔˆXh)−f(ˆXh(t))⟩ =⟨dfdX(ˆXh(t)):ΔˆXh+12d2fdX2(ˆXh(t)):ΔˆXh⊗ΔˆXh⟩+O(h2) =hˆGh(ˆXh(t))−1:(−dfdX(ˆXh(t))⊗dUdX(ˆXh(t))+kBTd2fdX2(ˆXh(t))) +O(h2). | (4.5) |
Here, the average ⟨∗⟩ is taken over the discrete stochastic processes from t to t+h, and the symbol –8–8represents the dot product of two vectors or two matrices. Let ˆPh(ˆXh,t) be the probability density function of SimHec-RC in (3.7). Then, we have the following relationship between the integrals over the conformation spaces ˆΩh(t) and ˆΩh(t+h) at times t and t+h, respectively:
∫⟨f(ˆXh(t)+ΔˆXh)⟩ˆPh(ˆXh(t),t)dˆΩh(t)=∫f(ˆXh(t+h))ˆPh(ˆXh(t+h),t+h)dˆΩh(t+h). | (4.6) |
Here, the averaging operation ⟨f(ˆXh+ΔˆXh)⟩ on the left-hand side is taken over ΔˆXh. By using Eq (4.6) followed by Eq (4.5) and applying integration by parts, we obtain the following equations:
∫f(ˆXh)1h(ˆPh(ˆXh,t+h)−ˆPh(ˆXh,t))dˆΩh=1h∫(⟨f(ˆXh+ΔˆXh)⟩−f(ˆXh))ˆPh(ˆXh,t)dˆΩh=∫{ˆGh(ˆXh)−1:(−dfdX(ˆXh)⊗dUdX(ˆXh)+kBTd2fdX2(ˆXh))}ˆPh(ˆXh,t)dˆΩh+O(h)=∫f(ˆXh){ddX:(ˆGh(X)−1dUdX(X)ˆPh(X,t)+kBTddX:(ˆGh(X)−1ˆPh(X,t)))|X=ˆXh}dˆΩh+O(h). | (4.7) |
Because the above equation holds for any smooth function f, we obtain
1h(ˆPh(X,t+h)−ˆPh(X,t))=ddX:{ˆGh(X)−1dUdX(X)ˆPh(X,t)+kBTddX:(ˆGh(X)−1ˆPh(X,t))}+O(h). | (4.8) |
This is a discrete version of the FP equation in which ˆGh(X) corresponds to the friction coefficients. Similarly, from (4.4), we obtain the following equation for the probability density function ˜Ph(X,t) for the stochastic process given by SimHec in (3.6) without the random force correction:
1h(˜Ph(X,t+h)−˜Ph(X,t))=ddX:{ˆGh(X)−1dUdX(X)˜Ph(X,t)+kBTddX:(ˆGh(X)−1GˆGh(X)−1˜Ph(X,t))}+O(h). | (4.9) |
Here, we see that the dispersion is overdamped by a factor of ˆGh(˜X(t))−1G compared with (4.8). The advantage of the random force correction is clear when ˆGh is a constant matrix. The term inside the brackets on the right-hand side of (4.8) is rewritten as
ˆGh(X)−1dUdX(X)ˆPh(X,t)+kBTddX:(ˆGh(X)−1ˆPh(X,t))=ˆGh(X)−1{dUdX(X)ˆPh(X,t)+kBTdˆPhdX(X,t)}+kBT{ddX:ˆGh−1}(X)ˆPh(X,t). | (4.10) |
Thus, we see that the probability density of the FP equation converges to the Boltzmann distribution, i.e.,
limt→∞ˆPh(X,t)∝exp(−U(X)kBT). | (4.11) |
Equation (4.10) also suggests that the accuracy of the obtained time transients of the distribution depends on the magnitude of ‖ddX:Gh−1‖. Therefore, it may be important to minimize ‖ddX˜H‖ in the construction of ˜H. Physically, ddX:Gh−1≠0 means that the thermal effect of the solvent depends on the conformation X. Thus, the probability density does not converge to the Boltzmann distribution in thermal equilibrium for a finite h.
Here, we consider combining the ˜H(X) and ˜Ξh that satisfy the conditions in (3.8). In general, the potential function U(X) in MD is given as a sum of elemental potentials between two, three, or four particles, as follows:
U(X)=∑kVk(PkX), | (5.1) |
where Pk represents projection onto the subspace Ωk of the subset of particles for which the potential Vk is defined. For every potential Vk, if we can construct ˜Hk and ˜Ξk,h on Ωk such that the conditions
{⟨Ξh(t)h⊗˜Ξk,h(t)h⟩=0,⟨˜Ξh,k(t)h⊗˜Ξh,l(t)h⟩=δkl2kBTh˜Hk(PkX(t))h, | (5.2) |
are fulfilled, then it is straightforward to show that the sums
{˜H(X(t))=∑k˜Hk(PkX(t)),˜Ξh(t)=∑k˜Ξk,h(t), | (5.3) |
satisfy the conditions in (3.8). Here, we identify ˜Hk and ˜Ξk,h on the subspace Ωk with PkT˜HkPk and PkT˜Ξk,h on the total space, respectively, for the sake of simplicity. We can realize the uncorrelatedness of random forces between different elemental potentials using sequences of uncorrelated random numbers. Therefore, the remaining problems are (ⅰ) constructing ˜Hk for Vk, and (ⅱ) constructing a ˜Ξk,h(t) that fulfills the condition:
⟨˜Ξh,k(t)h⊗˜Ξh,k(t)h⟩=2kBTh˜Hk(Xk(t))h,Xk(t)=PkX(t). | (5.4) |
As an example, we consider the potential V(x1,x2)=φ(r1,2), which is a function of the distance r1,2=‖x2−x1‖. In this case, the Hessian matrix is
H1,2=[∂2V∂x12∂2V∂x1x2∂2V∂x2x1∂2V∂x22]=¨φ(r1,2)A1,2+˙φ(r1,2)r1,2B1,2, | (5.5) |
where the matrices A1,2 and B1,2 are defined as
A1,2=[n⊗n−n⊗n−n⊗nn⊗n],B1,2=[I3−n⊗n−I3+n⊗n−I3+n⊗nI3−n⊗n], | (5.6) |
with the unit vector
n=x2−x1‖x2−x1‖. | (5.7) |
Note that the matrices A1,2 and B1,2 satisfy
{A1,22=2A1,2,B1,22=2B1,2,A1,2B1,2=B1,2A1,2=0. | (5.8) |
Therefore, for two given positive coefficients ˜α and ˜β which may depend on r1,2, a pair of ˜H1,2 and ˜Ξh satisfying Eq (5.4) can be given as
{˜H1,2=˜αA1,2+˜βB1,2,˜Ξh={√kBT˜αA1,2+√kBT˜βB1,2}[ξG1ξG2], | (5.9) |
with the normalized Gaussian noise satisfying
⟨[ξG1ξG2]⊗[ξG1ξG2]⟩=I3×2. | (5.10) |
From (5.8), the eigenspaces of H1,2 are composed of those of A1,2 and B1,2. The null space is spanned by the parallel displacements [dd],∀d. The eigenspace of A1,2 is spanned by the vector [n−n], which corresponds to oscillation along n. The eigenspace of B1,2 is spanned by the vectors [m−m],m⊥n, which corresponds to the rotation about the center (x1+x2)/2 (Figure 3). Because the nonzero eigenvalues of A1,2 and B1,2 are equal to 2, the nonzero eigenvalues of H1,2 are 2¨φ(r1,2) and 2˙φ(r1,2)/r1,2. In a case of a bond between a pair of adjacent particles (i,i+1),
φ(r)=cB(r−r0)2, | (5.11) |
The coefficient of B1,2 in (5.2) is negative for r<r0. In this paper, we replace (r−r0)/r with a nonnegative constant bB as follows:
{˜αB(r)=2cB,˜βB(r)=2cBmax((r−r0)/r,bB). | (5.12) |
Hereafter, we refer to the parameter bB as SimHec(bB) or SimHec-RC(bB). In the following, a nearly optimal parameter is chosen for each approach.
To test the efficiency and the accuracy of SimHEC-RC, we simulated the entropic elasticity of a freely jointed chain consisting of n particles (n=100) with potential
U(X)=n−1∑i=1cB(‖xi−xi+1‖r+1)2. | (6.1) |
We adopted the parameters of the coarse-grained protein model CafeMol [11], which only considers the position of a single bead at the Cα position for each residue in the polymer chain. A friction coefficient of γi=168.7CafeMolTime⋅ kcal/(mol Å2) was used. This value was derived by assuming the viscosity of water: 8×10−4kg/(m s), and the radius r+1 of the amino acids: 3.82 Å. The time unit in CafeMol (CafeMolTime) is approximately 49 fs. A value of cB=110.4 kcal/(mol Å2) was used for all bonds; the limitation on the time step size of the EM scheme due to the bond stiffness was estimated as h≤γ/2cB=0.76 CafeMolTime.
First, we examined the accuracy for a transient problem where one end of the straight chain was released (Figure 4A). In this numerical test, one end x1 was always fixed, and the other end xn was released at t=0. The trajectories of the coordinates of xn in the stretched direction were analyzed using the numerical results of 64 samples. We observed that the time transients of edge coordinate averages with 64 samples and 256 samples were almost identical in SimHec-RC with h=100. Therefore, we performed the comparisons with 64 samples. The EM scheme with h=0.5 shows a severe discrepancy (Figure 4B). The SimHec and SimHec-RC schemes were compared with the EM scheme with h=0.125 (Figure 4C). The accuracy of SimHec deteriorated as the time step h increased. The shortening speed decreased with excessive friction as shown in (4.4) and (4.9). However, SimHec-RC retained its accuracy by correcting the random force.
To confirm the accuracy for multiple steps as estimated in Figures 1 and 2, we examined the distributions of (x,y)=(rcosθ,rsinθ), where r and θ are defined using the changes of xi+nd−xi from t to t+kh as follows:
{r=‖xi+nd(t+kh)−xi(t+kh)‖,cosθ=xi+nd(t+kh)−xi(t+kh)‖xi+nd(t+kh)−xi(t+kh)‖⋅xi+nd(t)−xi(t)‖xi+nd(t)−xi(t)‖,0≤θ<π. | (6.2) |
In Figure 5, we compared the distributions superposed for all pairs (i,i+nd) with nd = 10 obtained using the EM scheme with h=0.125 and SimHec-RC(0.01) with h=100. There are substantial differences between the k=1 and 10 cases of SimHec-RC and their counterparts, namely, the k=800 and 8000 cases of the EM scheme. However, as expected from Figure 1A and B, the difference is diminished for k=100.
Note that hk = 104CafeMolTime ~ 0.5 ns is much smaller than the total time 107 CafeMolTime ~ 0.5 μs, which needed to be reduced to close to zero (Figure 4).
Next, we evaluated the force acting on the end xn trapped by the spring at the position pL=[nr+1,0,0]T as
U(X)=n−1∑i=1cB(‖xi−xi+1‖r+1)2+Ka‖xn−pL‖2. | (6.3) |
In the numerical experiment, we adopted Ka = 0.1 kcal/(mol Å2). In Figure 6A, the average values of the force
⟨f1⟩=2Ka⟨nr+1−xn,1⟩ | (6.4) |
computed over the time range [0,5×109]CafeMolTime (~[0, 0.25]s) for a single process are compared with the values calculated using the explicit scheme with h=0.25. The EM scheme with h=0.5 shows a severe discrepancy (Figure 6B) as in the case of shortening of the free end. In Figure 6C, we compared the average force as computed by SimHec and SimHec-RC, where cB and Ka were increased by factors of one hundred and ten, respectively, with the theoretical values in the limit of cB→∞:
⟨f1⟩∞(L)=kBTr+1F−1(Lnr+1), | (6.5) |
where the Langevin function F is given by
F(x)=ex+e−xex−e−x−1x. | (6.6) |
To assess the computational efficiency, we compared the overhead of SimHec-RC with that of EM. The overheads consist of the cost of computing the elemental Hessian matrices, the random force corrections, and the linear solutions. The computation times per time step for EM, SimHec, and SimHec-RC were 3.2, 16.9, and 22.6 μs, respectively, for a simulation of 100 particles using an Intel(R) Xeon(R) Gold 6448Y (4.1 GHz) single core processor. In the linear solution, the band structure with bandwidth 5 was exploited in the Cholesky factorization. The overhead of SimHec-RC was larger than that of SimHec owing to the random force correction. Comparing EM with h = 0.25 and SimHec-RC with h = 100 (resp. h = 1000), we found a speedup of a factor of 56.6 (resp. 566).
In this study, we analyzed the relationship between the corrected Hessian matrix in a semi-implicit scheme and the random force concerning the accuracies for both the linear problem and the derived FP equations of the nonlinear problem. The proposed correction to the random force worked well for the simple MD problem involving bonded interactions. Although our approach is potentially extendable to more general potentials with bond angles and dihedral angles [3], the appropriate corrections to the Hessian matrix must be developed to avoid negative eigenvalues.
Toshiaki Hisada designed the project; Takumi Washio and Akihiro Fujii developed the semi-implicit Hessian correction scheme with random force correction, ran the simulations, analyzed the simulation data, and wrote the manuscript with input from Toshiaki Hisada; Takumi Washio developed the simulation code with input from Akihiro Fujii and Toshiaki Hisada. All authors contributed to the article and approved the submitted version.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was supported by MEXT as "Program for Promoting Researches on the Supercomputer Fugaku–8(Project ID: hp230216). We thank Edanz (https://jp.edanz.com/ac) for editing a draft of this manuscript.
The authors declare no conflict of interest.
[1] |
K. Bathe, A. Cimento, Some practical procedures for the solution of nonlinear finite element equations, Comput. Method. Appl. M., 22 (1980), 59–85. https://doi.org/10.1016/0045-7825(80)90051-1 doi: 10.1016/0045-7825(80)90051-1
![]() |
[2] |
N. Schafer, D. Negrut, A quantitative assessment of the potential of implicit integration methods for molecular dynamics simulation, J. Comput. Nonlin. Dyn., 5 (2010), 031012. https://doi.org/10.1115/1.4001392 doi: 10.1115/1.4001392
![]() |
[3] |
T. Washio, R. Kanada, X. Cui, J. Okada, S. Sugiura, S. Takada, et al., Semi-implicit time integration with Hessian eigenvalue corrections for a larger time step in molecular dynamics simulations, J. Chem. Theory Comput., 17 (2021), 5792–5804. https://doi.org/10.1021/acs.jctc.1c00398 doi: 10.1021/acs.jctc.1c00398
![]() |
[4] | P. Kloeden, E. Platen, Numerical solution of stochastic differential equations, Berlin: Springer, 1992. https://doi.org/10.1007/978-3-662-12616-5 |
[5] |
N. Grønbech-jensen, S. Doniach, Long-time overdamped Langevin dynamics of molecular chains, J. Comput. Chem., 15 (1994), 997–1012. https://doi.org/10.1002/jcc.540150908 doi: 10.1002/jcc.540150908
![]() |
[6] |
C. Sweet, P. Petrone, V. Pande, J. Izaguirre, Normal mode partitioning of Langevin dynamics for biomolecules, J. Chem. Phys., 128 (2008), 145101. https://doi.org/10.1063/1.2883966 doi: 10.1063/1.2883966
![]() |
[7] | C. Gardiner, Handbook of stochastic methods, Berlin: Springer, 1985. https://doi:10.1007/978-3-662-02377-8 |
[8] | H. Risken, The Fokker Planck equation: methods of solution and applications, Berlin: Springer, 1996. https://doi.org/10.1007/978-3-642-61544-3 |
[9] | W. Coffey, Y. Kalmykov, J. Waldron, The Langevin equation: with applications to stochastic problems in physics, chemistry, and electrical engineering, Singapore: World Scientific, 2004. https://doi.org/10.1142/9789812795090 |
[10] | K. Itô, On stochastic differential equations, New York: American Mathematical Society, 1951. https://doi.org/10.1090/memo/0004 |
[11] | CafeMolV3.2 CafeMol 3.2 manual, 2021. Available from: https://www.cafemol.org/doc/. |