12
From generalized Langevin equations to Brownian dynamics and embedded Brownian dynamics Lina Ma, Xiantao Li, and Chun Liu Citation: The Journal of Chemical Physics 145, 114102 (2016); doi: 10.1063/1.4962419 View online: http://dx.doi.org/10.1063/1.4962419 View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/145/11?ver=pdfcov Published by the AIP Publishing Articles you may be interested in BDflex: A method for efficient treatment of molecular flexibility in calculating protein-ligand binding rate constants from Brownian dynamics simulations J. Chem. Phys. 137, 135105 (2012); 10.1063/1.4756913 An immersed boundary method for Brownian dynamics simulation of polymers in complex geometries: Application to DNA flowing through a nanoslit with embedded nanopits J. Chem. Phys. 136, 014901 (2012); 10.1063/1.3672103 Inertial solvent dynamics and the analysis of spectral line shapes: Temperature-dependent absorption spectrum of β-carotene in nonpolar solvent J. Chem. Phys. 120, 4344 (2004); 10.1063/1.1644534 From the Langevin equation to the fractional Fokker–Planck equation AIP Conf. Proc. 502, 375 (2000); 10.1063/1.1302409 A simple model for Brownian motion leading to the Langevin equation Am. J. Phys. 67, 1248 (1999); 10.1119/1.19111 Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29 Oct 2016 03:27:46

From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

From generalized Langevin equations to Brownian dynamics and embeddedBrownian dynamicsLina Ma, Xiantao Li, and Chun Liu Citation: The Journal of Chemical Physics 145, 114102 (2016); doi: 10.1063/1.4962419 View online: http://dx.doi.org/10.1063/1.4962419 View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/145/11?ver=pdfcov Published by the AIP Publishing Articles you may be interested in BDflex: A method for efficient treatment of molecular flexibility in calculating protein-ligand binding rateconstants from Brownian dynamics simulations J. Chem. Phys. 137, 135105 (2012); 10.1063/1.4756913 An immersed boundary method for Brownian dynamics simulation of polymers in complex geometries:Application to DNA flowing through a nanoslit with embedded nanopits J. Chem. Phys. 136, 014901 (2012); 10.1063/1.3672103 Inertial solvent dynamics and the analysis of spectral line shapes: Temperature-dependent absorptionspectrum of β-carotene in nonpolar solvent J. Chem. Phys. 120, 4344 (2004); 10.1063/1.1644534 From the Langevin equation to the fractional Fokker–Planck equation AIP Conf. Proc. 502, 375 (2000); 10.1063/1.1302409 A simple model for Brownian motion leading to the Langevin equation Am. J. Phys. 67, 1248 (1999); 10.1119/1.19111

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 2: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

THE JOURNAL OF CHEMICAL PHYSICS 145, 114102 (2016)

From generalized Langevin equations to Brownian dynamicsand embedded Brownian dynamics

Lina Ma,a) Xiantao Li,b) and Chun Liuc)

Department of Mathematics, The Pennsylvania State University, University Park, Pennsylvania 16802-6400,USA

(Received 10 June 2016; accepted 28 August 2016; published online 16 September 2016)

We present the reduction of generalized Langevin equations to a coordinate-only stochastic model,which in its exact form involves a forcing term with memory and a general Gaussian noise. It willbe shown that a similar fluctuation-dissipation theorem still holds at this level. We study the approxi-mation by the typical Brownian dynamics as a first approximation. Our numerical test indicates howthe intrinsic frequency of the kernel function influences the accuracy of this approximation. In thecase when such an approximate is inadequate, further approximations can be derived by embeddingthe nonlocal model into an extended dynamics without memory. By imposing noises in the auxiliaryvariables, we show how the second fluctuation-dissipation theorem is still exactly satisfied. Publishedby AIP Publishing. [http://dx.doi.org/10.1063/1.4962419]

I. INTRODUCTION

The Langevin dynamics (LD) model plays a crucial rolein the stochastic modeling of bio-molecules.39 In general, theLD model can be expressed as follows:

mx = f (x) − γ x + ξ(t). (1)

Here m denotes the mass, f (x) is the inter-molecular force,and γ is the friction coefficient. In addition, ξ(t) is anexternal white noise, which satisfies the fluctuation-dissipationtheorem (FDT),24

ξ(t)ξ(t ′)T = 2kBTγδ(t − t ′). (2)

This condition ensures that the dynamical system settles tothe correct equilibrium.24

In the Langevin dynamics model (1), the dampingcoefficient and the random noise are modeling the influence ofthe solvent particles. One interesting regime is where γ ≫ 1.The asymptotic analysis has been the subject of theoreticalinterest. The analysis can be done either through the Fokker-Planck equation,41,42 using the Chapman-Enskog expansion,or directly based on the stochastic differential equation.38 InRef. 38, the LD model was written in the first order form, andby solving the second equation, the velocity variables can beeliminated. This analysis confirms that, when the damping islarge, the inertial term can be neglected, leading to

γ x = f (x) + ξ(t). (3)

This model has been referred by some authors as the Browniandynamics (BD) model.21,36,44 Notice that in this reducedmodel, ξ(t) still obeys the FDT (2).

The derivation of the Brownian dynamics is of obvioustheoretical interest. In practice, the BD (3) neglects thetransient time scale, and as a result, the dynamics occurs

a)Electronic mail: [email protected])Electronic mail: [email protected])Electronic mail: [email protected]

on a much larger time scale. Therefore, from a computationalviewpoint, the time step can be much larger than that of theoriginal Langevin dynamics (1). This model is of considerableadvantage in bio-molecular simulations, in which the timescale of interest is often out of reach when the full model (1)is used.

There have been important theoretical works toward theasymptotic analysis in the limit of large damping coefficients.The question of high friction limit also arises in thestudy of plasma gases, especially those involving nonlocalCoulomb interactions, with kinetic descriptions.14,17,18,33,35

In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflectionboundary conditions for charge distributions, which showedthe convergence to the macroscopic Poisson-Nernst-Planck(PNP) system. The purpose there was trying to justify thePNP system as a macroscopic model for the transport ofmulti-species ions in dilute solutions.

The BD model (3) can be viewed as a first orderapproximation to the full Langevin dynamics model (1).Meanwhile, there have been numerous attempts to derivehigh order approximations to improve the modeling accuracywithout having to resolve the small time scales. A remarkablesuccess is the inertial Brownian dynamics model (IBD),developed by Beard and Schlick.4,5 Using further asymptoticexpansions, the IBD incorporates correction terms into theBD model and takes into account the transient part of thedynamics.

Motivated by this effort, we studied the BD approximationof the generalized Langevin equation (GLE). In contrast tothe Langevin dynamics (1), the GLE includes a history-dependent friction term and a correlated Gaussian noise.The memory term often comes from a coarse-graining step,e.g., Refs. 2, 8–11, 15, 16, 19, 22, 25–27, 29, 30, 32, 34,40, 43, and 46, which has also been a recent emergingarea of interest in molecular modeling. Among the manyapproximation schemes6–8,12,20,23,32,34 for the GLE model, the

0021-9606/2016/145(11)/114102/11/$30.00 145, 114102-1 Published by AIP Publishing.

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 3: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-2 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

BD model (3) is clearly the simplest. To understand howthe BD approximation can naturally come about, we firstpresent a derivation which eliminates momentum variablesfrom the GLE, so that the resulting model only involves thecoordinates of the molecular variables. It will be proved thatthe reduced model still exhibits history-dependence, and thesecond fluctuation-dissipation theorem, an important propertyof the GLE, is inherited by the reduced model, just like theBD approximation of the full Langevin dynamics.

At this level, the BD approximation arises as the firstapproximation, in which the memory kernel is approximatedby a Dirac delta function. To understand the accuracy andpossibly the limitation of this approximation, we considerthe GLE derived by Adelman and Doll2,3 and conductedseveral numerical tests. Our results indicate that the accuracyhinges on the frequency associated with the memory kernel:When the frequency is large, compared to that of the meanforce, the approximation by BD is quite promising. Otherwise,the accuracy is quite limited and high order approximationsare needed. To improve the modeling accuracy, we derivetwo other approximations where the memory kernel isapproximated by rational functions in the Laplace domain.The main motivation has been twofold: On one hand, in thetime domain, the memory can be eliminated by introducingauxiliary variables. As a result, there is no need to keep thehistory of the solution and evaluate the integral associated withthe memory term at every step. This significantly suppressesthe computational cost. Meanwhile, it will be shown that withproper choices of the stochastic forces, applied to the auxiliaryvariables, the noise in the GLE is approximated by a colorednoise, in such a way that the second FDT is exactly satisfied.

II. MATHEMATICAL DERIVATION

We start with the generalized Langevin equations (GLEs),

x = v,

v = f (x) − γv − t

0θ(t − τ)v(τ)dτ + ξ(t). (4)

The mass has been set to unity. The random noise is assumedto be a mean-zero Gaussian process, with time correlationgiven by

ξ(t)ξ(t ′)T = 2kBTγδ(t − t ′) + kBTθ(t − t ′). (5)

This is a general form of the fluctuation-dissipation theorem(FDT).24 Equation of this form has been derived in ourprevious work,32 where the full model is the Langevindynamics.

We first assume that θ and γ are symmetric, matrix-valued functions. Our goal is simply solving the secondequation, and then making a direct substitution into the firstequation to eliminate v(t) entirely. For this purpose, we definea matrix-valued function χ, which satisfies the followingequation:

χ = −γ χ − t

0θ(t − τ)χ(τ)dτ, χ(0) = I . (6)

Through Laplace transform, one can show that thefunction χ is symmetric. Now χ can be viewed as thefundamental function associated with the GLE model (4). Inparticular, we have

v(t) = χ(t)v(0) + t

0χ(t − τ)ξ(τ)dτ

+

t

0χ(t − τ) f

�x(τ)�dτ. (7)

Let us first look at the first two terms, denoted by w(t),

w(t) = χ(t)v(0) + t

0χ(t − τ)ξ(τ)dτ. (8)

We assume that v(0) is sampled from its equilibrium. Namelyit is Gaussian with mean zero and covariance kBT I. Then thefollowing result can be established using the FDT (5): w(t)is a stationary Gaussian process with mean zero and timecorrelation given by

w(t)w(t ′)T = kBT χ(t − t ′), ∀t ≥ t ′. (9)

The proof is postponed to the Appendix due to some non-trivialcalculations.

With this result, we arrive at a Brownian dynamics withmemory,

x = t

0χ(t − τ) f

�x(τ)�dτ + w(t). (10)

Together with (8), this forms a closed set of equations, sincea stationary Gaussian process is uniquely determined by itstime correlation.13

To arrive at a conventional BD model (3), we mayapproximate the kernel function by

χ(t) ≈ 2χ∞δ(t), χ∞ =

+∞

0χ(t)dt, (11)

and we obtain

x(t) = χ∞ f (x(t)) + w(t). (12)

In order to preserve the FDT (9), the noise now has to beapproximated by a white noise, with covariance given by

w(t)w(t ′)T = 2kBT χ∞δ(t − t ′). (13)

This is exactly the standard Brownian dynamics model (3).The coefficients are determined from the formula (11), whichis clearly of linear-response type. Namely, we have

χ∞ =1

2kBT

+∞

0

w(t)w(0)Tdt .

In the Sec. III, we will examine the validity of thisapproximation.

III. A CASE STUDY

To test the approximation (12) presented in the Sec. II,we consider the GLE model derived by Adelman and Doll,2,3

in which a gas molecule near a solid surface was considered.We write the position of the gas molecule as y(t) and theposition of the solid atom at the surface as x(t). Starting from

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 4: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-3 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

a one-dimensional chain of solid atoms, Adelman and Doll2

used Laplace transform, together with a spatial reductionprocedure, and derived a GLE for x(t) and y(t) with theremaining solid atoms eliminated. The GLE can be written as

M y = −φ′(x − y),x = φ′(x − y) − γ x −

t

0θ(t − τ)x(τ)dτ + R(t). (14)

Here M is the mass of the gas molecule, and as in the originalwork,2 we set the mass of the solid atom to unity. We writethe memory term as a friction term with x involved. This way,the FDT is directly expressed as

R(t)R(t ′)� = kBTθ(t − t ′) + 2kBTγδ(t − t ′). (15)

We also add the damping term with coefficient γ to offer aslightly more general setup.

When the one-dimensional chain of solid atoms is infinite,an explicit formula for the kernel function can be found viaLaplace transform, and it is given by

θ(t) = ω0

tJ1(2ω0t), (16)

where J1 is the Bessel function of the first kind, and ω0 =√

Kwith K being the spring constant of the atom chain. This modelhas been considered in Refs. 28, 29, and 31 to study boundaryconditions for molecular dynamics models. Although thisexplicit formula might be already known to others, we havenot found earlier references to it. This kernel function has avery slow decay rate, which makes it a highly non-trivial testproblem. In light of the natural time scaling in (16), we willconsiderω0 as the natural frequency associated with the kernelfunction, and it represents the corresponding time scale. Tomake this model more explicit, we choose the Morse potentialfor the interaction of the gas/solid molecules at the interface,

ϕ(u) = (e−au − 1)2. (17)

a is chosen as 1, so the vibration frequency associated withthe force f (x) is ω =

√2. To further simplify the model, we

fix y = 0, so that only the second equation in (14) needs to besolved.

The Laplace transform of the kernel function is given by

Θ(s) =

s2 + 4ω20

2− s

2. (18)

This will be needed in the approximation schemes presentedhere and the ones in the Sec. IV. In particular, we obtain

χ∞ =1

γ + ω0. (19)

To examine the accuracy, we first run the full dynamicswith 8192 solid atoms and generate the trajectory of thesolid atom at the gas/solid interface. We use the stochasticvelocity-Verlet method44 to simulate the Langevin dynamicsmodel. The time correlation function

x(t), x(0) as a dynamic

property is regarded as the exact result. As comparison,we solve the BD model (12) with the effective dampingcoefficients given by (19) using the Euler-Maruyama method.We first choose γ = 2, and K = 4, which implies ω0 = 2.Figure 1 displayed the two time correlation functions. One

FIG. 1. Comparison of the time correlation functions for the full model andthe BD model (12). γ = 2,K = 4.

can see that the results agree quite well, indicating that in thisregime, the approximation by the BD model (12) is good.

In the next test, we set K = 0.2, which gives a much lowerfrequency. The results are shown in Figure 2, from which weobserve discrepancies between the results. In this case, theapproximation leads to a much bigger error.

Next, we repeat these experiments with γ = 0. In this case,the damping mechanism, e.g., χ∞, can be directly attributedto the friction kernel θ. When K = 4, it turns out that BDis still a reasonable approximation. However, when K = 0.2,the time correlation exhibits pronounced oscillations, whichis not captured by the BD model, as shown in Figure 3.

IV. EMBEDDED BROWNIAN DYNAMICS

It has been observed from the Sec. III that in the regimewhere ω0 ≪ ω, the Brownian dynamics might not be a goodapproximation to the original dynamics. In this case, we

FIG. 2. Comparison of the time correlation functions for the full model andthe BD model (12). γ = 2,K = 0.2.

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 5: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-4 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

FIG. 3. Comparison of the time correlation functions for γ = 0,K = 4 (left) and K = 0.2 (right).

propose to embed the reduced model (10) in an extendedsystem, where the memory is removed and the generalGaussian noise is generated from the additional equations.

More specifically, the approximation is done in terms ofthe Laplace transform of the kernel function, given by

X(s) = sI + γ + Θ(s)−1

. (20)

We seek a rational approximation, in the form of R(s)= Q(s)−1P(s), where P and Q are matrix polynomials, givenby

P(s) = A0sn−1 + A1sn−1 + · · · + An,

Q(s) = snI − B0sn−1 − B1sn−1 + · · · + Bn.(21)

Here n indicates the level of approximation.For the zeroth level approximation, we have a constant

function and we choose to match the limit as s → 0. We obtain

lims→0

X =γ + M∞

−1, (22)

with M∞ = lims→0Θ. This limit is defined as χ∞. Namely, wechoose R(s) = χ∞. With this constant approximation in theLaplace domain, we have a corresponding approximation inthe time domain, which gives exactly the BD model (12).

We now turn to higher-order rational approximations.For n = 1, we write R1,1(λ) = [1 − λB]−1Aλ, where λ = 1/s.Similarly we rewrite X as a function of λ,

X(λ) = I + λγ + λΘ(λ)−1

λ, (23)

and approximate this Laplace transform by R1,1.From the definition of the Laplace transform, it is clear

that X,R → 0, as λ → 0. To determine the matrices A and B,we impose the following matching conditions:

limλ→+∞

X(λ) = limλ→+∞

R1,1(λ),

limλ→0+

ddλ

X(λ) = limλ→0+

ddλ

R1,1(λ).(24)

Direct calculations yield

A = I, B = −(γ + M∞). (25)

This is equivalent to approximating χ(t) as matrix-exponentialfunction,

χ(t) ≈ χI(t) def= eBt . (26)

The rational approximation in the Laplace domain makesit possible to eliminate the memory by introducing auxiliaryvariables. We set z =

t

0 χI(t − τ) f (x(τ))dτ in (10) as thememory function, then we write an ODE for z,

z = Bz + f , z(0) = 0. (27)

We now show that by adding a white noise in thisequation, i.e., we have

x = z,z = Bz + f (x) + ξ(t), (28)

the Gaussian noise w(t) in the exact reduced model (10) canbe approximated so that the correlation function is exactlyconsistent with the approximation by χI(t).

To prove this statement, let Σ be the covariance of noise,i.e.,

⟨ξ(t)ξ(t ′)T⟩ = Σδ(t − t ′).Then by solving z analytically, we have the expression

z(t) = eBtz(0) + t

0eB(t−τ)ξ(τ)dτ +

t

0eB(t−τ) f (x(τ))dτ.

Notice the third term of z(t) is exactly the approximation ofthe memory term with the memory kernel given by χI(t).Let the covariance of z(0) be Q = kBT I. Then the first twoterms form a stationary Gaussian process, provided that thecovariance Σ satisfies the Lyapunov equation,37

− Σ = BQ +QBT . (29)

Furthermore, the time correlation of the stationary noise isgiven by eB(t−t′)Q, which is exactly χI(t), therefore the FDTas in (9) is satisfied.

It is clear that this procedure can be extended to higherorder. For example, to construct the approximation for n = 2,we set

R2,2 = [I − λB0 − λ2B1]−1[λA0 + λ2A1]. (30)

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 6: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-5 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

FIG. 4. Comparison of the kernel functions for the exact result, the first order rational approximation (26), and the second order approximation (C4). γ = 2,K = 4 (left) and K = 0.2 (right).

To determine the coefficient matrices, we impose the followingfour conditions:

limλ→+∞

X(λ) = limλ→+∞

R1,1(λ),limλ→0+

dℓ

dλℓX(λ) = lim

λ→0+

dℓ

dλℓR1,1(λ), ℓ = 1,2,3.

(31)

The explicit form of the solutions for the coefficients A0, A1,B0,and B1 can be found in the Appendix.

We let the corresponding approximate kernel functionbe χII(t), which is the inverse Laplace transform of R2,2.Similar to the previous approximation, the memory term canbe embedded into an extended system with auxiliary variablesz and z1, together with the white noises ξ1 and ξ. The extendedsystem of equations read

x = z,z1 = B1z + A1 f + ξ1,

z = z1 + B0z + A0 f + ξ.

(32)

Through Laplace transform and direct substitutions, onecan easily check that the memory function χ is approximatedby χII(t), with the inverse Laplace transform given by (C4).The random noise w(t) is also approximated by a Gaussiannoise that satisfies the FDT (9) with proper choices of the

initial conditions for z(t) and z1(t) and the covariance of thenoises ξ1 and ξ. The details can be found in Appendix C.

Now we turn to the test problem and compare the resultsfrom these models. In Figure 4, we present a comparisonbetween different orders of approximations to the exact kernelfunction for γ = 2. The left panel is for K = 4 while theright panel is for K = 0.2. We did not plot the zeroth orderapproximation (the BD model (12)) since it is a delta function.We want to comment on how we produced the exact result.While the explicit expression of θ(t) is stated in (16), the exactform of χ(t) is however nontrivial, which involves integralof highly oscillatory Bessel functions. We therefore adopta numerical Euler method to invert Laplace transform1 andname it as Exact Euler. This exact result is tested againsthigh-order numerical quadrature to ensure its accuracy.

One can observe from the graph in Figure 4 that forsmaller frequency in the memory kernel θ(t), the kernelfunction χ(t) is more flat, as shown in the graph on theright, and we have improved accuracy from the rationalapproximations R1,1. At the same time, it is obvious that withhigher order of approximation R2,2, we achieve better results.Next we repeat test for γ = 0. In this case, χ can be explicitlyexpressed as

χ(t) = J1(2ω0t)tω0

.

FIG. 5. Comparison of the kernel functions. γ = 0, K = 4 (left) and K = 0.2 (right).

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 7: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-6 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

FIG. 6. Comparison of the time correlation functions for the full model, theBD model (12), the first order rational approximation (28), and the secondorder approximation (32). γ = 2 and K = 4.

Results are shown in Figure 5, and it is clear that forthis case, much larger error is introduced in the first orderapproximation. Though the second order approximation offersbetter agreement, there is still obvious discrepancy. Thereforewe carried out a third order approximation in a similaralgorithm and clearly observed the tendency of convergenceto the exact kernel function. Similar to the situation for γ = 4,for smaller frequency in θ(t), the kernel function is moreflat.

Next, we compare the time correlation of all thecases mentioned. For γ = 2, K = 4, the results arecollected in Figure 6. In this case, all the approximations,including the BD model (12), the first order rationalapproximation (28), and the second order approximation(32), yield reasonable results. The second order methodoffers the best approximation over short time scale, this isconsistent with increased accuracy near t = 0 in the approxi-mations.

Meanwhile, for a lower frequency from the kernel, whereK = 0.2, the results are shown in Figure 7. We observethat as the order of approximation increases, the accuracy issignificantly improved. The insets in both Figures 6 and 7

FIG. 7. Comparison of the time correlation functions. γ = 2 and K = 0.2.

showed that, for higher order of approximations, derivativesnear t = 0 are better predicted.

We also repeated these experiments with γ = 0. WhenK = 4, all the approximations give reasonable predictions.One can clearly see the improvement near zero asapproximation order gets higher, since more conditionsare matched at zero. However, when K = 0.2, the largeoscillations in the time correlation are not captured by theBD model as shown in Figure 8. The first order methoddoes predict oscillations, but with the wrong magnitude. Thesecond order method provides the better approximation, andthird order is definitely the best among all.

V. SUMMARY AND DISCUSSIONS

In this paper we considered the reduction of thegeneralized Langevin equations to a Brownian dynamics withmemory (nonlocal) and correlated noise, then to the standardBrownian dynamics as well as higher order approximations.Due to the considerable advantage of the Brownian dynamicsmodel over Langevin or generalized Langevin dynamics inallowing much larger time steps and longer time scale, it isimportant to understand the accuracy of such approximations.

FIG. 8. Comparison of the time correlation functions among the approximate models for γ = 0,K = 4 (left) and K = 0.2 (right).

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 8: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-7 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

Our findings gave insight into when to use the simplest BDmodel and how to reduce the error by using high orderapproximations.

In our numerical experiments, we chose the memorykernel of the GLE model derived by Adelman and Doll.2

The corresponding memory kernel exhibits very slowdecay, which makes it a good candidate to understandthe modeling error. Our interest was in finding theappropriate regime where the Brownian dynamics is agood approximation for the GLE model. In particular,numerous experiments with various frequencies (ω0) of thememory kernel were conducted, while fixing the vibrationfrequency ω associated with the potential of mean forces.Our observations have been that the Brownian dynamics is areasonably good approximation in cases when ω0 is relativelylarge.

Meanwhile, the Brownian dynamics model will generatelarge error in the case of low frequency in the memorykernel. We derived high order approximations, by embeddingthe nonlocal Brownian dynamics model into an extended,local stochastic dynamics. We have shown that by correctlychoosing the covariance for the additive noises andthe auxiliary variables, the second fluctuation-dissipationtheorem is exactly satisfied. Our numerical results showedimproved accuracy as the order of the approximation isincreased.

At the same time, the current approach for constructinghigh order approximations can be extended/improved.For example, in constructing our models, we only usedinterpolations of the Laplace transform at λ = 0 and λ = +∞.It is possible to introduce more interpolation points to betterresolve the intermediate rescales.

ACKNOWLEDGMENTS

This research was supported by NSF under Grant Nos.DMS-1412005, DMS-1216938, and DMS-1619661.

APPENDIX A: STATIONARY GAUSSIAN PROCESS w (t )In this section, we will prove the statement made in

Section II. Namely, we will show that w(t) is indeed astationary Gaussian process satisfying the stated property.More specifically, we want to show that

w(t)w(t ′)T is a

function of (t − t ′), which is equivalent to showing

dds

w(t + s)w(t ′ + s)T = 0,

for any s. Consequently, by taking s = −t ′,w(t)w(t ′)T =

w(t − t ′)w(0)T,which shows the stationarity of the random process.

Recall that

w(t) = χ(t)v(0) + t

0χ(t − τ)ξ(τ)dτ,

and ξ(t)ξ(t ′)T = 2kBTγδ(t − t ′) + kBTθ(t − t ′),

together with the property that χ(t) is symmetric. For t ≥ t ′,the correlation can be written as

w(t + s)w(t ′ + s)T= χ(t + s)v(0)v(0)Tχ(t ′ + s)+

t+s

0

t′+s

0χ(t + s − τ)ξ(τ)ξ(τ′)T

× χ(t ′ + s − τ′)dτ′dτ= kBT χ(t + s)χ(t ′ + s)+ 2kBT

t′+s

0χ(t + s − τ′)Γ χ(t ′ + s − τ′)dτ′

+ kBT t+s

0

t′+s

0χ(t + s − τ)θ(τ − τ′)

× χ(t ′ + s − τ′)dτ′dτ.

We denote these terms as term I, II, and III, respec-tively.

Taking derivatives with respect to s to each term, we find:For Term I, we have

dds

χ(t + s)χ(t ′ + s)= χ(t + s)χ(t ′ + s) + χ(t + s) χ(t ′ + s)

=

(−χ(t + s)Γ −

t+s

0χ(τ)θ(t + s − τ)dτ

)χ(t ′ + s)

+ χ(t + s) *,−Γ χ(t ′ + s) −

t′+s

0θ(t ′ + s − τ)χ(τ)dτ+

-

= −2χ(t + s)Γ χ(t ′ + s) − t+s

0χ(τ)θ(t + s − τ)

× χ(t ′ + s)dτ − t′+s

0χ(t + s)θ(t ′ + s − τ)χ(τ)dτ,

for Term II, we have

2dds

t′+s

0χ(t + s − τ′)Γ χ(t ′ + s − τ′)dτ′

= 2χ(t + s − (t ′ + s))Γ χ(t ′ + s − (t ′ + s))

+ 2 t′+s

0χ(t + s − τ′)Γ χ(t ′ + s − τ′)dτ

+ 2 t′+s

0χ(t + s − τ′)Γ χ(t ′ + s − τ′)dτ′

= 2χ(t − t ′)Γ χ(0) − 2 t′+s

0

ddτ′

(χ(t + s − τ′)

× Γ χ(t ′ + s − τ′)) dτ′

= 2χ(t − t ′)Γ χ(0) − 2χ(t + s − τ′)Γ χ(t ′ + s − τ′)���t′+s

τ′=0

= 2χ(t + s)Γ χ(t ′ + s),

and for Term III

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 9: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-8 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

dds

t+s

0

t′+s

0χ(t + s − τ)θ(τ − τ′)

× χ(t ′ + s − τ′)dτ′dτu=t′+s−τ′==========

dds

t+s

0

t′+s

0χ(t + s − τ)θ(τ − (t ′ + s − u))

× χ(u)dudτ

v=t+s−τ=========

dds

t+s

0

t′+s

0χ(v)θ(t − t ′ − v + u)χ(u)dudv

=

t′+s

0χ(t + s)θ(t − t ′ − (t + s) + u)χ(u)du

+

t+s

0χ(v)θ(t − t ′ − v + (t ′ + s))χ(t ′ + s)dv

=

t′+s

0χ(t + s)θ(t ′ + s − τ)χ(τ)dτ

+

t+s

0χ(τ)θ(t + s − τ))χ(t ′ + s)dτ.

Notice that the last equation used the property that θ is aneven function. It is clear that the summation of all the threeterms gives zero, which shows the stationarity.

Now we have for t ≥ t ′,w(t)w(t ′)T=

w(t − t ′)w(0)T

= ⟨ *,χ(t − t ′)v(0) + t−t′

0χ(t − t ′ − τ)ξ(τ)dτ+

-v(0)T⟩

= kBT χ(t − t ′).This is precisely the stated result. �

APPENDIX B: THE DERIVATION OF THE SECONDORDER SYSTEM (32)

We approximate the Laplace transform of χ in thefollowing form:

(s2 − sB0 − B1)X = sA0 + A1,

this is equivalent as

sX1 − B1X = A1,

X1 = sX − B0X − A0.

Taking the inverse of Laplace transform, we arrive at thedifferential equations,

χ1 = B1χ, χ1(0) = A1,

χ = B0χ + χ1, χ(0) = A0.

In this calculation, we have used the matching conditions atλ = 0+ (31). Therefore, incorporating the values of X(λ) atλ = 0+ seems to be important to convert the model to the timedomain.

The solution χ will be denoted by χII. Now we want toderive a system for z =

t

0 χII(t − τ) f (x(τ))dτ. By taking the

time derivative, we get

z = χ(0) f (x(t)) + t

0χII(t − τ) f (x(τ))dτ

= A0 f (x(t)) + t

0(B0χII + χ1)(t − τ) f (x(τ))dτ

= A0 f (x(t)) + B0z + z1,

where z1 = t

0 χ1(t − τ) f (x(τ))dτ. Similarly, by takingderivative of z1, we find

z1= χ1(0) f (x(t)) + t

0χ1(t − τ) f (x(τ))dτ

= A1 f (x(t)) + t

0B1χII(t − τ) f (x(τ))dτ

= A1 f (x(t)) + B1z.

APPENDIX C: THE INITIAL COVARIANCE MATRIXFOR THE SECOND ORDER SYSTEM (32)

We will derive the initial covariance matrix for system(32). The linear form of the system can be written as

˙*,

z1

z+-= D *

,

z1

z+-+

A1 f + ξ1

A0 f + ξ

, (C1)

and we assume the initial covariance matrix is

*,

z1(0)z(0)

+-

(z1(0) z(0) )

= Q, (C2)

where

D =

0 B1

I B0

, Q =

Q1 Q12

QT12 Q2

. (C3)

We consider the case when ξ1 and ξ2 are uncorrelated. Inorder for the Lyapunov equation to hold, we need DQ to beasymmetric, that is,

−QT2 BT

1 = Q1 + B0Q12.

Using linear differential equation (C1) and solving for theanalytic solution directly, we have

*,

z1(t)z(t)

+-= eDt *

,

z1(0)z(0)

+-+

t

0eD(t−s) *

,

ξ1(s)ξ(s)

+-

ds

+

t

0eD(t−s) *

,

A1

A0

+-

f (x(s))ds.

Then we substitute it back to x = z, and the equation is in theform of

x = (0 I)eDt *,

z1(0)z(0)

+-+ (0 I)

t

0eD(t−s) *

,

ξ1(s)ξ(s)

+-

ds

+

t

0(0 I)eD(t−s) *

,

A1

A0

+-

f (x(s))ds.

In particular, the colored noise is given by

w(t) = (0 I)eDt *..,

z1(0)z(0)

+//-+ (0 I)

t

0eD(t−s) *..

,

ξ1(s)ξ(s)

+//-

ds,

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 10: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-9 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

and thus the kernel function under this approximation can be expressed in terms of a matrix exponential,

χII(t) = (0 I)eDt *,

A1

A0

+-. (C4)

Collecting terms, we have

⟨w(t)wT(t ′)⟩ = (0 I)eDtQeDT t′ *..

,

0

I

+//-+

t′

0(0 I)eD(t−s′)

ΣeDT (t′−s′) *..

,

0

I

+//-

ds′

= (0 I)eDtQeDT t′ *..

,

0

I

+//-+ (0 I)eD(t−s′)QeD

T (t′−s′) *..,

0

I

+//-

���t′

0

= (0 I)eD(t−t′)Q*..,

0

I

+//-.

To match the FDT, we need

Q1 Q12

QT12 Q2

*..,

0

I

+//-=*..,

Q12

Q2

+//-=*..,

A1

A0

+//-,

and this leads us to the conditions

Q12 = A1, Q2 = A0, Q1 = −AT0 BT

1 − B0AT1 .

With these choices, the FDT is exactly satisfied.

APPENDIX D: THE COEFFICIENTSFOR THE APPROXIMATION WITH n = 2

The moments of second order rational approximation are

[I − λB0 − λ2B1]−1[λA0 + λ2A1]= A0λ + (B0A0 + A1)λ2

+ (B0A1 + B20 A0 + B1A0)λ3 +O(λ4). (D1)

We first check the moment expansion for the kernelfunction,

χ(λ) = [I + λγ + λΘ(λ)]−1λ.

The Taylor expansion for the previous few terms is written as

χ(λ) = M0λ + M1λ2 + M2λ

3 +O(λ4).Direct calculations yield

M0 =1 − λ2Θ′(λ)

[I + λγ + λΘ(λ)]2 (0) = I,

M1 = −12

2γ + 2Θ(λ) + 4Θ′(λ)λ + (I + γ)Θ′′(λ)λ2 + (Θ′′(λ)Θ(λ) − 2Θ′(λ)2)λ3

[I + λγ + λΘ(λ)]3 (0) = −γ − Θ(0),M2 = (Θ(0) + γ)2 − Θ′(0).

Combining with the expansion for the rational function,we found

A0 = M0, B0A0 + A1 = M1,

B0M1 + B1A0 = M2, B1 = −A1(γ + M∞).

From this linear system, one easily finds that

B0 = M2(γ + M∞)(I + γ + M∞)−1. (D2)

The rest of the coefficients can be determined directly fromthese equations.

For the particular example, we take

Θ(λ) = 12*,

1λ2 + 4ω2

0 −1λ+-,

with expansion at λ = 0 given by

Kλ − K2

mλ3 +O(λ5).

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 11: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-10 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

Therefore, we obtain

Θ(0) = 0, Θ′(0) = K.

APPENDIX E: A SHORT DERIVATIONOF THE ADELMAN AND DOLL MODEL

We start with the equations for the solid atoms,mx j = K(x j+1 − 2x j + x j−1), with x j(0) = x j(0) = 0 for j ≥ 0.Taking the Laplace transform, we arrive at

X j+1 −�2 + m/Ks2�X j + X j−1 = 0, j ≥ 0. (E1)

The general solution can be written as X j = Cξ j, in which

ξ = 1 + m2K s2 − ms

2K

s2 + 4 K

m. The other root has modulus

larger than one and it has to be ruled out since it leadsto unbounded solutions. Thus we have X1 = ξX0, and inthe time domain, it is given by x1(t) =

t

0 β(t − τ)x0(τ)dτ.To completely eliminate x1, we consider the force onx0 : K(x1 − x0). To proceed, we let θ(t) = K

+∞t β(τ)dτ.

As a result, the Laplace transform is related as follows:Θ(s) = K(ξ(0) − ξ(s))/s = m

2

s2 + 4 K

m− m

2 s. In terms of thenew kernel function, the force on x0 is reduced to

K(x1 − x0) = θ(t)x0(0) − t

0θ(t − τ)x0(τ)dτ. (E2)

In the case when the initial conditions for x j, j > 0, isnonzero, an external force f ex(t) can be derived using the linearsuperposition principle. When the initial state is in thermalequilibrium, this force, together with θ(t)x0(0), gives rise tothe stationary Gaussian noise, i.e., R(t) = f ex(t) + θ(t)x0(0).This part of the calculation is less straightforward, and thereaders are referred to Ref. 29 for a derivation using theMori-Zwanzig formalism.

1J. Abate and W. Whitt, “A unified framework for numerically in-verting Laplace transforms,” INFORMS J. Comput. 18(4), 408–421(2006).

2S. A. Adelman and J. D. Doll, “Generalized Langevin equation approachfor atom/solid-surface scattering: Collinear atom/harmonic chain model,”J. Chem. Phys. 61, 4242 (1974).

3S. A. Adelman and J. D. Doll, “Generalized Langevin equation approach foratom/solid-surface scattering: General formulation for classical scatteringoff harmonic solids,” J. Chem. Phys. 64, 2375 (1976).

4D. A. Beard and T. Schlick, “Inertial stochastic dynamics. I. Long-time-step methods for Langevin dynamics,” J. Chem. Phys. 112(17), 7313–7322(2000).

5D. A. Beard and T. Schlick, “Inertial stochastic dynamics. II. Influenceof inertia on slow kinetic processes of supercoiled DNA,” J. Chem. Phys.112(17), 7323–7338 (2000).

6M. Berkowitz, J. D. Morgan, and J. Andrew McCammon, “GeneralizedLangevin dynamics simulations with arbitrary time-dependent memory ker-nels,” J. Chem. Phys. 78, 3256 (1983).

7M. Berkowitz, J. D. Morgan, D. J. Kouri, and J. A. McCammon, “Mem-ory kernels from molecular dynamics,” J. Chem. Phys. 75(5), 2462–2463(1981).

8M. Chen, X. Li, and C. Liu, “Computation of the memory functions in thegeneralized Langevin models for collective dynamics of macromolecules,”J. Chem. Phys. 141, 064112 (2014).

9A. J. Chorin and P. Stinis, “Problem reduction, renormalization, and mem-ory,” Commun. Appl. Math. Comput. Sci. 1, 1–27 (2005).

10G. Ciccotti and J.-P. Ryckaert, “On the derivation of the generalizedLangevin equation for interacting Brownian particles,” J. Stat. Phys. 26(1),73–82 (1981).

11S. Curtarolo and G. Ceder, “Dynamics of an inhomogeneouslycoarse grained multiscale system,” Phys. Rev. Lett. 88(25), 255504(2002).

12E. Darve, J. Solomon, and A. Kia, “Computing generalized Langevin equa-tions and generalized Fokker-Planck equations,” Proc. Natl. Acad. Sci.106(27), 10884–10889 (2009).

13J. L. Doob, “The elementary Gaussian processes,” Ann. Math. Stat. 15,229–282 (1944).

14N. El Ghani, N. Masmoudi et al., “Diffusion limit of the Vlasov-Poisson-Fokker-Planck system,” Commun. Math. Sci. 8(2), 463–479(2010).

15P. Espanol, “Statistical mechanics of coarse-graining,” in Novel Methods inSoft Matter Simulations (Springer, 2004), pp. 69–115.

16D. J. Evans and G. P. Morriss, Statistical Mechanics of NonequilibriumLiquids (Academic Press, 2008).

17T. Goudon, “Hydrodynamic limit for the Vlasov–Poisson–Fokker–Plancksystem: Analysis of the two-dimensional case,” Math. Models MethodsAppl. Sci 15(05), 737–752 (2005).

18T. Goudon, J. Nieto, F. Poupaud, and J. Soler, “Multidimensional high-fieldlimit of the electrostatic Vlasov–Poisson–Fokker–Planck system,” J. Differ.Equations 213(2), 418–442 (2005).

19E. Guàrdia and J. A. Padró, “Generalized langevin dynamics simulation ofinteracting particles,” J. Chem. Phys. 83(4), 1917–1920 (1985).

20C. Hijón, P. Español, E. Vanden-Eijnden, and R. Delgado-Buscalioni,“Mori–Zwanzig formalism as a practical computational tool,” FaradayDiscuss. 144, 301–322 (2010).

21G. A. Huber and J. Andrew McCammon, “Browndye: A software packagefor Brownian dynamics,” Comput. Phys. Commun. 181(11), 1896–1905(2010).

22S. Izvekov and G. A. Voth, “Modeling real dynamics in the coarse-grained representation of condensed phase systems,” J. Chem. Phys. 125,151101–151104 (2006).

23D. Kauzlaric, J. T. Meier, P. Español, S. Succi, A. Greiner, and J. G. Korvink,“Bottom-up coarse-graining of a simple graphene model: The blob picture,”J. Chem. Phys. 134(6), 064106 (2011).

24R. Kubo, “The fluctuation-dissipation theorem,” Rep. Prog. Phys. 29(1),255–284 (1966).

25H. Lei, N. Baker, and X. Li, “The generalized Langevin equation and theparameterization from data,” preprint arXiv:1606.02596 (2016).

26X. Li, “A coarse-grained molecular dynamics model for crystalline solids,”Int. J. Numer. Methods Eng. 83, 986–997 (2010).

27X. Li, “Coarse-graining molecular dynamics models using an extendedGalerkin projection,” Int. J. Numer. Methods Eng. 99, 157 (2014).

28X. Li and W. E, “Variational boundary conditions for molecular dynamicssimulations of solids at low temperature,” Commun. Comput. Phys. 1,136–176 (2006).

29X. Li and W. E, “Boundary conditions for molecular dynamics simulationsat finite temperature: Treatment of the heat bath,” Phys. Rev. B 76, 104107(2007).

30Z. Li, X. Bian, X. Li, and G. Em Karniadakis, “Incorporation of memory ef-fects in coarse-grained modeling via the Mori-Zwanzig formalism,” J. Chem.Phys. 143(24), 243128 (2015).

31W. K. Liu, E. G. Karpov, and H. S. Park, Nano Mechanics and Mate-rials: Theory, Multiscale Methods and Applications (John Wiley & Sons,2006).

32L. Ma, X. Li, and C. Liu, “Derivation and approximation of coarse-grained dynamics from Langevin dynamics,” e-print arXiv:1605.04886 [math.NA] (2016).

33J. Nieto, F. Poupaud, and J. Soler, “High-field limit for the Vlasov-Poisson-Fokker-Planck system,” Arch. Ration. Mech. Anal. 158(1), 29–59(2001).

34B. Oliva, X. Daura, E. Querol, F. X. Avilés, and O. Tapia, “A generalizedLangevin dynamics approach to model solvent dynamics effects on proteinsvia a solvent-accessible surface. The carboxypeptidase a inhibitor protein asa model,” Theor. Chem. Acc. 105(2), 101–109 (2000).

35F. Poupaud and J. Soler, “Parabolic limit and stability of theVlasov–Fokker–Planck system,” Math. Models Methods Appl. Sci10(07), 1027–1045 (2000).

36A. Ricci and G. Ciccotti, “Algorithms for Brownian dynamics,” Mol. Phys.101(12), 1927–1931 (2003).

37H. Risken, Fokker-Planck Equation (Springer, 1984).38M. San Miguel and J. M. Sancho, “A colored-noise approach to Brownian

motion in position space. Corrections to the Smoluchowski equation,” J. Stat.Phys. 22(5), 605–624 (1980).

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46

Page 12: From generalized Langevin equations to Brownian dynamics ... · In our recent work,45 we studied a Vlasov-Poisson-Fokker-Planck (VPFP) system in a bounded domain with reflection boundary

114102-11 Ma, Li, and Liu J. Chem. Phys. 145, 114102 (2016)

39T. Schlick, Molecular Modeling and Simulation: An Interdisciplinary Guide(Springer-Verlag, 2002).

40M. Stepanova, “Dynamics of essential collective motions in proteins: The-ory,” Phys. Rev. E 76, 051918 (2007).

41U. M. Titulaer, “Corrections to the Smoluchowski equation in thepresence of hydrodynamic interactions,” Physica A 100(2), 251–265(1980).

42U. M. Titulaer, “A systematic solution procedure for the Fokker-Planckequation of a Brownian particle in the high-friction case,” Physica A 91(3),321–344 (1978).

43J. C. Tully, “Dynamics of gas–surface interactions: 3d generalized Langevinmodel applied to fcc and bcc surfaces,” J. Chem. Phys. 73(4), 1975–1985(1980).

44W. F. Van Gunsteren and H. J. C. Berendsen, “Algorithms for Browniandynamics,” Mol. Phys. 45(3), 637–647 (1982).

45H. Wu, T.-C. Lin, and C. Liu, “Diffusion limit of kinetic equations for mul-tiple species charged particles,” Arch. Ration. Mech. Anal. 215(2), 419–441(2015).

46R. Zwanzig, “Nonlinear generalized Langevin equations,” J. Stat. Phys. 9,215–220 (1973).

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 132.174.254.159 On: Sat, 29

Oct 2016 03:27:46