the effect of nonlinearity in hybrid kmc-continuum models

12
The Effect of Nonlinearity in Hybrid KMC-Continuum models. Ariel Balter, Guang Lin, and Alexandre M. Tartakovsky Recently there has been interest in developing efficient ways to model heterogeneous surface reactions with hybrid computational models that couple a KMC model for a surface to a finite difference model for bulk diffusion in a continuous domain. We consider two representative problems that validate a hybrid method and also show that this method captures the combined effects of nonlinearity and stochasticity. We first validate a simple deposition/dissolution model with a linear rate showing that the KMC-continuum hybrid agrees with both a fully deterministic model and its analytical solution. We then study a deposition/dissolution model including competitive adsorption, which leads to a nonlinear rate, and show that, in this case, the KMC-continuum hybrid and fully deterministic simulations do not agree. However, we are able to identify the difference as a natural result of the stochasticity coming from the KMC surface process. Because KMC captures inherent fluctuations, we consider it to be more realistic than a purely deterministic model. Therefore, we consider the KMC-continuum hybrid to be more representative of a real system. I. INTRODUCTION Kinetic Monte Carlo (KMC) is an established stochas- tic method for simulating dynamic, spatially inhomoge- neous surface phenomena at the atomistic level. This method samples the master equation, a probabilistic de- scription of surface processes, such as reactions, depo- sition/dissolution, diffusion, etc. [5]. KMC is advanta- geous for predicting or understanding experimentally ob- served surface processes, such as morphology [10] or reac- tion rates [4], that are complex and do not permit analyt- ical solutions. One important application is in modeling catalytic reactors [7, 11, 13]. The stochastic nature of the processes involved can drive emergent behavior. It is well known that in systems containing nonlinearities, mi- croscopic fluctuations can instigate both microscopically and macroscopically observed behavior that would not be predicted in a deterministic model [8, 14, 15, 23]. One can construct mean field models (deterministic models for the average behavior of the surface), but these cannot capture the full complexity of a surface process. For non- linear systems, a purely deterministic mean field model may not even predict the ensemble average behavior. For the most part, KMC has been used to model pro- cesses on (or in) a solid interface, and only recently have there been a few attempts to couple a KMC surface model to bulk phase diffusion, for instance, a surface immersed in a solution with a reactive species, or other multi-scale models. Saedi, Drews et al., and Pricer et al. mod- eled electrodeposition of copper [6, 16, 18, 19] and Mei and Lin modeled the CO oxidation over a Ruthenium catalyst substrate [12]. KMC models of surface kinet- ics are known show different behavior than mean field models of the same surface. Therefore, it is very rea- sonable that a coupled KMC/diffusion model could show different behavior than a mean field model for the same system. In the previous works, the purpose of the KMC portion of the model is to generate realistic surface kinet- ics that supply the bulk diffusion portion of the model with a realistic surface energy [18, 19] or turnover rate [12]. Saedi did compare a coupled KMC/diffusion model to a mean field model with a deterministic surface process and demonstrated that these to models produce different results. However, he did not address the origin of this dif- ference. Our goal is to focus on the effect of fluctuations on nonlinear rate laws on a more fundamental level. Real surface processes are very complex. Adsorption, desorption, surface diffusion, and reactions all alter spa- tial distribution and even the contour of a surface. By altering the immediate neighborhood of each site these processes, in turn, alter the free energy of each site which affects future surface evolution. Furthermore, each of these processes evolves randomly. A mean field model for the surface, especially a deterministic one, can not include this complexity. Therefore, one would expect that a model that couples diffusion to a deterministic, mean-field model of surface evolution would have very different behavior than a hybrid that couples diffusion to a more realistic KMC model. However, it is not immedi- ately apparent if we can understand at a theoretical level what the differences would be and how they would arise. Towards this end, we study two representative problems, one with a linear rate law and one with a nonlinear rate law. In both problems, an adsorption/desorption pro- cess exchanges a single species between a solid phase (s) localized to a surface and a gas-phase (g) that diffuses freely in a continuous domain: s g. (I.1) Deposition and dissolution occur when a solid phase is in contact with a solution. Individual molecules may deposit onto the solid or dissolve into the solution. When this happens, the surface acquires velocity, growing or re- ceding in height. This velocity may be spatially hetero- geneous, leading to spatially heterogeneous surface mor- phology. In our simple model, we ignore the shape and velocity of the surface, making the surface an infinite source and sink – there is always plenty of room for de- position and an unlimited supply for dissolution. The rate equation for this model is linear. In catalysis, one or more reactant species adsorb to a fixed surface of a different species, the catalyst. The cat- arXiv:1108.6222v1 [cond-mat.stat-mech] 26 Aug 2011

Upload: independent

Post on 15-May-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

The Effect of Nonlinearity in Hybrid KMC-Continuum models.

Ariel Balter, Guang Lin, and Alexandre M. Tartakovsky

Recently there has been interest in developing efficient ways to model heterogeneous surfacereactions with hybrid computational models that couple a KMC model for a surface to a finitedifference model for bulk diffusion in a continuous domain. We consider two representative problemsthat validate a hybrid method and also show that this method captures the combined effects ofnonlinearity and stochasticity. We first validate a simple deposition/dissolution model with a linearrate showing that the KMC-continuum hybrid agrees with both a fully deterministic model and itsanalytical solution. We then study a deposition/dissolution model including competitive adsorption,which leads to a nonlinear rate, and show that, in this case, the KMC-continuum hybrid and fullydeterministic simulations do not agree. However, we are able to identify the difference as a naturalresult of the stochasticity coming from the KMC surface process. Because KMC captures inherentfluctuations, we consider it to be more realistic than a purely deterministic model. Therefore, weconsider the KMC-continuum hybrid to be more representative of a real system.

I. INTRODUCTION

Kinetic Monte Carlo (KMC) is an established stochas-tic method for simulating dynamic, spatially inhomoge-neous surface phenomena at the atomistic level. Thismethod samples the master equation, a probabilistic de-scription of surface processes, such as reactions, depo-sition/dissolution, diffusion, etc. [5]. KMC is advanta-geous for predicting or understanding experimentally ob-served surface processes, such as morphology [10] or reac-tion rates [4], that are complex and do not permit analyt-ical solutions. One important application is in modelingcatalytic reactors [7, 11, 13]. The stochastic nature ofthe processes involved can drive emergent behavior. It iswell known that in systems containing nonlinearities, mi-croscopic fluctuations can instigate both microscopicallyand macroscopically observed behavior that would not bepredicted in a deterministic model [8, 14, 15, 23]. Onecan construct mean field models (deterministic modelsfor the average behavior of the surface), but these cannotcapture the full complexity of a surface process. For non-linear systems, a purely deterministic mean field modelmay not even predict the ensemble average behavior.

For the most part, KMC has been used to model pro-cesses on (or in) a solid interface, and only recently havethere been a few attempts to couple a KMC surface modelto bulk phase diffusion, for instance, a surface immersedin a solution with a reactive species, or other multi-scalemodels. Saedi, Drews et al., and Pricer et al. mod-eled electrodeposition of copper [6, 16, 18, 19] and Meiand Lin modeled the CO oxidation over a Rutheniumcatalyst substrate [12]. KMC models of surface kinet-ics are known show different behavior than mean fieldmodels of the same surface. Therefore, it is very rea-sonable that a coupled KMC/diffusion model could showdifferent behavior than a mean field model for the samesystem. In the previous works, the purpose of the KMCportion of the model is to generate realistic surface kinet-ics that supply the bulk diffusion portion of the modelwith a realistic surface energy [18, 19] or turnover rate[12]. Saedi did compare a coupled KMC/diffusion model

to a mean field model with a deterministic surface processand demonstrated that these to models produce differentresults. However, he did not address the origin of this dif-ference. Our goal is to focus on the effect of fluctuationson nonlinear rate laws on a more fundamental level.

Real surface processes are very complex. Adsorption,desorption, surface diffusion, and reactions all alter spa-tial distribution and even the contour of a surface. Byaltering the immediate neighborhood of each site theseprocesses, in turn, alter the free energy of each site whichaffects future surface evolution. Furthermore, each ofthese processes evolves randomly. A mean field modelfor the surface, especially a deterministic one, can notinclude this complexity. Therefore, one would expectthat a model that couples diffusion to a deterministic,mean-field model of surface evolution would have verydifferent behavior than a hybrid that couples diffusion toa more realistic KMC model. However, it is not immedi-ately apparent if we can understand at a theoretical levelwhat the differences would be and how they would arise.Towards this end, we study two representative problems,one with a linear rate law and one with a nonlinear ratelaw. In both problems, an adsorption/desorption pro-cess exchanges a single species between a solid phase (s)localized to a surface and a gas-phase (g) that diffusesfreely in a continuous domain:

s� g. (I.1)

Deposition and dissolution occur when a solid phaseis in contact with a solution. Individual molecules maydeposit onto the solid or dissolve into the solution. Whenthis happens, the surface acquires velocity, growing or re-ceding in height. This velocity may be spatially hetero-geneous, leading to spatially heterogeneous surface mor-phology. In our simple model, we ignore the shape andvelocity of the surface, making the surface an infinitesource and sink – there is always plenty of room for de-position and an unlimited supply for dissolution. Therate equation for this model is linear.

In catalysis, one or more reactant species adsorb to afixed surface of a different species, the catalyst. The cat-

arX

iv:1

108.

6222

v1 [

cond

-mat

.sta

t-m

ech]

26

Aug

201

1

2

alyst facilitates the transformation of these species intoone or more product species, which may desorb back intothe solution. However, the catylist itself remains intact.Because a finite surface has only a finite number of bind-ing sites (in our model, s0), the surface can “fill up”,leading to competitive adsorption [9]. Under these con-ditions, the rate law has a nonlinear term.

In both of our models we solve the diffusion equationwith the same deterministic forward Euler finite differ-ence (FEFD) algorithm. However, we evolve the surfaceprocess in two different ways: (1) a deterministic FEFD,and (2) a stochastic master equation solved using thekinetic Monte Carlo (KMC) algorithm. The surface pro-cess model (KMC or mean field) supplies the boundarycondition at the surface.

We validate our numerical methods using the linearmodel (deposition/dissolution) which has an analyticalsolution for 1D diffusion with a deterministic boundarycondition corresponding to the mean field model. Wefind perfect agreement between the analytic, mean fieldnumerical and KMC hybrid models. We then comparethe numerical solutions of both the mean field model, andthe KMC hybrid model, for the system with a nonlinearsurface rate. For a wide range of parameters, we find thatthe mean field model overestimates the average surfaceoccupancy, and this result agrees with our mathematicalanalysis. In Sec. II we describe the mathematical model,a partial differential equation (PDE), that corresponds tothe physical system we consider in this paper. In Sec. IIIwe present a moment equation analysis that shows hownonlinearity in the catalysis model alters the steady-statesurface equation when stochasticity is present. In Sec. IVwe describe the numerical algorithms we use integrate thePDE described in Sec. II. Sec. V contains the graphicalresults of our simulations. We analyze and discuss thesein Secs. V and VI.

II. MATHEMATICAL MODELS

Our model system consists of a 1D continuum domainwith a reactive boundary condition at one end and afixed concentration at the other. In the continuum do-main, gas particles diffuse leading to local changes inconcentration. At the reactive boundary, particles ad-sorb and desorb according to a rate law R(s(t), g(0, t)),where s(t) is the number of particles on the surface andg(0, t) is the concentration at the surface. In the depo-sition/dissolution model, s(t) and g(y, t) are completelydecoupled, R(s(t), g(0, t)) ≡ R(g(0, t)), so we do not needto explicitly track the surface concentration. For thecatalysis model, s(t) and g(y, t) are coupled. KMC isa particle based algorithm (for evolving a particle basedmodel), while the PDE models concentration. So, ex-cept as otherwise noted, we refer to concentration in thecontinuum domain and particle number on the surface.

The system we model is a quite simplified version of amore realistic system consisting of a 2D surface bound-

ing a 3D continuum domain. Only adsorption and des-orption occur at the surface, and there are no neighborinteractions of any kind. As a result, the entire surfaceis reduced to a single 0D lattice site. By having each gasphase voxel (grid cell) span the entire surface, we reducethe 3D gas-phase domain into a 1D domain discretizedinto a line of voxels. We illustrate the conceptual modeljust described and the actual model in Fig. 1. The fi-nite number of lattice sites s0 plays a role in our catal-ysis model, but not in the deposition/dissolution model.However, the individual binding sites in Fig. 1(b) are forconceptual clarity, and purely illustrative.

We can consider two ways to formulate a PDE modelfor diffusion with a reactive boundary: (1) We could im-pose a Robin boundary condition at the surface, or (2)Include a delta function (δ(y)) source at the boundaryand a Neumann (no flux) boundary condition. From apractical point of view, it is often simpler, more accurate[17], and quite common in numerical work [2] to use thesecond approach:

∂g

∂t= D

∂2g

∂x2−R(s(t), g(y, t))δ(y) (II.1a)

−D∂g

∂x= 0 at y = 0 (II.1b)

g(Y, t) = gY (II.1c)

ds

dt= R(s(t), g(0, t)) (II.1d)

where R(s(t), g(0, t) is the rate law describing the surfaceprocess. Surface rate coefficients are specified in terms ofinverse area [17], while bulk rate coefficients are specifiedin terms of inverse volume. The δ function has units ofV −1 and accounts for this.

II.1. Linear Model

The linear model is representative of a surface growthprocess involving adsorption and desorption [20–22].Such models may use a surface reaction of the form:

R(s, g(0, t)) = kong(0, t)− koff (II.2)

The model, comprised of Eqs. (II.1) and (II.2), permitsan analytic solution:

g(y, t) = Py +Q+

N∑n=1

bn sin [λn(y − Y )] e−Dλ2nt (II.3)

with

3

y(a)System Modeled: s0only applies to catalysis

model.

y

(b)Conceptualized System: s0 only appliesto catalysis model.

FIG. 1. (a): Actual model – 0D surface and 1D continuum domain. (b): Conceptual model – 2D surface and 3D continuumdomain. s0 only applies to catalysis model.

λn = −kon

DtanλnY (II.4)

P =kongY − koffkonY +D

Q =koffY +DgYkonY +D

and

bn =4

λn

P sinλnY − PλnY +Qλn cosλnY −Qλnsin 2λnY − 2λnY

(II.5)We see that the steady-state, t→∞, solution is

gss(y) =kongY − kloffkonY +D

y +koffY +DgYkonY +D

(II.6)

Since fluctuations should not affect the ensemble meanin the linear model (see Sec. III), we use the linearmodel to validate our numerical methods. When, fora range of parameters, we see a match between the an-alytic solution, the deterministic mean field model, andthe ensemble average of many realizations of the hybridKMC/diffusion model, then we assume this validates ournumerical methods.

II.2. Nonlinear Model

In the nonlinear model,

R(s, g(0, t)) =ds

dt(II.7)

= kong(0, t) (s0 − s(t))− koffs(t)

We see that, in the nonlinear case, Eq. (II.1) is coupledto the surface because, in this case, the surface rate is ex-plicitly a function of the number of particles on the sur-face s(t) and the gas concentration at the surface g(0, t).When the number of particles on the surface reaches asteady state, i.e. ds/dt = 0, this constrains Eq. (II.1) to:

∂g

∂t= D

∂2g

∂x2(II.8a)

−D∂g

∂x= 0 at y = 0 (II.8b)

g(Y, t) = gY (II.8c)

The steady state ∂g/∂t = 0 solution of Eq. (II.8) givesg(y,∞) = gY . Therefore, using Eq. (II.7), we can derivethe steady state number of particles on the surface

sss = s0kongY

kongY + koff(II.9)

4

III. EFFECT OF FLUCTUATIONS

At the molecular scale, chemical processes are randomprocesses. Molecular scale models such as KMC includethese random fluctuations. Linear stochastic differentialor partial-differential equations have the same ensemblemean as an equivalent deterministic equation. However,nonlinear equations may not. Using the method of mo-ment equations [3], we can sometimes predict the differ-ences.

Suppose x(t) is a stochastic function of time. Themethod of moments begins with Reynolds decomposi-tion where we divide x(t) into a deterministic part anda stochastic part: Then x(t) = x(t) + x′(t), wherex(t) = 〈x(t)〉ω is the ensemble average (over realizationsω) of x(t), and x′(t) is the stochastic part. Notice thatour definition ensures x′(t) = 0. The Reynolds decom-position of Eq. (II.1) is:

∂t( g(y, t) + g′(y, t)) = D

∂2

∂y2( g(y, t) + g′(y, t))

−[R (s(t), g(y, t)) +R′ (s(t), g(y, t))

]δ(y) (III.1a)

−D ∂

∂y( g(y, t) + g′(y, t)) = 0 at y = 0 (III.1b)

g(Y, t) + g′(Y, t) = gY (III.1c)

d

dt( s(t) + s′(t)) = R (s(t), g(0, t)) +R′ (s(t), g(0, t))

(III.1d)On account of the fact that Eq. (II.1) is entirely linear,the ensemble average returns exactly Eq. (II.1).

The Reynolds Decomposition of the linear surface rateEq. (II.2) gives:

R (s(t), g(0, t)) +R′ (s(t), g(0, t)) =

kon( g(0, t) + g′(0, t))− koff (III.2)

Again, the ensemble average of Eq. (III.2) is exactly Eq.(II.2). However, for the nonlinear surface rate, Eq. (II.7),the Reynolds Decomposition gives:

R (s(t), g(0, t)) +R′ (s(t), g(0, t)) =

kon( g(0, t) + g′(0, t))(s0 − s(t)− s′(t))− koff ( s(t) + s′(t)) (III.3)

Now the ensemble average includes a non-vanishing crossmoment:

R (s(t), g(0, t)) =

kon g(0, t)(s0 − s(t))− koff s(t)− kon g′(0, t)s′(t) (III.4)

At steady state, gss(0,∞) = gY and ddt s = 0, so:

kon( s′g′)ss = kongY (s0 − sss)− koff sss (III.5)

Thus, when we consider fluctuations, we have a newsteady state value for the number of particles on the sur-face:

sss =kongY s0 − kon( s′g′)ss

kongY + koff(III.6)

which can be either greater or less than the deterministiccase depending on the sign of the cross moment ( s′g′)ss.It is challenging to derive analytical expressions for thesecross moments. However, we may approximate them nu-merically. We can only approximate them due to thefact that fluctuations are rapid and individual realiza-tions evolve in asynchronous time steps. Thus it is im-possible to calculate the ensemble statistics at an exactpoint in time. We estimate the ensemble statistics in thesame way as we do for the ensemble statistics in our re-sults, Sec. V. In Fig. 2 we have chosen parameters suchthat the cross moment s′g′ss is particularly large, lead-ing to a clearly visible effect. On the other hand, thereare many parameter ranges where s′g′ss is small, andthus, so is the effect. See Sec. V for examples.

III.1. System Characterization

Q1 =kons0Y

D(III.7a)

Q2 =koffs0Y

Dg∗(III.7b)

The parameter g∗ is a characteristic gas-phase concentra-tion. These constants represent whether the on-reactionor off-reaction is reaction-limited or diffusion-limited. Inthe mean-field model, a natural choice for g∗ is thefixed concentration gY . However, in the master-equationmodel, we work with integral numbers of particles. SincegY is both the fixed boundary concentration and the equi-librium concentration, we will also use the representation:

Q2 =koffs0Y

2

DnY(III.8)

where nY = gY L is the total number of particles in thedomain when the gas concentration has reached steady-state. This is consistent with the units represented in ourgraphs.

5

0 1 2 3 4 5 6 7 8 9 100

2

4

6

8

10

12

s(t)

KMCDeterminsticPredicted determinstic steady−state surface concentrationKMC ensenble fluctuations

0 1 2 3 4 5 6 7 8 9 100

20

40

60

80

Time

〈g′ s

′ 〉(t)

FIG. 2. In the lower panel we see that the cross moment s′ssg′ss stabilizes to a value that is > 0, and sstochss < sdetss , as predictedby Eq. (III.6). (Ensemble size: N = 200, nY = 101, S0 = 104, kon = 10−3, koff = 10, Y = 102, D = 103)

IV. ALGORITHMS

IV.1. Finite Difference continuum model for bulkdiffusion

To solve Eq. (II.1) we used a standard FEFD algo-rithm. However, in the KMC hybrid we let the KMCtake the place of Eqs. (II.2) and (II.7). We discretizethe space between y = 0 and y = Y into N voxels cen-tered at ∆y(i− 1) where i = 1...N and ∆y = Y/(N − 1).With this model, the numerical integration step for thegas-phase concentration is

gt+∆ti = gti +D

∆t

∆y2

[gti+1 − 2gti + gti− 1

]−R(st, gt0)∆tδi,0/∆y for i = 1...N − 1 (IV.1a)

−Dgt1 − gt0∆y

= 0 =⇒ gt1 = gt0 (IV.1b)

gtN = gY (IV.1c)

We integrate the number of particles on the surface using

st+∆t = st +R(st, gt0)∆t (IV.2)

in which we have discretized the Dirac delta function δ(y)using the Kronecker delta function δi,0.

In the linear case (deposition/dissolution), the rate is

R(st, gt0) = kongt0 − koff (IV.3)

In the nonlinear case (catalysis), the rate is

R(st, gt0) = kongt0(s0 − st)− koffst (IV.4)

For stability, we chose ∆t = α ×MIN{∆y2/D, 1/(kongtsurf ), 1/koff )}, where α < 1is a stability factor, usually 0.1.

IV.2. Kinetic Monte Carlo for the surface process

Because one single bulk phase voxel spans the entirehypothetical surface, molecules in the voxel at y = 0have an equal likelihood of interacting with any surfacemolecule or empty site. Likewise, because each surfacemolecules reacts only with gas molecules, and not othersurface molecules, surface molecules each have the sameprobability of interacting with any gas molecule. There-fore, we treat the “surface” as a zero-dimensional pointlocated at y = 0. Traditionally, one would use kineticMonte Carlo (KMC) to generate individual realizationsof the master equation corresponding to spatially inho-mogeneous surface process. Since we have homogenizedour surface, a small volume that encompasses the surface(st) and adjacent bulk phase gt0 constitute a ”well-mixed”system. In this respect, our KMC reduces to the simplerGillespie algorithm for a small, well-mixed volume con-taining both surface and an adjacent small volume of the

6

bulk phase. We will continue using the term KMC so asto remind the reader that we are using it for a surfaceprocess.

We employ the ”next reaction” form of this algorithm,which we will briefly review. We have on and off transi-tion rates that are re-defined at each time step accordingto:

ron roff

Linear kongt0 koff

Nonlinear kongt0 (s0 − s) koffs

Then we define rtot = ron+roff , poff = roff/rtot, andpon = ron/rtot. The algorithm has the following steps:

Repeat until t > Tmax:

1. Choose a uniform random number r1 ∈ [0, 1]

2. If poff < r1 then ∆ = −1, else ∆ = 1

3. st → st + ∆, gt1 → gt1 −∆/dy

4. Choose another uniform random number r2 ∈ [0, 1]

5. ∆t = −log(r2)/rtot

6. t = t+ ∆t

The master equation for the surface describes the evo-lution of a discrete number of particles, while the PDE,Eq. (II.1), describes the evolution of concentration. So,when a single particle is exchanged between the surfacephase and the bulk phase, we exchange an amount ∆/∆yof concentration.

IV.3. Hybrid: KMC surface model withfinite-difference bulk diffusion

The hybrid model consists of alternating the KMCsteps described above, and the diffusion steps describedin Sec. IV.1. As described in Sec. II, we treat theKMC surface process as a source for the diffusion step.The surface process (stochastic or deterministic) requiresthe concentration at the surface. Thus we can say thatthe KMC step feeds concentration to the FEFD model,and the FEFD step feeds a concentration to the KMCmodel [1]. In addition, the KMC step produces its ownrandom time step. Being random, this time step canbe much smaller or much larger than the time step re-quired for stability in the diffusion step. Therefore, weimposed a lower and higher threshold for the diffusion

time step ∆tlow = αlow∆y2

D and ∆thigh = αhigh∆y2

D ;αlow < αhigh < 1/4. For each KMC step, we perform

anywhere from none to many diffusion time steps. In thisfollowing algorithm, ∆∗t is the cumulative time since thelast diffusion step.

1. ∆∗t = 0

2. Do KMC step → ∆t

3. ∆∗t = ∆∗t+ ∆t

4. if ∆∗t < ∆tlow

(a) Do not do a diffusion step

5. elseif ∆∗t > ∆thigh

(a) M = [∆∗t/∆thigh]

(b) δt = ∆t/M

(c) do M diffusion steps with δt

(d) ∆∗t = 0

6. else

(a) do 1 diffusion step with ∆∗t

(b) ∆∗t = 0

7. go to 2.

V. RESULTS

V.1. Validation – linear model

Fig. 3 shows time dependent solutions of the linearmodel, Eqs. (II.1) and (II.2), with the initial bulk phaseconcentration zero everywhere. Notice that we presentthe concentration field in units of particles, N(x, t) =g(y, t)∆y because our simulations are in a realm of smallparticle number. This enables simpler and more intuitivecomparison of the concentration at the surface, g(0, t)with the number of particles on the surface s(t). Fig.3 shows a clear agreement between the analytical solu-tion, the deterministic FEFD solution, and the hybridKMC/FEFD solution (ensemble mean, N=100). Thisdemonstrate the accuracy of our numerical method.

V.2. Nonlinear System

For the nonlinear system, we started from initial con-ditions in which the surface coverage was 0, and thebulk phase was initialized to the steady-state concentra-tion gY . The reason we did not look at evolution fromg(x, 0) = 0 in the bulk phase (such as we did for the linearmodel) is that in the nonlinear model, the surface processrate R(s(t), g(0, t)) is 0 until the gas reaches the surface(the linear rate law does not depend on s(t)). In a KMCmodel, a rate of 0 corresponds to an infinite waiting time.To accommodate this, we would need to suspend KMCsteps until gas had diffused to the surface. Rather than

7

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

20

40

60

80

100

120

Position (y)

Par

ticle

Num

ber

(N)

t = 0

t = 0.19

t = 0.56

t = 1.3

t = 3

Standard finite differenceAnalyticHybrid KMC (µ ± σ)

(a)kon = 0.001, koff = 10−1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

20

40

60

80

100

120

Position (y)

Par

ticle

Num

ber

(N)

t = 0

t = 0.029

t = 0.17

t = 0.71

t = 3

Standard finite differenceAnalyticHybrid KMC (µ ± σ)

(b)kon = 0.01, koff = 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

20

40

60

80

100

120

Position (y)

Par

ticle

Num

ber

(N)

t = 0

t = 0.0062

t = 0.05

t = 0.39

t = 3

Standard finite differenceAnalyticHybrid KMC (µ ± σ)

(c)kon = 0.1, koff = 10

FIG. 3. Time dependent solutions of the linear model with the initial bulk phase concentration zero everywhere. KMC-continuum: solid line with ensemble fluctuations. Deterministic continuum: dashed. Analytic: dotted. Concentration field isin units of particle number, N(y) = g(y)∆y. (Ensemble size: N = 100, nY = 102, Y = 1, D = 1)

running simulations with diffusion only until g diffusedto the surface at y = 0, it seemed reasonable to choose aninitial condition that allowed the process to ensue imme-diately with a non-zero rate. A corresponding physicalmodel would be that the solution was allowed to equi-librate to a fixed concentration, and then the catalyticsurface was suddenly immersed in the solution.

Pursuing this idea, we chose parameters with somephysical relevance. We set the surface capacity to s0 =

104 particles. We might think of this as a square surface100 molecules on a side, or approximately 1nm. If wechoose the physical dimensions of ∆y to also be 1nm,then the length of a domain with Y = 100 is 0.1µm.With these dimensions, a concentration of 1/∆y3 cor-responds to 1 mol/l. Dimensional analysis gives k =D/(L2gY ), and the diffusion constant of a particle ofmolecular size is roughly ≈ 10−13m2/s, giving values ofk ≈ 10−6 to 10−2 l · mol−1 · s−1 for the range of D and

8

Y we simulated. Reaction rates a few orders of magni-tude on either side of this are very reasonable on a phys-ical basis. Therefore, while we chose parameter valuesthat enabled us to run stable simulations in a reasonableamount of time, our parameters are also within reasonfor real systems.

The time dependent solution to these models, whenlooked at over the entire spatial domain, become exces-sively busy. The concentration initially drops and theempty surface fills, and then increases back to the equilib-rium value. This caused the spatial concentration fieldsat different times to obscure each other. In our nonlin-ear, catalysis model, the ensemble average concentrationmust equilibrate to the deterministic value. However, aswe discussed, the number of particles on the surface doesnot. Therefore, for clarity and simplicity, we show onlythe values of s(t) and g(0, t).

We have shown that when the stochastic and determin-istic versions of the nonlinear model give different results,this is due to the presence of fluctuations. In our mod-els, fluctuations result from the stochastic nature of thechemical reactions. However, these stochastic reactionsare also coupled to diffusion. When fluctuations enterthe system, via the reactions, faster than they can dif-fuse away, then the system is in a reaction-limited regime.It is in this regime that we expect the system to respondmost strongly to fluctuations.

Figs 4 to 6 show results from simulations in which wefixed all parameters but one. The graphs give the en-semble average and ensemble fluctuations for ensemblesof 100 independent realizations.

To assess the effects of nonlinearity, we look at twoqualitative measures in the graphs: (1) the initial dipin gas concentration during the transient, and (2) thesteady-state surface concentration. The initial dip ingas concentration occurs because, in accordance with ourproposed physical situation (see section V.2), the cat-alytic surface is initially empty. The gas concentrationinitially dips as material adsorbs onto the surface, butthen recovers to the steady-state value. The size of thisdip, relative to the steady-state concentration is one dif-ference we see as we make the system more reaction-limited. We have discussed in section III that due tothe nonlinearity, the steady-state surface concentrationin the master-equation model and the mean-field model.The size of this effect also becomes more pronounced aswe make the system more reaction-limited.

For instance, in Fig. 5, Fig. 5(a) both Q1 and Q2 are

much larger than in Fig. 5(b). We also see a much largertransient dip. In Fig. 6 Q2 is much larger in Fig. 6(a)where we see a much greater relative difference in thesteady-state surface concentration.

VI. CONCLUSION

We have successfully validated a hybrid numericalmodel that couples KMC for a surface process to finitedifference for bulk diffusion. We have demonstrated thata KMC surface model coupled to a finite difference bulkdiffusion model can exhibit behavior not predicted by adeterministic counterpart when the surface process ratelaw is nonlinear. We first validated a numerical methodby comparing results of a linear model to an analyticsolution. Then, we showed, through stochastic analysisof moment equations, that the observed difference is nota product of the numerical method, but a fundamentalproperty of the coupled nonlinear system. We used thisanalysis to successfully explain our observed results.

This study highlights some important concepts: Theevolution of a system modeled with deterministic mean-field surface model is different than with a KMC surfacemodel, this may have two causes. This difference mayarise from the mean-field surface model being unable tocapture the complexity of the surface processes. But itmay also arise when there is nonlinear coupling betweenthe surface process and the gas phase – simply becauseof the nonlinear coupling. Even when a deterministicmean-field model for a surface process is available, if therate equations are nonlinear, then one should still use aKMC driven surface model to obtain the most realisticresults. Such a case might be when there are only weakinteractions between solid phase molecules – a physicalsituation similar to the models we studied.

VII. ACKNOWLEDGEMENTS

This work was financially supported by the LaboratoryDirected Research and Development (LDRD) project atPacific Northwest National Laboratory (PNNL) and Ap-plied Mathematics program of the US DOE Office ofAdvanced Scientific Computing Research. The PacificNorthwest National Laboratory is operated by Battellefor the U.S. Department of Energy under Contract DE-AC05-76RL01830.

[1] If we had used the Robin boundary condition formula-tion, Sec. II, the KMC would supply a flux for the finitedifference portion.

[2] Nabeel Al-Rawahi and Gretar Tryggvason. Numeri-cal simulation of dendritic solidification with convection:Two-dimensional geometry. Journal of Computational

Physics, 180(2):471 – 496, 2002.[3] Benjamin Bolker and Stephen W. Pacala. Using mo-

ment equations to understand stochastically driven spa-tial pattern formation in ecological systems. TheoreticalPopulation Biology, 52(3):179 – 197, 1997.

9

[4] A. Chatterjee, M.A. Snyder, and D.G. Vlachos. Meso-scopic modeling of chemical reactivity. Chemical Engi-neering Science, 59(22-23):5559 – 5567, 2004. ISCRE18.

[5] Abhijit Chatterjee and Dionisios Vlachos. An overview ofspatial microscopic and accelerated kinetic monte carlomethods. Journal of Computer-Aided Materials Design,14:253–308, 2007.

[6] Timothy O. Drews, Sriram Krishnan, Jay C. Alameda,Jr., Dennis Gannon, Richard D. Braatz, and Richard C.Alkire. Multiscale simulations of copper electrodeposi-tion onto a resistive substrate. IBM J. Res. Dev., 49:49–63, January 2005.

[7] Milorad P. Dudukovic. Frontiers in reactor engineering.Science, 325(5941):698–701, 2009.

[8] D J Elderfield and D D Vvedensky. Appearance ofcorrelations and symmetry breaking in non-equilibriumreaction-diffusion systems. Journal of Physics A: Math-ematical and General, 19(3):L137, 1986.

[9] Richard I. Masel. Principles of Adsorption and Reactionon Solid Surfaces. John Wiley and Sons, 1996.

[10] Paul Meakin and Kevin M. Rosso. Simple kinetic montecarlo models for dissolution pitting induced by crystal de-fects. The Journal of Chemical Physics, 129(20):204106,2008.

[11] Donghai Mei, Jincheng Du, and Matthew Neurock. First-principles-based kinetic monte carlo simulation of nitricoxide reduction over platinum nanoparticles under lean-burn conditions. Industrial & Engineering Chemistry Re-search, 49(21):10364–10373, 2010.

[12] Donghai Mei and Guang Lin. Effects of heat and masstransfer on the kinetics of co oxidation over ruo2 cata-lyst. Catalysis Today, 165(1):56 – 63, 2011. TheoreticalCatalysis for Energy Production and Utilization: fromFirst Principles Theory to Microkinetics.

[13] Matthew Neurock, Sally A. Wasileski, and Donghai Mei.From first principles to catalytic performance: trackingmolecular transformations. Chemical Engineering Sci-ence, 59(22-23):4703 – 4714, 2004. ISCRE18.

[14] G Nicolis and I Prigogine. Fluctuations in nonequilib-rium systems. Proc Natl Acad Sci U S AProceedings ofthe National Academy of Sciences of the United States ofAmerica, 68:2102–2107, 1971 Sep.

[15] Peter Ortoleva, Enrique Merino, Craig Moore, and JohnChadam. Geochemical self-organization i; reaction-transport feedbacks and modeling approach. Am J Sci,287(10):979–1007, 1987.

[16] Timothy J. Pricer, Mark J. Kushner, and Richard C.Alkire. Monte carlo simulation of the electrodeposi-tion of copper. Journal of The Electrochemical Society,149(8):C396–C405, 2002.

[17] Emily M. Ryan, Alexandre M. Tartakovsky, and CristinaAmon. Pore-scale modeling of competitive adsorption inporous media. Journal of Contaminant Hydrology, 120-121:56 – 78, 2011. Reactive Transport in the Subsurface:Mixing, Spreading and Reaction in Heterogeneous Media.

[18] Amirmehdi Saedi. A study on mutual interaction be-tween atomistic and macroscopic phenomena during elec-trochemical processes using coupled finite difference - ki-netic monte carlo model: Application to potential steptest in simple copper sulfate bath. Journal of Electroan-alytical Chemistry, 588(2):267 – 284, 2006.

[19] Amirmehdi Saedi. A study on mutual interaction be-tween atomistic and macroscopic phenomena during elec-trochemical processes using fd-kmc model: Applicationto cv test in simple copper sulfate bath. Journal of Elec-troanalytical Chemistry, 592(1):95 – 102, 2006.

[20] A. M. Tartakovsky, G. Redden, P. C. Lichtner, T. D.Scheibe, and P. Meakin. Mixing-induced precipita-tion: Experimental study and multiscale numerical anal-ysis. WATER RESOURCES RESEARCH, 44:1–19, June2008.

[21] Alexandre M. Tartakovsky, Paul Meakin, Timothy D.Scheibe, and Rogene M. Eichler West. Simulations ofreactive transport and precipitation with smoothed par-ticle hydrodynamics. Journal of Computational Physics,222(2):654 – 672, 2007.

[22] Alexandre M. Tartakovsky, Paul Meakin, Timothy D.Scheibe, and Brian D. Wood. A smoothed particle hy-drodynamics model for reactive transport and mineralprecipitation in porous and fractured porous media. Wa-ter Resour. Res., 43(5):W05437–, May 2007.

[23] N G Van Kampen. Stochastic processes in physics andchemistry. North Holland, 2007.

10

0 50 100 150 200 250 300 350 400 450 500

99.8

99.9

100

100.1

N(0,t)

0 50 100 150 200 250 300 350 400 450 5000

20

40

60

80

100

T ime

S(t)

Determinstic

KMC +/−

(a)kon = 0.0001, Q1 = 10−1, Q2 = 103

0 50 100 150 200 250 300 350 400 450 50088

90

92

94

96

98

100

N(0,t)

0 50 100 150 200 250 300 350 400 450 5000

1000

2000

3000

4000

5000

T ime

S(t)

Determinstic

KMC +/−

(b)kon = 0.01, Q1 = 10, Q2 = 103

FIG. 4. Nonlinear model with all parameters fixed except the on rate kon. (Ensemble size: N = 100, nY = 102, S0 = 104,koff = 10−1, Y = 103, D = 104)

11

0 20 40 60 80 100 12098

99

100

101

N(0,t)

0 20 40 60 80 100 1200

20

40

60

80

100

T ime

S(t)

Determinstic

KMC +/−

(a)D = 1000, Q1 = 10−2, Q2 = 102

0 100 200 300 400 500 600 700 800 900 1000

50

60

70

80

90

100

N(0,t)

0 100 200 300 400 500 600 700 800 900 10000

20

40

60

80

100

T ime

S(t)

Determinstic

KMC +/−

(b)D = 0.01, Q1 = 103, Q2 = 107

FIG. 5. Nonlinear model with all parameters fixed except the diffusion constant D. Time and memory limitations preventedus from running the simulation for Fig. 5(b) until equilibrium. (Ensemble size: N = 100, nY = 102, S0 = 104, kon = 10−5,koff = 10−1, Y = 102)

12

0 200 400 600 800 1000 1200 1400 1600

0.85

0.9

0.95

1

1.05

1.1

N(0,t)

0 200 400 600 800 1000 1200 1400 16000

2

4

6

8

10

12

T ime

S(t)

Determinstic

KMC +/−

(a)nY = 10, Q1 = 1, Q2 = 106

0 200 400 600 800 1000 1200 1400 1600

95

96

97

98

99

100

N(0,t)

0 200 400 600 800 1000 1200 1400 16000

200

400

600

800

T ime

S(t)

Determinstic

KMC +/−

(b)nY = 10000, Q1 = 1, Q2 = 104

FIG. 6. Nonlinear model with all parameters fixed except the fixed concentration nY . (Ensemble size: N = 100, S0 = 104,kon = 10−4, koff = 10−1, D = 103, Y = 103)