kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via...

53
Kinetic modeling in heterogeneous catalysis. C. Franklin Goldsmith Modern Methods in Heterogeneous Catalysis Research Friday, January 27 th , 2012

Upload: tranmien

Post on 29-Jul-2018

227 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Kinetic modeling in heterogeneous catalysis.

C. Franklin Goldsmith

Modern Methods in Heterogeneous Catalysis ResearchFriday, January 27th, 2012

Page 2: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Microkinetic modeling is a valuable tool in catalysis.

2

• Disentangle complex phenomena

• Provide insight into what’s going on at the atomic level

• Make predictions for new conditions

• Save time and money

• Optimize reactor design

Page 3: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Review: Power Law

3

Typically a differential reactor is used:

• low loading, low concentration

• isothermal

• high flow rate.

Reaction order na, nb are determined experimentally.

r = k A[ ]na B[ ]nb

Page 4: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Review: Langmuir-Hinshelwood

4

A simple, global mechanism is assumed:

• adsorbates are equilibrated with gas phase

• surface reaction is rate determining step

Apparent reaction order varies between -1 and 1:

r = kKAKB A[ ] B[ ]

1+ KA A[ ] + KB B[ ]( )2

na ≡∂ ln r∂ ln A[ ] = 1−

2KA A[ ]1+ KA A[ ] + KB B[ ]

Page 5: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Typically the experimentally observed reaction order is used to interpret a Langmuir-Hinshelwood mechanism.

5

This approach suffers two flaws:• Experimental na, nb are only valid over

a narrow range of conditions.➡ it cannot be used to predict anything

• The LH mechanism is over simplistic.➡ it cannot be used to explain anything

Page 6: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

We need a modeling approach that is predictive, not postdictive.

6

Elementary reaction mechanism:

•attempt to describe real chemistry

• is valid over a broad range of conditions

An elementary reaction occurs in a single step, i.e.

• it can be described by a single reaction coordinate

• or it passes through a single transition state

Page 7: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Catalysis is a multiscale problem.

7

10 orders of magnitude in length.15 orders of magnitude in time.14 orders of magnitude in pressure.

Chorkendorff and Niemantsverdriet. Conc. of Mod. Cat.

Page 8: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Scale is not the only challenge in catalysis.

8

naturally to hybrid simulations, whose numerical solution posesnew challenges and opportunities for multiscale mathematics.The hybrid multiscale integration algorithm (Christofides, 2001;Drews et al., 2004; Lou and Christofides, 2003; Rusli et al., 2004;Vlachos, 1997), the equation-free approach (Gear et al., 2003;Kevrekidis et al., 2004) and the heterogeneous multiscale method(E et al., 2003) are typical names and variations of couplingalgorithms.

The aforementioned coupled multiscale simulations have beenapplied to prototype systems where the emphasis was on methoddevelopment rather than on the physical systems themselves. Asa result, the models were oversimplified. In catalysis, the majorityof hybrid multiscale models with detailed mechanisms has tacitlyignored the coupling i.e., only one-way coupling was consideredwhereby the smaller scale model passed information to the next(larger) scale model. One of the most physically interesting, one-directional multiscale models for reactions in zeolites has recentlybeen reported by Hansen et al. (2010). Some examples of two-directional coupling have also been reported (Makeev et al., 2002;Raimondeau and Vlachos, 2002).

Molecular modeling, such as molecular dynamics (MD) andkinetic Monte Carlo (KMC), is often limited to short length andtime (mesoscopic) scales. Over the past 10 years, various coarse-graining and acceleration methods have been developed to enablemolecular simulation of larger systems. The coarse-grained KMCmethod reviewed herein is such an example. Accelerated MDmethods are briefly discussed below (Section 11). Similar to hybrid

multiscale modeling, most coarse-grained KMC and to some extentMD simulations have again focused on method development. Thecoupling of mesoscopic and macroscopic scales of chemical reactorsis covered in Raimondeau and Vlachos (2002) and Vlachos (1997)and will not be further discussed here. Instead our focus will be onthe micro- and meso-scales where several exciting developmentshave taken place over the past decade.

Coupling of models across scales enables top-down modelingwhereby one defines optimization targets (e.g., maximum activityand/or selectivity) and then searches for materials with suitableelectronic properties. This opens up exciting opportunities forproduct design, rather than just process design (Vlachos et al.,2006). In Section 6 we discuss how this concept is put to use forcomputation-driven catalyst design, especially of materials exhi-biting emergent behavior.

2.3. Challenges in multiscale modeling of catalytic systems

Multiscale modeling of reactions and reactors imposes multi-ple challenges, as shown in Figs. 2 and 3. The first challenge indescribing heterogeneous catalytic chemistry via first principles isthat phenomena involving chemical reactions and reactors aremultiscale in nature (Raimondeau and Vlachos, 2002; Vlachos,2005). Individual reaction events take place on specific sites of acatalyst, at the sub-nanometer length scale and over picosecond tonanosecond time scales. Yet, the macroscopic behavior is determinedby the collective behavior of reaction events (ensemble average),

Multiple scales

(b) !Ei = 41 kcal/mol (c) !Ei = 23 kcal/mol (a) !Ei = 54 kcal/mol

Intrinsic heterogeneity in

adsorbatedistribution

catalyst sites

Many-body effects

Multiple phases

Catalyst dynamics

reconstruction

isomerization

edge corner

support(111)

(100)

Missing row reconstruction on (110) plane

Fig. 3. Overview of challenges in the multiscale modeling of catalytic systems and specific examples. In the many body effects, selected configurations (top view) in thepresence of 0, 1, and 3 oxygen atoms for the 1.2 H shift reaction (ethylene isomerization of CH2CH2-CHCH3) and corresponding activation energy (kcal/mol) from periodicDFT calculations on Pt(111) indicate that the presence of co-adsorbates can strongly affect the reaction barriers. White: H atoms, gray: C atoms, red: O atoms. (Forinterpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

M. Salciccioli et al. / Chemical Engineering Science 66 (2011) 4319–4355 4321

No single modeling approach can include all effects at all scales.

Salciccioni et al. Chem. Eng. Sci. 66 (2011) 4319-4355

Page 9: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Which modeling method you chose depends upon the length scale you intend to model.

9

techniques, and specifically Density-Functional Theory (DFT), arerevolutionizing our thinking of catalytic reactions. Still, our abilityto describe, and eventually control, chemical transformations byfirst-principles modeling, at the molecular level, is hindered bymultiple challenges.

In this paper, we provide a perspective on multiscale modelingfor the development and simulation of catalytic reaction mechan-isms. First, we provide an overview of the length and time scalesin reacting systems, of the objectives of multiscale modeling, andof the challenges in first-principles modeling of chemical reac-tions and reactors. We also underscore the need for detailedreaction models by way of a few examples. The greater part of thereview then focuses on mean-field microkinetic models and theirhierarchical multiscale refinement. Emerging topics in computa-tion-driven catalyst design and uncertainty quantification are alsoreviewed. Recent developments in ab initio kinetic Monte Carlosimulations are then presented, and structure-dependent micro-kinetic models are discussed. New methods to describe catalyticchemistry in solution are outlined and an example from thehomogeneous catalytic dehydrogenation of fructose to 5-hydro-xylmethylfurfural is summarized. Finally, concluding remarks andan outlook are given.

2. Overview of multiscale modeling of chemical reactions andreactors

2.1. Scales in reacting systems

There are at least three scales encountered in a chemical reactor(Fig. 1). At the microscopic, or electronic, length and time scales(bottom of Fig. 1), adsorbate–catalyst and adsorbate–adsorbateinteractions determine the potential energy surface and thus thefree energy barrier and entropy of the chemical transformation. Acoarse description at this scale is the free energy of transformationfrom reactants to the transition state (TS) and then to products. Thethermal rate constant is a convenient way of coarse-graining theinformation from this scale, and quantum mechanical methods areideally suited, at least in principle (see below), for this task.

Given a list of reaction events and their rate constants,adsorbates arrange themselves in spatial configurations or pat-terns, as a result of the collective behavior of the ensemble of allspecies. At this mesoscopic scale (middle of Fig. 1), the collectivebehavior has to be averaged over length and time scales that aremuch larger than the characteristic length and time scale of the

underlying pattern – or what is known as the correlation length –in order to compute the reaction rate. This can be achieved vianon-equilibrium statistical mechanics techniques. Due to the fastvibrations of adsorbates with respect to the reaction time scales,adsorbates are typically thermally equilibrated, and reactionevents can be thought of as rare events, i.e., over the time scaleof a chemical reaction, the system loses its memory and can beapproximated as a Markov process. The kinetic Monte Carlo(KMC) method is the most commonly used statistical techniquefor averaging spatiotemporal events and providing the reactionrate (Bortz et al., 1975; Chatterjee and Vlachos, 2007).

At the macroscopic (reactor) scale (top of Fig. 1), there aregradients in fluid flow, concentration and temperature fields overscales that are typically much larger than the spatial inhomo-geneity of the patterns of adsorbates. As a result, the reaction ratecomputed at the mesoscopic scale can be applied over a certainlength scale (discretization size) of a chemical reactor. Due tospatial macroscopic gradients, the rate has to be evaluated at alldiscretization points of the macroscopic (reactor) domain.

At each scale, computation can be done with various methodswhose accuracy and cost vary. As one moves from left to right ofthe graph at each scale, the accuracy increases at the expense ofcomputational intensity. Thus, at each scale, one can think of ahierarchy of methods. The accuracy of these methods does notvary in a continuous fashion, i.e., each method is different. Typicalmethods are depicted in Fig. 1. Hierarchy adds a new dimensionto multiscaling: at each length and time scale, more than onemodel can be employed in the same simulation scheme, in orderto refine the results or calculate error estimates.

2.2. Objectives of multiscale modeling

The early vision of multiscale modeling was rooted in the bottom-up modeling strategy for predicting the macroscopic (reactor)behavior from microscopic scale calculations (Raimondeau andVlachos, 2002), as shown in Fig. 2. This approach naturally leads toprocess design, control, and optimization with unprecedented accu-racy. It departs significantly from the empirical process design andcontrol strategies of the past, whereby fitting to experimental datawas essential to model building.

Due to the disparity in length and time scales over which varioustools apply (Fig. 2), the straightforward, if not the only, way to reachmacroscopic scales is by coupling models describing phenomena atdifferent scales. Over the past 15 years or so, several algorithms havebeen developed to achieve this bi-directional or two-way coupling(the branches of multiscale modeling are discussed elsewhere(Vlachos, 2005). The structural difference of models across scales(continuum vs. discrete and deterministic vs. stochastic) leads

Semi-empirical: UBI-QEP, TST

ab initio:DFT, TST, DFT-

MD

Continuum:MF-ODEs

Discrete: KMC

Ideal: PFR, CSTR, etc.

ComputationalFluid Dynamics

(CFD)

Mesoscopic:PDEs

Discrete:CGMC

Pseudo-homogeneous:Transport correlations

DFT-based correlations, BEPs

Catalyst/adsorbedphase:

Reaction rate

Reactor scale:Performance

Electronic:Parameter estimation

Accuracy, cost

Fig. 1. Schematic of three scales and a possible hierarchy of models at each scale.At each scale, additional models may exist. The accuracy and cost increase fromleft to right. Acronyms from top to bottom: PRF, plug flow reactor; CSTR,continuously stirred tank reactor; ODE, ordinary differential equation; PDE, partialdifferential equation; CG-KMC, coarse-grained kinetic Monte Carlo; KMC, kineticMonte Carlo; UBI-QEP, unity bond index-quadratic exponential potential; TST,transition state theory; DFT, density functional theory; GA, group additivity; BEP,Brønsted–Evans–Polanyi; QM/MM, quantum mechanics/molecular mechanics.

Fig. 2. Schematic of various models operating at various scales. Redrawn fromVlachos (2005).

M. Salciccioli et al. / Chemical Engineering Science 66 (2011) 4319–43554320

Kinetic Monte Carlo

Mean field theory

Transition state theory

Page 10: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Outline:

10

1. Transition State Theory2. Kinetic Monte Carlo3. Mean Field Theory4. Making sense of complexity

Page 11: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

techniques, and specifically Density-Functional Theory (DFT), arerevolutionizing our thinking of catalytic reactions. Still, our abilityto describe, and eventually control, chemical transformations byfirst-principles modeling, at the molecular level, is hindered bymultiple challenges.

In this paper, we provide a perspective on multiscale modelingfor the development and simulation of catalytic reaction mechan-isms. First, we provide an overview of the length and time scalesin reacting systems, of the objectives of multiscale modeling, andof the challenges in first-principles modeling of chemical reac-tions and reactors. We also underscore the need for detailedreaction models by way of a few examples. The greater part of thereview then focuses on mean-field microkinetic models and theirhierarchical multiscale refinement. Emerging topics in computa-tion-driven catalyst design and uncertainty quantification are alsoreviewed. Recent developments in ab initio kinetic Monte Carlosimulations are then presented, and structure-dependent micro-kinetic models are discussed. New methods to describe catalyticchemistry in solution are outlined and an example from thehomogeneous catalytic dehydrogenation of fructose to 5-hydro-xylmethylfurfural is summarized. Finally, concluding remarks andan outlook are given.

2. Overview of multiscale modeling of chemical reactions andreactors

2.1. Scales in reacting systems

There are at least three scales encountered in a chemical reactor(Fig. 1). At the microscopic, or electronic, length and time scales(bottom of Fig. 1), adsorbate–catalyst and adsorbate–adsorbateinteractions determine the potential energy surface and thus thefree energy barrier and entropy of the chemical transformation. Acoarse description at this scale is the free energy of transformationfrom reactants to the transition state (TS) and then to products. Thethermal rate constant is a convenient way of coarse-graining theinformation from this scale, and quantum mechanical methods areideally suited, at least in principle (see below), for this task.

Given a list of reaction events and their rate constants,adsorbates arrange themselves in spatial configurations or pat-terns, as a result of the collective behavior of the ensemble of allspecies. At this mesoscopic scale (middle of Fig. 1), the collectivebehavior has to be averaged over length and time scales that aremuch larger than the characteristic length and time scale of the

underlying pattern – or what is known as the correlation length –in order to compute the reaction rate. This can be achieved vianon-equilibrium statistical mechanics techniques. Due to the fastvibrations of adsorbates with respect to the reaction time scales,adsorbates are typically thermally equilibrated, and reactionevents can be thought of as rare events, i.e., over the time scaleof a chemical reaction, the system loses its memory and can beapproximated as a Markov process. The kinetic Monte Carlo(KMC) method is the most commonly used statistical techniquefor averaging spatiotemporal events and providing the reactionrate (Bortz et al., 1975; Chatterjee and Vlachos, 2007).

At the macroscopic (reactor) scale (top of Fig. 1), there aregradients in fluid flow, concentration and temperature fields overscales that are typically much larger than the spatial inhomo-geneity of the patterns of adsorbates. As a result, the reaction ratecomputed at the mesoscopic scale can be applied over a certainlength scale (discretization size) of a chemical reactor. Due tospatial macroscopic gradients, the rate has to be evaluated at alldiscretization points of the macroscopic (reactor) domain.

At each scale, computation can be done with various methodswhose accuracy and cost vary. As one moves from left to right ofthe graph at each scale, the accuracy increases at the expense ofcomputational intensity. Thus, at each scale, one can think of ahierarchy of methods. The accuracy of these methods does notvary in a continuous fashion, i.e., each method is different. Typicalmethods are depicted in Fig. 1. Hierarchy adds a new dimensionto multiscaling: at each length and time scale, more than onemodel can be employed in the same simulation scheme, in orderto refine the results or calculate error estimates.

2.2. Objectives of multiscale modeling

The early vision of multiscale modeling was rooted in the bottom-up modeling strategy for predicting the macroscopic (reactor)behavior from microscopic scale calculations (Raimondeau andVlachos, 2002), as shown in Fig. 2. This approach naturally leads toprocess design, control, and optimization with unprecedented accu-racy. It departs significantly from the empirical process design andcontrol strategies of the past, whereby fitting to experimental datawas essential to model building.

Due to the disparity in length and time scales over which varioustools apply (Fig. 2), the straightforward, if not the only, way to reachmacroscopic scales is by coupling models describing phenomena atdifferent scales. Over the past 15 years or so, several algorithms havebeen developed to achieve this bi-directional or two-way coupling(the branches of multiscale modeling are discussed elsewhere(Vlachos, 2005). The structural difference of models across scales(continuum vs. discrete and deterministic vs. stochastic) leads

Semi-empirical: UBI-QEP, TST

ab initio:DFT, TST, DFT-

MD

Continuum:MF-ODEs

Discrete: KMC

Ideal: PFR, CSTR, etc.

ComputationalFluid Dynamics

(CFD)

Mesoscopic:PDEs

Discrete:CGMC

Pseudo-homogeneous:Transport correlations

DFT-based correlations, BEPs

Catalyst/adsorbedphase:

Reaction rate

Reactor scale:Performance

Electronic:Parameter estimation

Accuracy, cost

Fig. 1. Schematic of three scales and a possible hierarchy of models at each scale.At each scale, additional models may exist. The accuracy and cost increase fromleft to right. Acronyms from top to bottom: PRF, plug flow reactor; CSTR,continuously stirred tank reactor; ODE, ordinary differential equation; PDE, partialdifferential equation; CG-KMC, coarse-grained kinetic Monte Carlo; KMC, kineticMonte Carlo; UBI-QEP, unity bond index-quadratic exponential potential; TST,transition state theory; DFT, density functional theory; GA, group additivity; BEP,Brønsted–Evans–Polanyi; QM/MM, quantum mechanics/molecular mechanics.

Fig. 2. Schematic of various models operating at various scales. Redrawn fromVlachos (2005).

M. Salciccioli et al. / Chemical Engineering Science 66 (2011) 4319–43554320

11

Kinetic Monte Carlo

Mean field theory

Transition state theory

Part I: Transition-State Theory

Page 12: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Part I: Transition-State Theory

12

Transition-state theory (TST):• an approximation to dynamic theory (classical or quantum).

• evaluates the reactive flux through a dividing plane on a potential energy surface.

There are many flavors of TST.For simplicity we will focus on canonical TST.

Page 13: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

TST Assumptions

13

1. The Born-Oppenheimer approximation is valid.

2. A dynamic bottleneck between reactants and products can be identified.

3. Reactant molecules are distributed in a Maxwell-Boltzmann distribution.

Page 14: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

For n internal coordinates, the potential energy surface (PES) is an (n+1)-dimensional hypersurface.

14

AB+

CA+B

+C

A+BC

PES

dAB

dBC

Page 15: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

The reaction coordinate is the lowest energy path between reactants and products.

15

dAB

dBC

Page 16: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

The transition state is the maximum along the reaction coordinate.

16

AB+C

A...B...C

A+BC

PES

Reaction Coordinate

Page 17: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

TST assumes that the reactant(s) and transition state are equilibrated.

17

RK† TS k†⎯ →⎯ P

d P[ ]dt

= k† TS[ ] = k†K †

kTST R[ ]

Page 18: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

The rate constant is proportional to the frequency of passes over the barrier times the transition state equilibrium constant.

18

kTST T( ) = kBThk†

QTS†

QR

e−E0 /RT

K†

Note the functional similarity to the Arrhenius equation:

kArr T( ) = Ae−Ea /RT

Page 19: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

The activation energy and barrier height are correlated but not equivalent:

19

Ea ≡ −R ∂ ln r∂1 T

= RT 2 ∂∂T

ln kBTh

QTS† T( )

QR T( )⎛⎝⎜

⎞⎠⎟+ E0

Equivalence can be obtained if we assume:1. classical oscillators2. change in all other modes is negligible

A ≈kBTh

QTS†

QR

≈ νn O 1013 s-1( )

Page 20: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Transition states are categorized as loose or tight.

20

Loose transition states:1. higher in entropy than reactants2. more energy levels to be occupied3. 1013 < A < 1017 s-1

Tight transition states:1. lower in entropy than reactants2. fewer energy levels to be occupied3. 109 < A < 1013 s-1

Page 21: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Calculating kTST from first principles requires electronic structure calculations (e.g. DFT) and statistical mechanics.

21

E0 is the difference in the zero-point corrected electronic energy. • Typical DFT error is 20 - 30 kJ/mol.• Need error less than 5 kJ/mol for chemical accuracy.

The partition function requires physical properties:• vibrational frequencies• reduced moments of inertia for weakly bound rotors

Q = QelecQtransQrotQvib

Page 22: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

4000

3000

2000

1000

0

reactions

87654321number of carbon atoms

0

250

500

750

1000

species

Most systems are too largefor first-principles calculations.

We need tools to generate large-scale mechanisms rapidly but accurately.

22

Page 23: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

We can combine methods to estimate the kinetic parameters.

DFT Compute atomic binding energies

Scaling relationsUBI-QEP Estimate adsorbate binding energies

BEPUBI-QEP

Estimate activation energies

Reaction type

Estimate pre-exponential factors

23

Page 24: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Two methods are commonly used to estimate energies.

1. Linear scaling relations (Nørskov)2. Unity Bond Index - Quadratic

Exponential Potential (UBI-QEP)

24

Page 25: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

The binding energy of a molecule is linearly proportional to the binding energy of the central atom.

A ! C, xmax ! 3 for A ! N, and xmax ! 2 for A ! O; S).Since "xmax # x$ is the valency of the AHx molecule, weconclude that for the four families of molecules consideredthe slope only depends on the valency of the adsorbate. Inthe following we will consider a model that allows us tounderstand the origin of this effect.

For some of the considered systems, simple valency orbond-counting arguments [14] can explain the results:Comparing CH, CH2, and CH3 on the close-packed sur-faces, we generally find CH (with a valency of 3) to preferthreefold adsorption sites, CH2 (valency of 2) to prefertwofold adsorption, and CH3 (valency of 1) to prefer one-fold adsorption. The implication of these trends is thatunsaturated bonds on the carbon atom form bonds tosurface metal atoms; in effect, each unsaturated sp3 hybridon the central C atom binds independently to the d states ofthe nearest neighbor metal atoms, consistent with theslopes in Fig. 1. However, this picture cannot includeadsorbed atomic C. Adsorbed C also adsorbs in a threefoldsite (neglecting long range reconstructions), but it does nothave four bonds as would be needed to explain all the Cdata in Fig. 1. We also note that the overall scaling behav-ior is independent of the adsorption geometry and hencethe details of the bonding; see Fig. 2. The scaling in Figs. 1and 2 must therefore have a more general explanation that

includes the argument above for CH, CH2, and CH3 as aspecial case.

We will base our analysis on the d-band model whichhas been used quite successfully to understand trends inadsorption energies from one transition metal to the next[8,15–19]. According to the d-band model, it is useful tothink of the formation of the adsorbate-surface bond astaking place in two steps. First, we let the adsorbate statesinteract with the transition metal sp states, and then weinclude the extra contribution from the coupling to the dstates:

!E ! !Esp % !Ed: (2)

The coupling to the metal sp states usually contributes thelargest part of the bonding and involves considerable hy-bridization and charge transfer. In terms of variations fromone transition metal to the next it can, however, be consid-ered to be essentially a constant; the sp bands are broad,and all the transition metals have one sp electron per metalatom in the metallic state [20]. According to the d-bandmodel, the main contribution to the variations in bondenergy from one transition metal to the next comes fromthe coupling to the metal d states; the d states form narrowbands of states close to the Fermi level, and the width andenergy of the d bands vary substantially between transitionmetals. According to the d-band model, all the variationsamong the metals observed in Fig. 1 should therefore begiven by !Ed. That means that the x dependence of!EAHx"x$ must be given by the d coupling alone: Let usassume for the moment that the d coupling for AHx isproportional to the valency parameter ! defined above:

!EAHxd ! !"x$!EA

d (3)

Using Eq. (1), this will lead to the kind of relationship inFig. 1. We can write the adsorption energy of moleculeAHx in terms of the adsorption energy of molecule A as

-6 -5 -4 -3 -2

!EC (eV)

-2.25

-2

-1.75

-1.5

-1.25

-1

-0.75

-0.5

!EC

H3 (

eV) Fit: y=0.24x-0.02

Fit: y=0.28x-0.27

IrPt

PdNiRh

Ru

CuAu

Ag

IrRuRh

Pt

PdNi

CuAu

Ag

Cu3Pt

Pd3AuCu3Pt

Pd3Au

Ontop adsorption siteMost stable adsorption site

FIG. 2 (color). Binding energies of CH3 plotted against thebinding energies of C for adsorption in the most stable sites(triangles) and in the case where both CH3 and C have been fixedin the on-top site (squares).

FIG. 1 (color). Adsorption energies of CHx intermediates(crosses: x ! 1; circles: x ! 2; triangles: x ! 3), NHx inter-mediates (circles: x ! 1; triangles: x ! 2), OH, and SH inter-mediates plotted against adsorption energies of C, N, O, and S,respectively. The adsorption energy of molecule A is defined asthe total energy of A adsorbed in the lowest energy positionoutside the surface minus the sum of the total energies of A invacuum and the clean surface. The data points represent resultsfor close-packed (black) and stepped (red) surfaces on varioustransition-metal surfaces. In addition, data points for metals inthe fcc(100) structure (blue) have been included for OHx.

PRL 99, 016105 (2007) P H Y S I C A L R E V I E W L E T T E R S week ending6 JULY 2007

016105-2

25

ΔEAHx = γΔEA + ξ

γ =xmax − xxmax

xmax = 4 for A = C= 3 for A = N= 2 for A = O, S

Abild-Pederson et al. Phys. Rev. Let. 99 (2007)

Page 26: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Brønsted-Evans-Polanyi (BEP) can be used to estimate the activation energy.

26

adsorbates and, second, the use of a BEP relation to predict theenergetics of the transition statesallows us to map out the entirePES for the overall reaction on several metal surfaces. This mappingis done in Figure 3 for 10 relevant catalytic close-packed metalsurfaces, where the energetics for C-C and C-O bond breakingin ethanol for all 24 possible C-C-O backbone intermediates oneach of the 10 metals are shown. We note that this PES can bederived for any metal surface, provided that we know only thebinding energy of CO, C, O, and H on that surface.

In principle, one could develop a full microkinetic model50 basedon this overall PES for each surface and use this model to predictthe overall rate of this reaction, but one can see that even for thisrelatively simple system, the number of possible pathways is large.Additionally, this information is incomplete, as only direct C-Oand C-C bond cleavages are given by the BEP relation consideredhere; for instance, disproportionation steps are not included in thisanalysis. In addition, C-H and O-H bond cleavage is notconsidered explicitly. While the use of the correlations as outlinedabove has vastly simplified the generation of large amounts of data,the challenge of extracting reaction rates from this data remains.As a first approximation, we have developed a simple mean-fieldkinetic model, as described below, to estimate reaction rates atsteady state.

Surface Intermediates. Energetics for species included in thekinetic model were calculated as described above using the scalingcorrelation28 and the DFT-calculated binding energies of H, C, O,and CO. The energy of the transition state is calculated from theBEP correlations developed on Ru(0001), as shown in Figure 2.The vibrational frequencies for all surface intermediates arecalculated on Ru(0001);39 these values are assumed to be inde-pendent of metal. Using these vibrational frequencies, we calculatethe entropy and zero-point energies of surface species. Sincevibrational frequencies are assumed to be independent of metal,both entropy and zero-point energies for surface species are alsoassumed to be independent of metal. Additionally, these twoquantities are assumed to be invariant with surface coverage.Entropy values for gas-phase intermediates are taken from publishedtables.51

The model assumes the presence of a most abundant surfaceintermediate (MASI). To determine the nature of the MASI, all

intermediates are first assumed to be in equilibrium with theappropriate gas-phase species and the stoichiometric amount of gas-phase hydrogen: CHxCHyOHz species are assumed to be inequilibrium with ethanol, CHxCHy species with ethane, CHxOHy

species with carbon monoxide, CHx species with methane, and OHx

species with water. The coverage of each species is calculated onthe basis of this equilibrium, and the highest-coverage species isassigned as the MASI. The gas-phase pressure of each species isset equal to the value determined by the respective experimentalmeasurement.

Our model includes the effect of surface coverage on the bindingenergy of reactive intermediates. In particular, we find that, underour experimental conditions, CO is the most abundant surfaceintermediate for all metals studied experimentally, except for Cu.For this reason, CO binding is calculated separately, rather than asa function of C binding on metal surfaces. As a first approximationfor adsorbate binding energy as a function of CO surface coverage,we use a correlation developed earlier for Pt(111):10

where BECO (eV) is the differential binding energy of CO (i.e., thebinding energy of the last CO added to the surface), evaluated atthe CO coverage of !CO. This correlation reflects the repulsiveinteraction between CO molecules on Pt(111) and accuratelypredicts the saturation coverage of CO on that surface. To a firstapproximation, we assume that the CO binding energy dependenceon !CO on other transition metals is identical to that on Pt(111). Inthat case, the constant term in the above relation changes from metalto metal, depending on the initial binding energy of CO on thatmetal, but the second and third terms are kept invariant across themetals studied. Using this function, we can solve for the coverageof CO, which is the MASI for most metals studied. For Cu, wherethe majority of the surface is comprised of vacant sites, no coverageeffects have been included in our analysis.

As the coverage of CO will also destabilize other surface species,we have included this effect in the model. To a first approximation,

(50) Cortright, R. D.; Dumesic, J. A. AdV. Catal. 2002, 46, 161.

(51) Afeefy, H.; Liebman, J. F.; Stein, S. E. NIST Chemistry Webbook,NIST Standard Reference Database 69; Linstrom, P. J., Mallard, W. G.,Eds.; National Institute of Standards and Technology: Gaithersburg,MD.

Figure 2. BEP correlation for C-C and C-O bond-breaking steps on Ru (0001)39 (in black) and Pt(111)30 (in red). The best fit line for Ru(0001) is ETS

(eV) ) 0.88EFS + 1.07 (R2 ) 0.98), where EFS is the energy of the final state for the reaction written in the exothermic direction, relative to the initial stategas-phase species and the clean slab. ETS is the energy of the transition state with the same reference.

BECO(!CO) ) -1.78 + 0.0065e4.79!CO + 0.031135!COe4.79!CO

5812 J. AM. CHEM. SOC. 9 VOL. 131, NO. 16, 2009

A R T I C L E S Ferrin et al.

Ferrin et al. JACS. 131, (2009)

α, β can be refined based upon type of bond broken, surface, etc.

Ea = αΔE + β

Page 27: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Linear scaling relations reduce a complex mechanism down to a minimum set of descriptors.

27 Grabow et al. Angew. Chem. 50, (2011)

energy also scales with the C adsorptionenergy, and the same is roughly true of theH adsorption energy (although here thevariation from one metal to the next isvery slight). The result is that in a !rstapproximation the activity of a given metalis given by only two descriptors: the C andthe O adsorption energies.Fig. 7 shows the calculated rate of

methanation under industrial conditions asa function of these two parameters. Theinput into the model is the set of scaling

relationships for reaction energies andtransition energies. The result is a volcano-shaped relationship between the catalyticrate and the adsorption energies. Themean absolute errors introduced by thescaling relations are on the order of 0.2–0.3eV, errors that are within the limit of DFT.These errors are very small comparedwith the energy scale on which the volcanois de!ned. Hence, trends will be well-described even based on approximatescaling relations. The volcano relationshipis a very fundamental concept in hetero-geneous catalysis (49, 50) because it allowsthe identi!cation of the best catalytic ma-terials for a given reaction as a function ofthe chosen descriptors (51). A search forgood catalysts therefore amounts to !ndingmaterials with descriptors that are close tothe maximum of the volcano.Themost critical issue is to determine the

active site for a speci!c reaction. For themethanation reaction, it is the step or kinksites on the metal particles that providethe proper structure for the breaking of thestrong CO bond. Each structure de!nesa new set of scaling and Brønsted–Evans–Polanyi relations. Hence, identifying whichsites are present during the reaction andidentifying which of the resulting relationslead to the highest rate are the essenceof determining the proper catalytic mate-rial for the process.For the methanation reactions, the ori-

gin of the volcano is simple. We !nd thatthe breaking of the CO bond is the rate-

determining step for relatively unreactivecatalysts such as Ni and that this reactionstep proceeds via bond breaking of the C-OH intermediate yielding adsorbed C andOH on the surface (52–55). On the morereactive metals, such as Fe, this reactionstep becomes fast, and the surface be-comes poisoned by adsorbed carbon andoxygen. A good methanation catalysttherefore represents a reasonable com-promise between a low CO dissociationbarrier and a high carbon and oxygen de-sorption rate (in terms of the CH4 andH2O desorption, respectively) (56). As canbe seen in the volcano relation in Fig. 7,the theoretical results describe experi-mental !ndings (49) very well: Ru andCo lie at the top of the volcano, whereasRh and Ni on the right and Fe on theleft are calculated to have somewhatlower activities.Industrially, the catalyst that is com-

monly used is based on Ni. Even though Ruand Co have been found to be the mosteffective (39), Ni is still preferred due toits lower price. Recently, DFT calculationsidenti!ed Ni-Fe catalysts as a cheaper al-ternative with a higher activity than Ni(see also Fig. 7). These !ndings have beenveri!ed experimentally (56).

Descriptor-Based Search for NewCatalystsThe identi!cation of Ni-Fe as a possiblecatalyst for methanation is a very simple,early example of descriptor-based searchesfor new heterogeneous catalysts. There areseveral others in the recent literature (7,25, 57–62). In the following, we will illus-trate the approach with another example,which is a little more complex because itinvolves not only the rate but also the se-lectivity of the process.We will take the selective hydrogenation

of acetylene in the presence of an excess ofethylene as an example. The reaction isimportant industrially, because it is used toremove trace amounts of acetylene fromthe ethylene used to make polymers. Fig. 8shows the calculated pathway for the hy-drogenation of acetylene to ethylene andthe further hydrogenation of ethylene toethane on the Pd(111) surface. The pro-cess starts with the adsorption and suc-cessive hydrogenation of acetylene toethylene. Once ethylene is formed, it candesorb or react further to produce ethane,which is unwanted in this process (if thiswere fast, the ethylene would all disap-pear). A selective catalyst should have anactivation barrier for the hydrogenation ofethylene that is higher than its desorptionenergy to favor desorption of the productrather than its further hydrogenation.Whereas these two energies are compa-rable for Pd(111), the desorption barrier issmaller than the activation barrier for thePdAg(111) surface (Fig. 8), making this

Meta-data

Catalytic properties

(Selectivity, turnover, ...)

Data(Energies, ...)

Experiment

Models

Descriptors

Quantumcalculations

Fig. 6. Illustration of the link between the mi-croscopic surface properties and the macroscopiccatalytic properties as measured in a catalyticconverter. They are directly linked through thekinetics, but it is far more instructive to developmodels that map the large number of microscopicproperties characterizing the catalyst onto a fewdescriptors.

Fig. 7. Theoretical volcano for the production of methane from syngas, CO, and H2. The turnoverfrequency (TOF) is plotted as a function of carbon and oxygen binding energies. The carbon and oxygenbinding energies for the stepped 211 surfaces of selected transition metals are depicted. Reactionconditions are 573 K, 40 bar H2, 40 bar CO.

940 | www.pnas.org/cgi/doi/10.1073/pnas.1006652108 Nørskov et al.

reactive Rh and Pt surfaces are covered with almost onemonolayer (1 ML) of carbon, which reacts with NHy frag-ments to give HCN. Multiple carbon layers deactivate thecatalyst, which provides evidence for a metal-catalyzedreaction where the C–N coupling may occur between differ-

ent CHx and NHy species under industrial conditions.[32,33]

Evidence for a direct coupling of atomic C and N on thePt(111) surface from NH3 and CH4 at approximately 500 K isprovided by ultrahigh-vacuum studies using temperature-programmed desorption (TPD), X-ray photoelectron spec-troscopy (XPS), and reflection–absorption IR spectroscopy(RAIRS) on Pt(111).[35,36] Based on experiments in a tempo-ral analysis of products (TAP) reactor on Pt black at 1173 K,Delagrange and Schuurman[37] suggested an HCN synthesisreaction mechanism in which NH3 decomposition is the rate-limiting step and HCN is formed by a C–N or HC–N couplingreaction. However, the TAP experiments do not allow theexclusion of other CHx–NHy coupling mechanisms. A verydetailed DFT study on Pt(111) suggests that the C–N couplingstep in the Degussa process occurs between C or CH andNH2.

[38] Lastly, early coupling reactions of CHx with NHy onthe catalyst surface and subsequent dehydrogenation as ahomogeneous gas-phase reaction have also been dis-cussed.[39,40]

To answer the mechanistic question we inspected thesurface coverages and rates of the various C–N couplingreactions. Figure 3 shows that the most abundant surfaceintermediate for all the investigated metals is carbon. On thevery strong binding metals W, Mo, and Fe, the carbon

coverage is 1 ML, whereas for all other metals qC< 0.95 ML,which corresponds to the coverage on Pt. This result is in goodagreement with the TPD and atomic-emission spectrometry(AES) measurements by Hasenberg and Schmidt.[32,33] Theonly other surface species that can be observed in notablequantities is N. High N coverages are predicted for very strongEN and weak EC but none of the metals studied herein fallsinto this region. High surface coverages can alter thestabilities of surface intermediates and transition states andusually lead to a destabilization through repulsive interac-tions. As has been shown for CO oxidation, one can estimatethe effect of adsorbate–adsorbate interactions to a firstapproximation by moving the metal points to weaker bindingenergies.[41] Across different transition-metal surfaces theseinteractions have a similar effect on the stability of surfaceintermediates and change the quantitative results, but theobserved trends are expected to remain identical.

The individual rates of each possible CHx and NHy speciesare presented as a grid in Figure 4. The panels from left toright in the first row show the coupling rates of C with N, NH,and NH2, respectively. Similarly, the first column from top tobottom shows the coupling rates of N with C, CH, and CH2.Although it appears as if the direct C–N coupling mechanismis most active in a large region of the investigated parameterspace, we note that for Pt the C–NH2 coupling mechanism isdominant and C–N coupling can be neglected. In contrast, the

Figure 2. Calculated logarithmic TOF for a) HCN production and b) N2

production as function of the carbon (EC) and nitrogen (EN) bindingenergies. The error bars indicate an uncertainty of 0.2 eV for EC andEN.

Figure 3. Calculated coverages of C (top) and N (bottom) as afunction of the carbon (EC) and nitrogen (EN) binding energies. Errorbars and metal labels except Pt and Rh have been omitted for clarity.

4603Angew. Chem. Int. Ed. 2011, 50, 4601 –4605 ! 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

Nørskov et al. PNAS. 108, 3, (2011)

CO + 3H2 → CH4 + H2O CH4 + NH3 → HCN + 3H2

Page 28: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

UBI-QEP is a semi-empirical method for coverage-dependent activation energies.

28

Requires atomic binding energies and gas-phase molecular dissociation energies.

1. Two-body interactions are described by a quadratic potential, exponential in distance (Morse).

2. Total energy of many-body system is the sum of two-body interactions.

Fast, easy, popular -- but can lead to big errors for larger molecules.

Maestri and Reuter. Angew. Chem. 50 (2011) Shustorovich and Sellers. Surf. Sci. Rep. 31 (1998)

Page 29: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

A kinetic mechanism must conserve enthalpy and entropy.

reactants products

ΔHrxn and ΔSrxn must be the same for either path.

29

Page 30: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

ΔH rxn = Ea, f − Ea,r

ΔSrxn = R lnAf

Ar

Thermodynamic consistency constrains the Arrhenius parameters:

30

Most mechanisms are not consistent! Typically entropy is not conserved;equilibrium constants are off by orders of magnitude.

Page 31: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

There are two approaches to enforcing thermodynamic consistency:

1. Constrain all Ea,rev and Arev according to a basis set*.

• Requires accurate kadsorption, kdesorption for all intermediates

2. Compute kr directly from the equilibrium constant.

• Requires accurate H(T), S(T) for all intermediates

31 *Mhadeshwar, Wang, & Vlachos. J. Phys. Chem. B., 107, 2003

Page 32: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

techniques, and specifically Density-Functional Theory (DFT), arerevolutionizing our thinking of catalytic reactions. Still, our abilityto describe, and eventually control, chemical transformations byfirst-principles modeling, at the molecular level, is hindered bymultiple challenges.

In this paper, we provide a perspective on multiscale modelingfor the development and simulation of catalytic reaction mechan-isms. First, we provide an overview of the length and time scalesin reacting systems, of the objectives of multiscale modeling, andof the challenges in first-principles modeling of chemical reac-tions and reactors. We also underscore the need for detailedreaction models by way of a few examples. The greater part of thereview then focuses on mean-field microkinetic models and theirhierarchical multiscale refinement. Emerging topics in computa-tion-driven catalyst design and uncertainty quantification are alsoreviewed. Recent developments in ab initio kinetic Monte Carlosimulations are then presented, and structure-dependent micro-kinetic models are discussed. New methods to describe catalyticchemistry in solution are outlined and an example from thehomogeneous catalytic dehydrogenation of fructose to 5-hydro-xylmethylfurfural is summarized. Finally, concluding remarks andan outlook are given.

2. Overview of multiscale modeling of chemical reactions andreactors

2.1. Scales in reacting systems

There are at least three scales encountered in a chemical reactor(Fig. 1). At the microscopic, or electronic, length and time scales(bottom of Fig. 1), adsorbate–catalyst and adsorbate–adsorbateinteractions determine the potential energy surface and thus thefree energy barrier and entropy of the chemical transformation. Acoarse description at this scale is the free energy of transformationfrom reactants to the transition state (TS) and then to products. Thethermal rate constant is a convenient way of coarse-graining theinformation from this scale, and quantum mechanical methods areideally suited, at least in principle (see below), for this task.

Given a list of reaction events and their rate constants,adsorbates arrange themselves in spatial configurations or pat-terns, as a result of the collective behavior of the ensemble of allspecies. At this mesoscopic scale (middle of Fig. 1), the collectivebehavior has to be averaged over length and time scales that aremuch larger than the characteristic length and time scale of the

underlying pattern – or what is known as the correlation length –in order to compute the reaction rate. This can be achieved vianon-equilibrium statistical mechanics techniques. Due to the fastvibrations of adsorbates with respect to the reaction time scales,adsorbates are typically thermally equilibrated, and reactionevents can be thought of as rare events, i.e., over the time scaleof a chemical reaction, the system loses its memory and can beapproximated as a Markov process. The kinetic Monte Carlo(KMC) method is the most commonly used statistical techniquefor averaging spatiotemporal events and providing the reactionrate (Bortz et al., 1975; Chatterjee and Vlachos, 2007).

At the macroscopic (reactor) scale (top of Fig. 1), there aregradients in fluid flow, concentration and temperature fields overscales that are typically much larger than the spatial inhomo-geneity of the patterns of adsorbates. As a result, the reaction ratecomputed at the mesoscopic scale can be applied over a certainlength scale (discretization size) of a chemical reactor. Due tospatial macroscopic gradients, the rate has to be evaluated at alldiscretization points of the macroscopic (reactor) domain.

At each scale, computation can be done with various methodswhose accuracy and cost vary. As one moves from left to right ofthe graph at each scale, the accuracy increases at the expense ofcomputational intensity. Thus, at each scale, one can think of ahierarchy of methods. The accuracy of these methods does notvary in a continuous fashion, i.e., each method is different. Typicalmethods are depicted in Fig. 1. Hierarchy adds a new dimensionto multiscaling: at each length and time scale, more than onemodel can be employed in the same simulation scheme, in orderto refine the results or calculate error estimates.

2.2. Objectives of multiscale modeling

The early vision of multiscale modeling was rooted in the bottom-up modeling strategy for predicting the macroscopic (reactor)behavior from microscopic scale calculations (Raimondeau andVlachos, 2002), as shown in Fig. 2. This approach naturally leads toprocess design, control, and optimization with unprecedented accu-racy. It departs significantly from the empirical process design andcontrol strategies of the past, whereby fitting to experimental datawas essential to model building.

Due to the disparity in length and time scales over which varioustools apply (Fig. 2), the straightforward, if not the only, way to reachmacroscopic scales is by coupling models describing phenomena atdifferent scales. Over the past 15 years or so, several algorithms havebeen developed to achieve this bi-directional or two-way coupling(the branches of multiscale modeling are discussed elsewhere(Vlachos, 2005). The structural difference of models across scales(continuum vs. discrete and deterministic vs. stochastic) leads

Semi-empirical: UBI-QEP, TST

ab initio:DFT, TST, DFT-

MD

Continuum:MF-ODEs

Discrete: KMC

Ideal: PFR, CSTR, etc.

ComputationalFluid Dynamics

(CFD)

Mesoscopic:PDEs

Discrete:CGMC

Pseudo-homogeneous:Transport correlations

DFT-based correlations, BEPs

Catalyst/adsorbedphase:

Reaction rate

Reactor scale:Performance

Electronic:Parameter estimation

Accuracy, cost

Fig. 1. Schematic of three scales and a possible hierarchy of models at each scale.At each scale, additional models may exist. The accuracy and cost increase fromleft to right. Acronyms from top to bottom: PRF, plug flow reactor; CSTR,continuously stirred tank reactor; ODE, ordinary differential equation; PDE, partialdifferential equation; CG-KMC, coarse-grained kinetic Monte Carlo; KMC, kineticMonte Carlo; UBI-QEP, unity bond index-quadratic exponential potential; TST,transition state theory; DFT, density functional theory; GA, group additivity; BEP,Brønsted–Evans–Polanyi; QM/MM, quantum mechanics/molecular mechanics.

Fig. 2. Schematic of various models operating at various scales. Redrawn fromVlachos (2005).

M. Salciccioli et al. / Chemical Engineering Science 66 (2011) 4319–43554320

32

Kinetic Monte Carlo

Mean field theory

Transition state theory

Part II: Kinetic Monte Carlo

Page 33: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Part II: Kinetic Monte Carlo

33

KMC is a method for simulating state-to-state dynamics of a rare event system.

• Can span a large range of time scales by neglecting unimportant ultrafast phenomena.

• Explicitly accounts for spatial heterogeneity in competing processes:

- adsorption/desorption

- surface diffusion

- surface reactions

Page 34: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Coarse-grain time scales allows us to model chemical reactions.

34

lution at the atomic level are molecular dynamics (MD)simulations, which correspond to a numerical integrationof Newton’s equations of motion.14 Starting in any oneof the two minima and using the forces provided by thePES gradient, a MD trajectory would therefore explicitlytrack the entire thermal motion of the adsorbate. In or-der to accurately resolve the picosecond-scale vibrationsaround the PES minimum this requires time steps in thefemtosecond range. Surmounting the high barrier to getfrom one basin to the other is only possible, if enoughof the random thermal energy stored in all other degreesof freedom gets coincidentally united in just the right di-rection. If this happens as in our example only everymicrosecond or so, a MD simulation would have to firstcalculate of the order of 109 time steps until one of thereally relevant di!usion events can be observed. Evenif the computational cost to obtain the forces would benegligible (which as we will see below is certainly notthe case for PESs coming from first-principles calcula-tions), this is clearly not an e"cient tool to study thelong-term time evolution of such a rare-event system. Infact, such long MD simulations are presently computa-tionally not feasible for any but the most simple modelsystems, and spending CPU time into any shorter MDtrajectory would only yield insight into the vibrationalproperties of the system.

B. State-to-state dynamics and kMC trajectories

The limitations to a direct MD simulation of rare-eventsystems are therefore the long time spans, in which thesystem dwells in one of the PES basins before it escapesto another one. Precisely this feature can, however, alsobe seen as a virtue that enables a very e"cient accessto the system dynamics. Just because of the long timespent in one basin, it should not be a bad assumptionthat the system has forgotten how it actually got in therebefore it undergoes the next rare transition. In this case,each possible escape to another PES basin is then com-pletely independent of the preceding basins visited, i.e.of the entire history of the system. With respect to therare jumps between the basins, the system thus performsnothing but a simple Markov walk.6

Focusing on this Markovian state-to-state dynamics,i.e. coarse-graining the time evolution to the discreterare events, is the central idea behind a kMC simulation.Methodologically, this is realized by moving to a stochas-tic description that focuses on the evolution with time t ofthe probability density function Pi(t) to find the systemin state i representing the corresponding PES basin i.This evolution is governed by a master equation, whichin view of the just discussed system properties is of asimple Markovian form6,

dPi(t)dt

= !!

j !=i

kijPi(t) +!

j !=i

kjiPj(t) , (1)

where the sums run over all system states j. The equa-

Molecular Dynamics:the whole trajectory

Kinetic Monte Carlo:coarse-grained hops

FIG. 2: Schematic top view explaining the di!erences be-tween a MD (left panel) and a kMC (right panel) trajectory.Sketched is the path covered by an adsorbate that di!uses overthe surface by rare hops to nearest-neighbor sites. Whereasthe MD trajectory resolves the short-time vibrational dynam-ics around the stable adsorption sites explicitly, this is coarse-grained into the rate constants in the kMC simulations so thatthe corresponding trajectory consists of a sequence of discretehops from site to site.

tion thus merely states that the probability to find thesystem in a given state i at any moment in time t isreduced by the probabilities to jump out of the presentstate i into any other basin j, and is increased by theprobabilities to jump from any other basin j into thepresent state i. These various probabilities are expressedin the form of rate constants kij , which give the averageescape rate from basin i to basin j in units of inversetime. Because of the Markovian nature of the state-to-state dynamics, also these quantities are independent ofthe system history and thus an exclusive function of theproperties of the two states involved. Assuming for nowthat all rate constants for each system state are known,the task of simulating the time evolution of the rare-eventsystem is in this description then shifted to the task ofsolving the master equation, Eq. (1).

As we will see in the application example below, thetypical number of states in models for a reactive surfacechemistry is so huge that an analytic solution of the corre-sponding high-dimensional master equation is unfeasible.Following the usual stochastic approach of Monte Carlomethods14,15, the idea of a kMC algorithm is instead toachieve a numerical solution to this master equation bygenerating an ensemble of trajectories, where each trajec-tory propagates the system correctly from state to statein the sense that the average over the entire ensemble oftrajectories yields probability density functions Pi(t) forall states i that fulfill Eq. (1).

Since this ensemble aspect of kMC trajectories is quitecrucial, let me return to our example of the di!using ad-sorbate to further illustrate this point. Figure 2 showsa schematic top view of our model surface sketching thepath covered by the adsorbate in a time span coveringa few of the rare di!usion events. The left panel showsthe trajectory as it would have been obtained in a MD

Molecular Dynamics wastes time modeling the 109 thermal vibrations between diffusion events.

Reuter, K. Wiley (2009)

Page 35: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

35

start with

complete list of processes at

t=0 Execute random process from list

Advance time clockUpdate processes

finish

Page 36: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

36

kMC can yield accurate results for model systems.

inequivalent rate constants that need to be computed,and this number increases steeply when accounting forthe constellation in the second, third and so on nearest-neighbor shells. On the other hand, depending on thenature of the surface chemical bond such lateral interac-tions also decay away more or less quickly with distance,so that either a truncation or a suitable interpolation45

is possible, in either case with only a quite finite amountof di!erent DFT rate constant calculations remaining.

This leaves the problem of how to determine in thefirst place which lateral interactions are actually opera-tive in a given system, keeping additionally in mind thatlateral interactions especially at metal surfaces are by nomeans necessarily restricted to just the pairwise inter-actions discussed in the example. For site-specific ad-sorption the rigorous approach to this problem is to ex-pand DFT energetics computed for a number of di!erentordered superstructures into a lattice-gas Hamiltonian.2.This cluster expansion technique46–48 is to date primarilydeveloped with the focus on lateral interactions betweensurface species in their (meta)stable minima2,45,49–53, butgeneralizing the methodology to an expansion of the lo-cal environment dependence of the transition states isstraightforward. Due to the computational constraintsimposed by the expensive transition-state searches, ex-isting expansions in first-principles kMC work in the fieldare often rather crude, truncating the dependence on thelocal environment often already at the most immediateneighbor shells. In this situation, it has been suggested54

that a less rigorous alternative could be to resort to semi-empirical schemes like the unity bond index-quadraticexponential potential (UBI-QEP) method55 to accountfor the e!ects of the local environment. In either case,great care has again to be taken that applying any suchapproximations does not lead to sets of rate constantsthat violate the detailed balance criterion. Particularlyin models with di!erent site types and di!erent surfacespecies, this is anything but a trivial task.

III. A SHOWCASE

Complementing the preceding general introduction tothe underlying concepts, let me continue in this Sectionwith a demonstration of how first-principles kMC simu-lations are put into practice. Rather than emphasizingthe breadth of the approach with a multitude of di!er-ent applications, this discussion will be carried out us-ing one particular example, namely the CO oxidation atRuO2(110). With a lot of theoretical work done on thesystem, this focus enables a coherent discussion of thevarious facets, which I feel is better suited to provide animpression of the quality and type of insights that first-principles kMC simulations can contribute to the fieldof heterogeneous catalysis (with refs. 56,57 e.g. pro-viding similar compilations for epitaxial growth relatedproblems). Since the purpose of the example is primar-ily to highlight the achievements and limitations of the

cus site bridge sitebridge

sitescussites

[001

]

3.12

Å

[110]_

6.43 Å

FIG. 5: Top view of the RuO2(110) surface showing thetwo prominent active sites (bridge and cus). Ru = light,large spheres, O = dark small spheres. When focusing onthese two site types, the surface can be coarse-grained to thelattice model on the right, composed of alternating rows ofbridge and cus sites. Atoms lying in deeper layers have beenwhitened in the top view for clarity.

methodological approach, su"ce it to say that the moti-vation for studying this particular system comes largelyfrom the extensively discussed pressure gap phenomenonexhibited by “Ru”-catalysts, with the RuO2(110) surfacepossibly representing a model for the active state of thecatalyst under technologically relevant O-rich feeds (seee.g. ref. [58] for more details and references to the origi-nal literature).

A. Setting up the model: Lattice, energetics andrate constant catalog

To some extent the system lends itself to a modelingwith first-principles kMC simulations, as extensive sur-face science experimental59 and DFT-based theoreticalstudies29 have firmly established that the surface kinet-ics is predominantly taking place at two prominent activesites o!ered by the rutile-structured RuO2(110) surface,namely the so-called coordinately unsaturated (cus) andthe bridge (br) site. As illustrated in Fig. 5 this leadsnaturally to a lattice model where these two sites are ar-ranged in alternating rows, and to consider as elementaryprocesses the adsorption and desorption of O and COat the bridge and cus sites, as well as di!usion and sur-face chemical reactions of both reaction intermediates ad-sorbed at these sites. With a very small DFT-computedCO2 binding energy to the surface29, the surface reac-tions can furthermore be modeled as associative desorp-tions, i.e. there is no need to consider processes involvingadsorbed CO2.

Another benign feature of this system is its extremelocality in the sense that DFT-computed lateral interac-tions at the surface are so small that they can be ne-glected to a first approximation.29 Considering the ex-haustive list of non-correlated, element-specific processesthat can occur on the two-site-type lattice then leads to26 di!erent elementary steps, comprising the dissociative

110-20 10-10

log(TOF)

pO (atm)2

p CO

(atm

)

110-20 10-1010-20

10-10

COCObrbr/CO/COcuscus

COCObrbr//--

-- // -- OObrbr/O/Ocuscus

1

350K

FIG. 9: Left panel: Equivalent plot of the first-principleskMC-determined surface composition as in the middle panelof Fig. 8, yet at T = 350 K (see Fig. 8 for an explana-tion of the labeling). Right panel: Map of the correspondingturn-over frequencies (TOFs) in 1015 cm!2s!1: White areashave a TOF < 103 cm!2s!1, and each increasing gray levelrepresents one order of magnitude higher activity. Thus, theblack region corresponds to TOFs above 1011 cm!2s!1, whilein a narrow (pCO, pO2) region the TOFs actually peak over1012 cm!2s!1. (From ref. [29]).

much lower reactant partial pressures, which, however,corresponds exactly to the same range of O2 and COchemical potentials as in the corresponding kMC diagramat T = 600K in the middle panel of Fig. 8. This is doneto briefly address the general notion of thermodynamicscaling, which expects equivalent surface conditions inthermodynamically similar gas phases and which is thusoften employed to relate results from surface science stud-ies performed under ultra-high vacuum conditions andlow temperatures to catalytically relevant environmentsat ambient pressures and elevated temperatures. If sucha scaling applies, the topology of the two diagrams atthe two temperatures would be exactly the same, withonly the width of the white coexistence regions varyingaccording to the changing configurational entropy. Com-paring the two kMC diagrams in Figs. 8 and 9 it is clearthat scaling is indeed largely present in the sense that thetransition from O-poisoned to CO-poisoned state occursroughly at similar chemical potentials. Nevertheless, indetail notable di!erences due to the surface kinetics canbe discerned, cautioning against a too uncritical use ofthermodynamic scaling arguments and emphasizing thevalue of explicit kinetic theories like kMC to obtain thecorrect surface structure and composition at finite tem-peratures.

C. Parameter-free turnover frequencies

Besides the surface populations another importantgroup of quantities that is straightforward to evaluatefrom kMC simulations of steady-state operation are theaverage frequencies with which the various elementaryprocesses occur. Apart from providing a wealth of in-

xx xx x x xxxx

x

x

x

x

x

xxxx

xx x xxx

x

x

xxx

xx x

2

pco = 10-10 atm pO2 = 10-10 atm

pCO (10-9 atm)0.0 1.0 2.0 3.0pO2 (10-9 atm)

0.0 1.0 2.0 3.0

3

2

1

0

TOF C

O2

(1012

cm-2

s-1)

6543210

FIG. 10: Comparison of the steady-state TOFs at T = 350 Kfrom first-principles kMC (solid line) and the experimentaldata from Wang et al.68 (dotted line). Shown is the depen-dence with pO2 at pCO = 10!10 atm (left), and the depen-dence with pCO at pO2 = 10!10 atm (right), cf. the overallTOF-plot in Fig. 9. (From ref. [29]).

formation illuminating the on-going chemistry, this alsoyields the catalytic activity as the sum of the averagedfrequencies of all surface reaction events. Properly nor-malized to the surface area, a constant (T, pO2 , pCO) first-principles kMC-run thus provides a parameter-free accessto the net rate of product formation or turn-over fre-quency, TOF (measured in molecules per area and time).

A corresponding TOF-plot for the same range of gas-phase conditions as discussed for the “kinetic phase di-agram” at T = 350K is also included in Fig. 9.29 Thecatalytic activity is narrowly peaked around gas-phaseconditions corresponding to the transition region whereboth O and CO are present at the surface in appreciableamounts, with little CO2 formed in either O-poisoned orCO-poisoned state. On a conceptual level, this is notparticularly surprising and simply confirms the view ofheterogeneous catalysis as a “kinetic phase transition”phenomenon60,63,64, stressing the general importance ofthe enhanced dynamics and fluctuations when the sys-tem is close to an instability (here the transition from O-covered to CO-covered surface). Much more intriguing isthe quantitative agreement that is achieved with existingexperimental data68 as illustrated in Fig. 10. Recallingthat the calculations do not rely on any empirical inputthis is quite remarkable. In fact, considering the multi-tude of uncertainties underlying the simulations, in par-ticular the approximate DFT-GGA energetics enteringthe rate constants, such an agreement deserves furthercomment and I will return to this point in more detail inthe final frontiers Section of this Chapter.

In the catalytically most active coexistence region be-tween O-poisoned and CO-poisoned state, the kinetics ofthe on-going surface chemical reactions builds up a sur-face population in which O and CO compete for eithersite type at the surface. This competition is reflected bythe strong fluctuations in the site occupations visible inFig. 7, but leads also to a complex spatial distribution ofthe reaction intermediates at the surface. In the absence

110-20 10-10

log(TOF)

pO (atm)2

p CO

(atm

)

110-20 10-1010-20

10-10

COCObrbr/CO/COcuscus

COCObrbr//--

-- // -- OObrbr/O/Ocuscus

1

350K

FIG. 9: Left panel: Equivalent plot of the first-principleskMC-determined surface composition as in the middle panelof Fig. 8, yet at T = 350 K (see Fig. 8 for an explana-tion of the labeling). Right panel: Map of the correspondingturn-over frequencies (TOFs) in 1015 cm!2s!1: White areashave a TOF < 103 cm!2s!1, and each increasing gray levelrepresents one order of magnitude higher activity. Thus, theblack region corresponds to TOFs above 1011 cm!2s!1, whilein a narrow (pCO, pO2) region the TOFs actually peak over1012 cm!2s!1. (From ref. [29]).

much lower reactant partial pressures, which, however,corresponds exactly to the same range of O2 and COchemical potentials as in the corresponding kMC diagramat T = 600K in the middle panel of Fig. 8. This is doneto briefly address the general notion of thermodynamicscaling, which expects equivalent surface conditions inthermodynamically similar gas phases and which is thusoften employed to relate results from surface science stud-ies performed under ultra-high vacuum conditions andlow temperatures to catalytically relevant environmentsat ambient pressures and elevated temperatures. If sucha scaling applies, the topology of the two diagrams atthe two temperatures would be exactly the same, withonly the width of the white coexistence regions varyingaccording to the changing configurational entropy. Com-paring the two kMC diagrams in Figs. 8 and 9 it is clearthat scaling is indeed largely present in the sense that thetransition from O-poisoned to CO-poisoned state occursroughly at similar chemical potentials. Nevertheless, indetail notable di!erences due to the surface kinetics canbe discerned, cautioning against a too uncritical use ofthermodynamic scaling arguments and emphasizing thevalue of explicit kinetic theories like kMC to obtain thecorrect surface structure and composition at finite tem-peratures.

C. Parameter-free turnover frequencies

Besides the surface populations another importantgroup of quantities that is straightforward to evaluatefrom kMC simulations of steady-state operation are theaverage frequencies with which the various elementaryprocesses occur. Apart from providing a wealth of in-

xx xx x x xxxx

x

x

x

x

x

xxxx

xx x xxx

x

x

xxx

xx x

2

pco = 10-10 atm pO2 = 10-10 atm

pCO (10-9 atm)0.0 1.0 2.0 3.0pO2 (10-9 atm)

0.0 1.0 2.0 3.0

3

2

1

0

TOF C

O2

(1012

cm-2

s-1)

6543210

FIG. 10: Comparison of the steady-state TOFs at T = 350 Kfrom first-principles kMC (solid line) and the experimentaldata from Wang et al.68 (dotted line). Shown is the depen-dence with pO2 at pCO = 10!10 atm (left), and the depen-dence with pCO at pO2 = 10!10 atm (right), cf. the overallTOF-plot in Fig. 9. (From ref. [29]).

formation illuminating the on-going chemistry, this alsoyields the catalytic activity as the sum of the averagedfrequencies of all surface reaction events. Properly nor-malized to the surface area, a constant (T, pO2 , pCO) first-principles kMC-run thus provides a parameter-free accessto the net rate of product formation or turn-over fre-quency, TOF (measured in molecules per area and time).

A corresponding TOF-plot for the same range of gas-phase conditions as discussed for the “kinetic phase di-agram” at T = 350K is also included in Fig. 9.29 Thecatalytic activity is narrowly peaked around gas-phaseconditions corresponding to the transition region whereboth O and CO are present at the surface in appreciableamounts, with little CO2 formed in either O-poisoned orCO-poisoned state. On a conceptual level, this is notparticularly surprising and simply confirms the view ofheterogeneous catalysis as a “kinetic phase transition”phenomenon60,63,64, stressing the general importance ofthe enhanced dynamics and fluctuations when the sys-tem is close to an instability (here the transition from O-covered to CO-covered surface). Much more intriguing isthe quantitative agreement that is achieved with existingexperimental data68 as illustrated in Fig. 10. Recallingthat the calculations do not rely on any empirical inputthis is quite remarkable. In fact, considering the multi-tude of uncertainties underlying the simulations, in par-ticular the approximate DFT-GGA energetics enteringthe rate constants, such an agreement deserves furthercomment and I will return to this point in more detail inthe final frontiers Section of this Chapter.

In the catalytically most active coexistence region be-tween O-poisoned and CO-poisoned state, the kinetics ofthe on-going surface chemical reactions builds up a sur-face population in which O and CO compete for eithersite type at the surface. This competition is reflected bythe strong fluctuations in the site occupations visible inFig. 7, but leads also to a complex spatial distribution ofthe reaction intermediates at the surface. In the absence

Reuter, K. Wiley (2009)

Page 37: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

37

A key advantage of kMC is the spatial resolution and the ability to model adsorbate-adsorbate interactions.

Neighboring molecules can affect the stability of a species or transition state.

ΔHi = ΔHi,0 + cjδ j ,nnj=1

Nspecies

∑ + cj ,kδ j ,nnδk ,nnk=1

Nspecies

∑j=1

Nspecies

∑ + ...

Ea = Ea,0 + ε jδ j ,nnj=1

Nspecies

∑ + ε j ,kδ j ,nnδk ,nnk=1

Nspecies

∑j=1

Nspecies

∑ + ...

Jansen, APJ. chembond.catalysis.nl/kMC

Page 38: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Summary:Kinetic Monte Carlo

38

Pros:- detailed chemistry over large time scale.

- accurate representation surface heterogeneity.

Cons:- limited in length scale.

- difficult to couple with continuum(i.e. no transport limitations).

- “home cooked” code.

Page 39: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

techniques, and specifically Density-Functional Theory (DFT), arerevolutionizing our thinking of catalytic reactions. Still, our abilityto describe, and eventually control, chemical transformations byfirst-principles modeling, at the molecular level, is hindered bymultiple challenges.

In this paper, we provide a perspective on multiscale modelingfor the development and simulation of catalytic reaction mechan-isms. First, we provide an overview of the length and time scalesin reacting systems, of the objectives of multiscale modeling, andof the challenges in first-principles modeling of chemical reac-tions and reactors. We also underscore the need for detailedreaction models by way of a few examples. The greater part of thereview then focuses on mean-field microkinetic models and theirhierarchical multiscale refinement. Emerging topics in computa-tion-driven catalyst design and uncertainty quantification are alsoreviewed. Recent developments in ab initio kinetic Monte Carlosimulations are then presented, and structure-dependent micro-kinetic models are discussed. New methods to describe catalyticchemistry in solution are outlined and an example from thehomogeneous catalytic dehydrogenation of fructose to 5-hydro-xylmethylfurfural is summarized. Finally, concluding remarks andan outlook are given.

2. Overview of multiscale modeling of chemical reactions andreactors

2.1. Scales in reacting systems

There are at least three scales encountered in a chemical reactor(Fig. 1). At the microscopic, or electronic, length and time scales(bottom of Fig. 1), adsorbate–catalyst and adsorbate–adsorbateinteractions determine the potential energy surface and thus thefree energy barrier and entropy of the chemical transformation. Acoarse description at this scale is the free energy of transformationfrom reactants to the transition state (TS) and then to products. Thethermal rate constant is a convenient way of coarse-graining theinformation from this scale, and quantum mechanical methods areideally suited, at least in principle (see below), for this task.

Given a list of reaction events and their rate constants,adsorbates arrange themselves in spatial configurations or pat-terns, as a result of the collective behavior of the ensemble of allspecies. At this mesoscopic scale (middle of Fig. 1), the collectivebehavior has to be averaged over length and time scales that aremuch larger than the characteristic length and time scale of the

underlying pattern – or what is known as the correlation length –in order to compute the reaction rate. This can be achieved vianon-equilibrium statistical mechanics techniques. Due to the fastvibrations of adsorbates with respect to the reaction time scales,adsorbates are typically thermally equilibrated, and reactionevents can be thought of as rare events, i.e., over the time scaleof a chemical reaction, the system loses its memory and can beapproximated as a Markov process. The kinetic Monte Carlo(KMC) method is the most commonly used statistical techniquefor averaging spatiotemporal events and providing the reactionrate (Bortz et al., 1975; Chatterjee and Vlachos, 2007).

At the macroscopic (reactor) scale (top of Fig. 1), there aregradients in fluid flow, concentration and temperature fields overscales that are typically much larger than the spatial inhomo-geneity of the patterns of adsorbates. As a result, the reaction ratecomputed at the mesoscopic scale can be applied over a certainlength scale (discretization size) of a chemical reactor. Due tospatial macroscopic gradients, the rate has to be evaluated at alldiscretization points of the macroscopic (reactor) domain.

At each scale, computation can be done with various methodswhose accuracy and cost vary. As one moves from left to right ofthe graph at each scale, the accuracy increases at the expense ofcomputational intensity. Thus, at each scale, one can think of ahierarchy of methods. The accuracy of these methods does notvary in a continuous fashion, i.e., each method is different. Typicalmethods are depicted in Fig. 1. Hierarchy adds a new dimensionto multiscaling: at each length and time scale, more than onemodel can be employed in the same simulation scheme, in orderto refine the results or calculate error estimates.

2.2. Objectives of multiscale modeling

The early vision of multiscale modeling was rooted in the bottom-up modeling strategy for predicting the macroscopic (reactor)behavior from microscopic scale calculations (Raimondeau andVlachos, 2002), as shown in Fig. 2. This approach naturally leads toprocess design, control, and optimization with unprecedented accu-racy. It departs significantly from the empirical process design andcontrol strategies of the past, whereby fitting to experimental datawas essential to model building.

Due to the disparity in length and time scales over which varioustools apply (Fig. 2), the straightforward, if not the only, way to reachmacroscopic scales is by coupling models describing phenomena atdifferent scales. Over the past 15 years or so, several algorithms havebeen developed to achieve this bi-directional or two-way coupling(the branches of multiscale modeling are discussed elsewhere(Vlachos, 2005). The structural difference of models across scales(continuum vs. discrete and deterministic vs. stochastic) leads

Semi-empirical: UBI-QEP, TST

ab initio:DFT, TST, DFT-

MD

Continuum:MF-ODEs

Discrete: KMC

Ideal: PFR, CSTR, etc.

ComputationalFluid Dynamics

(CFD)

Mesoscopic:PDEs

Discrete:CGMC

Pseudo-homogeneous:Transport correlations

DFT-based correlations, BEPs

Catalyst/adsorbedphase:

Reaction rate

Reactor scale:Performance

Electronic:Parameter estimation

Accuracy, cost

Fig. 1. Schematic of three scales and a possible hierarchy of models at each scale.At each scale, additional models may exist. The accuracy and cost increase fromleft to right. Acronyms from top to bottom: PRF, plug flow reactor; CSTR,continuously stirred tank reactor; ODE, ordinary differential equation; PDE, partialdifferential equation; CG-KMC, coarse-grained kinetic Monte Carlo; KMC, kineticMonte Carlo; UBI-QEP, unity bond index-quadratic exponential potential; TST,transition state theory; DFT, density functional theory; GA, group additivity; BEP,Brønsted–Evans–Polanyi; QM/MM, quantum mechanics/molecular mechanics.

Fig. 2. Schematic of various models operating at various scales. Redrawn fromVlachos (2005).

M. Salciccioli et al. / Chemical Engineering Science 66 (2011) 4319–43554320

39

Kinetic Monte Carlo

Transition state theory

Part III: Mean field theory

Mean field theory

Page 40: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Part III: Mean Field Theory

40

MFT assumes a uniform distribution of adsorbates and catalyst sites.

• Can span a large range of length and time scales.

• Computationally efficient

• Only real option for process modeling.

• Lots of software available (CHEMKIN, CANTERA)

Page 41: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

MFT is a poor approximation for heterogeneous processes.

Mean Field Approximationprobably not bad probably not good

41

Page 42: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

MFT generally is more accurate at higher temperatures.

42

Disordered surfaces are higher in entropy.• desorption increases with temperature, yielding more

empty sites.

• repulsive lateral interactions increase homogeneity.

ΔG = ΔH − TΔS

Page 43: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

One advantage of MFT is the ability to couple detailed chemistry with fluid mechanics.

43

platinum incorporating both gas-phase and surface mechanisms[7], but this geometry is of considerably lesser complexity ascompared to the above-referred gauze.

For the first time, this paper provides a rational approach formodeling the CPO of CH4 over Pt-gauze reactor. To this end,detailed kinetic schemes for gas-phase and surface chemistrieshave been developed [8], evaluated and implemented in a three-dimensional gauze reactor model, coupled with the flow fieldand heat transfer. This model closely describes realistic reactorfeatures. Our results indicate that both homogeneous andheterogeneous processes should be simultaneously implemen-ted in order to accomplish a solid reactor modeling.

2. Reactor modeling

2.1. Gauze reactor and computational domain

The gauze used in our simulations, is schematically shown inFig. 1. The gauze catalyst consists of two rows of six parallelplatinum wires placed on top of each other (Fig. 1(a)). Theaverage distance between the centers of two individual wires(diameter = 0.20 mm) was 8.2 ! 10"4 m (Fig. 1(b)). Due tosymmetry considerations, only a quarter of a single mesh of thegauze catalyst had been taken into account into the computa-tional grid, including a single contact point between two wires.The domain extends for half of the distance between centers of

two individual wires (amount 4.1 ! 10"4 m) in the y-direction(Fig. 1(c)). At the reactor inlet, the reactive mixture (volumetricCH4/O2 ratio = 2.5 diluted in 80 vol.% He) flows at a uniforminlet velocity of 10 m/s and at 600 K.

2.2. Numerical model

Modeling the gauze reactor involves a numerical solutionfor the Navier Stokes equations, including the coupling ofdetailed gas-phase and surface reaction mechanisms byassuming interaction between them via stable and radicalspecies (see Section 3.1) and considering that energy andmass balances should be satisfied at the interface solid–gas.Arrhenius type expressions were adopted to model thereaction rate in the chemical source terms, which weredivided into homogeneous (gas-phase) and heterogeneous(surface) reactions. The enthalpy of the species wascalculated by integrating the heat capacity at constantpressure (cp) over the working temperature range, the cpvalues were obtained by a polynomial approach using datafrom the JANAF thermodynamic database [9]. The flow fieldwas computed using the commercial computational fluiddynamics code FLUENT [10] coupled with externalsubroutines developed in our group to model gas-phase andsurface chemistries. Details on this are extensively describedin the DETCHEM manual [11].

R. Quiceno et al. / Catalysis Today 119 (2007) 311–316312

Fig. 1. Platinum gauze modeled in this study: (a) original gauze and (b) detail of wire intersections. Shadow area in (b) corresponds to the selected computational

domain presented in (c).

ref. [8] was simplified and only the most relevant reactionswere considered in the 3D simulations, the reduced mechan-ism consisted of 150 reactions and 30 species and it isschematically shown in Fig. 2.

3.1.2. Surface mechanismIgnition and steady-state operation during catalytic combus-

tion of methane over platinum was previously simulated [14].The mechanism used in ref. [14] was completed by includingadditional elementary steps [19]: initial adsorption for methanewith adsorbed atomic oxygen and hydroxyl species anddesorption pathways for the fuel. The extended kinetic

mechanism consists of 11 species and 32 elementary reactionsand can be downloaded from ref. [11].

3.2. Model evaluation with experimental results

Our model has been evaluated with the experimental resultsreported in refs. [5,6]. These authors reported CO, CO2 andH2O as the main reaction products at T < 1270 K. Fig. 3displays experimental data (points) and the results from oursimulations (lines). CH4 and O2 conversions were nottemperature dependent, while the selectivity of CO is stronglyinfluenced by the reaction temperature. Simulations coupling

R. Quiceno et al. / Catalysis Today 119 (2007) 311–316314

Fig. 4. (a–d) Results from the 3D simulations, including profiles of temperature, velocity and concentration of the main species. Conditions: volumetric CH4/O2

ratio = 2.5, V0 = 10 m/s, T0 = 600 K and P = 1.3 bar. (e and f) V0 = 1 m/s.Quinceno et al. Cat. Today, 119 (2007)

Page 44: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Conservation of mass, energy, and momentum share a common formalism.

44 Quinceno et al. Cat. Today, 119 (2007)

accumulation = bulk transport+ molecular transport+ chemical reaction

MassEnergyMomentum

⎢⎢⎢

⎥⎥⎥in

MassEnergyMomentum

⎢⎢⎢

⎥⎥⎥out

Page 45: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Simple systems can be modeled with ideal reactors.

45

Batch reactor:perfect mixing

concentration versus time

Perfectly stirred reactor:perfect mixing

lower conversion per unit volume (generally)

concentration versus time

Plug flow reactor:no radial gradients

higher conversion per unit volume

concentration versus position

Page 46: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Networks of ideal reactors can model more complex phenomena.

46

Page 47: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Two-dimensional reactor models include mass transport limitations.

47

CO*

Pt*O*

O2(z,r)

Page 48: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

More complex geometries or problems require computational fluid dynamics.

48

Most CFD codes are developed for fluid mechanics (chemistry is an afterthought).• CFD code spends >95% computer time on chemistry,

not transport

• Simplified, lumped models must be used➡New computational paradigms are needed

Page 49: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Summary:Mean Field Theory

49

Pros:- covers a broad range of time and length scales.

- Allows for easy modeling of transport limitations.

- easily coupled with reactor models and CFD.

- standard software available.

Cons:- mean field is a poor approximation for inherently

heterogeneous phenomena.

Page 50: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Part IV: Understanding the results

50

So you’ve successfully modeled a system with 100’s of species and 1000’s of reactions.

Now what?

Page 51: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Sensitivity analysis:determine which rates are most important.

51

xi =∂ ln TOFCO2( )

∂ ln ki( ) si =∂ ln C2H4[ ]( )

∂ ln ki( )

ature range we obtain an apparent activation energy of Eapp = 2.85eV, by fitting a straight line to the kMC data.

We know the activation energies of all elementary processes inour kMC simulation and none is even close to 2.85 eV, which dem-onstrates immediately that the deduced apparent activation en-ergy does not reflect the activation energy of one bottleneckprocess. Indeed, in the temperature range in which the straight linefitting was performed there are instead three processes having asizeable xri : The adsorption of CO on the cus sites, the desorptionof CO from the cus sites and the dissociative adsorption of oxygeninto a pair of cus sites. Of these, only the desorption of COcus is acti-vated, having a barrier of 1.3 eV11,12. Since x!i for this process is "2throughout the temperature range of interest, Eq. (12) gives Eapp "

2.6 eV, which considering the approximations made is fairly closeto the true value of 2.85 eV.

4.5. Summary

We have used first-principles kinetic Monte Carlo simulationsto study the usefulness of two definitions of the DRC: One givenby Campbell [4,6] and one defined here. Both definitions studythe ‘‘linear response” of the turn-over frequency to a change inone of the rate constants of the reaction network. Analyzing thecomplementary insight provided by the two definitions over awide range of gas-phase conditions we conclude that they cor-rectly reflect the knowledge we have about the system from thedetailed data available from the first-principles kMC simulations,i.e. the total TOF and the contribution to it from the various reac-tion mechanisms, the surface coverages, as well as the occurrenceof the individual elementary processes. The conclusions reached bycalculating the DRC are furthermore non-trivial in the sense thatthey could not have been reached by merely examining the magni-tude of the rate constants or of the activation energies of the ele-mentary processes.

In the pressure regime in which the catalyst is most active theDRC analysis identifies an entire group of processes to which theTOF is very sensitive. While there is thus no single ‘‘rate limitingstep” this number of processes controlling the overall CO2 produc-tion is small. This indicates that if the rate constants of these pro-cesses are known accurately, the kMC procedure will producecorrect results even if the other rate constants are inaccurate. Wehave confirmed this by direct calculation of the variation of theTOF with some of the unimportant rate constants. In some casesone can change a rate constant by several orders of magnitude withno effect on the TOF. Apart from providing a tool for analyzing themechanism of a complex set of catalytic reactions, the DRC tells ustherefore which aspects of the reaction mechanism must be trea-ted accurately and which can be studied by less accurate and moreefficient methods. In this sense we argue that a sensitivity analysisbased on the DRC can be a useful tool towards establishing a con-trol of the propagation of error from the electronic structure calcu-lations to the statistical simulations in first-principles kineticMonte Carlo approaches.

References

[1] M. Boudard, K. Tamaru, Catal. Lett. 9 (1991) 15.[2] J.A. Dumesic, J. Catal. 185 (1999) 496.[3] A. Baranski, Solid State Ionics 117 (1999) 123.[4] C.T. Campbell, Topics Catal. 1 (1994) 353.[5] J.A. Dumesic, J. Catal. 204 (2001) 525.[6] C.T. Campbell, J. Catal. 204 (2001) 520.[7] K. Reuter, M.V. Ganduglia-Pirovano, C. Stampfl, M. Scheffler, Phys. Rev. B 65

(2002) 165403.[8] K. Reuter, M. Scheffler, Phys. Rev. B 65 (2001) 035406.[9] K. Reuter, M. Scheffler, Phys. Rev. B 68 (2003) 045407.[10] K. Reuter, M. Scheffler, Phys. Rev. Lett. 90 (2003) 046103.[11] K. Reuter, D. Frenkel, M. Scheffler, Phys. Rev. Lett. 93 (2004) 116105.[12] K. Reuter, M. Scheffler, Phys. Rev. B 73 (2006) 045433.[13] K. Reuter, C. Stampfl, M.V. Ganduglia-Pirovano, M. Scheffler, Chem. Phys. Lett.

352 (2002) 311.[14] B. Temel, H. Meskine, K. Reuter, M. Scheffler, H.J. Metiu, Chem. Phys. 126

(2007) 204711.[15] J. Dumesic, in: G. Ertl, H. Knözinger, F. Schüth, J. Weitkamp (Eds.), Handbook of

Heterogeneous Catalysis, second ed., Wiley-VCH, Weinheim, 2008.[16] I. Chorkendorff, J.W. Niemantsverdriet, Concepts in Modern Catalysis and

Kinetics, Wiley-VCH, Weinheim, 2003.[17] H. Lynggard, A. Andreasen, C. Stegelmann, P.I. Stoltze, Prog. Surf. Sci. 77 (2004)

71.

Fig. 5. Upper panel: Plot of the logarithm of the simulated steady-state turnoverfrequency for CO2 production versus the inverse temperature (1/kBT) for pO2

= 1 atmand pCO = 2 atm. Lower panel: Computed xi and xri for these gas-phase conditions.See text for an explanation of the nomenclature used to describe the differentelementary processes. The DRC values for all processes not shown are practicallyzero on the scale of this figure.

1730 H. Meskine et al. / Surface Science 603 (2009) 1724–1730

Page 52: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

52 Blaylock et al. JPCC, 113 (2009)

Flux path analysis:determine which intermediates are most important.

Page 53: Kinetic modeling in heterogeneous catalysis. · describing heterogeneous catalytic chemistry via first principles is that phenomena involving chemical reactions and reactors are

Scaling relations adsorbate binding energies

activation energiespre-exponential factors

BEP

Reaction type

Sensitivity+ Flux Determine the most important

parameters and refine them.

We can combine all these techniques to build accurate mechanisms with minimal computational effort.

Microkinetic model

Reactor simulations(e.g. TOF, concentration profiles)

53