Sie sind auf Seite 1von 10

MOLECULAR PHYSICS, 20 FEBRUARY 2004, VOL. 102, NO.

4, 361369

First-principles-based kinetic Monte Carlo simulation of nitric oxide decomposition over Pt and Rh surfaces under lean-burn conditions
DONGHAI MEI1, QINGFENG GE1, MATTHEW NEUROCK1*, LAURENT KIEKEN2 and JAN LEROU2 1 Department of Chemical Engineering, University of Virginia, Charlottesville, Virginia 22904, USA 2 NovoDynamics, Inc., Ann Arbor, Michigan 48104, USA
(Received 6 July 2003; accepted 7 January 2004) First-principles-based kinetic Monte Carlo simulation was used to track the elementary surface transformations involved in the catalytic decomposition of NO over Pt(100) and Rh(100) surfaces under lean-burn operating conditions. Density functional theory (DFT) calculations were carried out to establish the structure and energetics for all reactants, intermediates and products over Pt(100) and Rh(100). Lateral interactions which arise from neighbouring adsorbates were calculated by examining changes in the binding energies as a function of coverage and different coadsorbed congurations. These data were tted to a bond order conservation (BOC) model which is subsequently used to establish the effects of coverage within the simulation. The intrinsic activation barriers for all the elementary reaction steps in the proposed mechanism of NO reduction over Pt(100) were calculated by using DFT. These values are corrected for coverage effects by using the parametrized BOC model internally within the simulation. This enables a site-explicit kinetic Monte Carlo simulation that can follow the kinetics of NO decomposition over Pt(100) and Rh(100) in the presence of excess oxygen. The simulations are used here to model various experimental protocols including temperature programmed desorption as well as batch catalytic kinetics. The simulation results for the temperature programmed desorption and decomposition of NO over Pt(100) and Rh(100) under vacuum condition were found to be in very good agreement with experimental results. NO decomposition is strongly tied to the temporal number of sites that remain vacant. Experimental results show that Pt is active in the catalytic reaction of NO into N2 and NO2 under lean-burn conditions. The simulated reaction orders for NO and O2 were found to be +0.9 and 0.4 at 723 K, respectively. The simulation also indicates that there is no activity over Rh(100) since the surface becomes poisoned by oxygen.

1. Introduction Potential solutions for NOx removal include direct NO decomposition, selective catalytic reduction and NOx storage. Despite the substantial progress that has been made from both fundamental and applied efforts, there is still an incomplete understanding of the mechanism and no breakthrough material leads. Direct lean burn has become an important problem for highthroughput efforts. Theory, simulation and modelling can help to provide a bridge between surface science studies and actual catalytic performance studies. In addition, it may prove to be invaluable in the generation of ideas for the synthesis of novel materials.

The development of novel catalytic materials that can operate under lean-burn conditions would enable the broad commercialization of spark injection lean-burn engines which offer substantially improved fuel economy [1]. Current three-way exhaust catalysts, however, are inactive under lean-operating conditions. A number of obstacles including the stringent regulations on NOx emissions, oxygen and SOx poisoning, and material stability to operative condition have hindered the development of a working catalyst. Much of the effort over the past ve years has focused on the search for selective NOx reduction [26] and NOx storagereduction [711] catalysts. While fundamental studies have revealed a number of important features of the surface chemistry, the elementary mechanisms and

*Author for correspondence. e-mail: mn4n@virginia.edu


Molecular Physics ISSN 00268976 print/ISSN 13623028 online # 2004 Taylor & Francis Ltd http://www.tandf.co.uk/journals DOI: 10.1080/00268970410001668471

362

D. Mei et al. of the local environment on the kinetics [1723]. This approach has been used to simulate surface kinetic processes such as TPD as well as a microcatalytic system thus providing an initial bridge for both the pressure and materials gaps. This combined DFT/KMC approach maintains the atomic structure and can therefore be used to probe how changes in the specic atomic structure of the catalytic surface might impact its activity and selectivity. This could ultimately be used to help complement experimental design efforts. In this paper, we present a comparative simulation study on the catalytic decomposition of NO over welldened Pt(100) and Rh(100) surfaces under lean-burn conditions. First-principles DFT calculations were used to construct a fairly comprehensive intrinsic kinetic database for the potential species and reaction pathways involved in this system. This database was used along with the parametrized bond order conservation (BOC) model within a kinetic Monte Carlo algorithm to simulate kinetic behaviour of NO decomposition over Pt(100) and Rh(100) under both ultrahigh vacuum (UHV) and lean-burn conditions. The simulation results are compared with experiments.

microkinetics for NO reduction under lean-burn conditions are still incomplete. A detailed understanding of the elementary kinetic processes and how they work together to control lean NOx decomposition over Pt and Rh could be invaluable in the development of novel bimetallic catalysts for NO direct reduction under lean-burn conditions. Advances in rst-principle quantum chemical methods and atomistic simulation algorithms have reached the stage whereby they can be used to elucidate surface kinetics for more complex systems with greater degrees of sophistication. Density functional theory (DFT) has been used to study elementary processes which occur over a wide range of different materials including transition metals, metal oxides, metal sulphides and zeolites. Despite its widespread use, DFT is still limited in its accuracy. Typical errors for adsorption energies are on the order of 5 kcal mol1. In addition, there are systems that are outliners which can be even further off. First-principle calculations, however, have been very effective in complementing experiment and offer guidance into the design of new materials. For example, rst-principles DFT is being used on a routine basis to elucidate plausible mechanisms for model catalytic reactions [1215]. While ab initio methods offer the resolution of elementary steps, catalysis requires following the dynamics over much longer time scales and spatial dimensions. Kinetic Monte Carlo (KMC) simulation offers an effective bridge between the changes in the electronic and atomic structure and the kinetics for catalytic processes. Kinetic Monte Carlo simulation can be used to explicitly track the temporal behaviour of all the surface species as a function of time and processing conditions [16]. We have shown previously how DFT calculations can be used to construct a rst-principles kinetic database and parametrize a semi-empirical lateral interaction model that can be called upon in situ of the Monte Carlo simulations to determine the effects
Table 1.

2. Reaction mechanism and intrinsic kinetics 2.1. Reaction mechanism NO decomposition can proceed either by direct dissociation to form nitrogen and oxygen adatoms or by the coupling of two NO molecules to form N2O which can subsequently react to form N2 and chemisorbed atomic oxygen. At higher coverages, the NO coupling reaction may be as important as the direct NO decomposition path [9, 13, 14, 24, 25]. These paths along with other potential adsorption, desorption and surface reaction steps have been assembled into a plausible mechanism which is consistent with most experimental speculations. The resulting mechanism is presented in table 1. The mechanism proceeds by

DFT results for the atomization energies and binding energies for all proposed reaction intermediates on Pt(100) and Rh(100) surfaces. Binding energy kJ mol1 Atomization energy kJ mol1 569.4 960.6 671.5 1073.4 1225.3 Pt(100) Atop 136.6 118.8 26.5 Bridge 403.1 394.0 214.3 74.4 11.9 4-fold hollow 414.4 359.4 157.8 Atop 353.6 376.8 219.6 177.4 53.9 Rh(100) Bridge 474.4 475.4 263.8 4-fold hollow 547.8 496.7 256.6

Species Oxygen (O2) Nitrogen (N2) Nitrogen (N) Oxygen (O) Nitric Oxide (NO) Nitric Dioxide (NO2) Nitrous Oxide (N2O)

First-principles-based kinetic Monte Carlo simulation of nitric oxide decomposition molecular adsorption of NO. The NO molecules can adsorb on any of the possible sites (atop, 2-fold bridge, 4-fold hollow) available. Molecular oxygen dissociatively adsorbs to form atomically adsorbed oxygen which prefers either the bridge or hollow sites. The adsorbed NO molecule can subsequently dissociate and form atomic nitrogen and oxygen. Atomic nitrogen also prefers the bridge and hollow sites. In addition, adsorbed NO can react with neighbouring oxygen atoms or nitrogen atoms to form NO2 or N2O, respectively. DFT results indicate that the atop sites are the most favourable sites for NO2 and N2O. The adsorbed NO2 and N2O can subsequently desorb from the surface. Finally, two atomic oxygen or nitrogen atoms can recombine and desorb as O2 and N2 respectively. 2.2. Intrinsic kinetics and lateral interactions An intrinsic kinetic database was built from a comprehensive set of periodic DFT calculations on the adsorption energies for the different intermediates at different coverages, the heats of reaction as well as the activation barriers for all of the steps which comprise the mechanism shown in table 1. Plane wave gradient corrected periodic DFT calculations were performed using the VASP code developed by Kresse and co-workers [26, 27] to determine the lowest energy structures and their corresponding energies for all of the results reported herein. The PerdewWang form of the generalized gradient approximation (GGA) was used to calculate exchange and correlation corrections [28]. The metal surface was described by a periodic super cell which contained between foursix metal layers along with a 12 A vacuum layer to electronically separate the repeating slab structure in the z direction. The top layers of the slab, as well as the adsorbate structure, were allowed to freely optimize. The bottom layer of the slab

363

was held xed at the bulk lattice spacing of the metal. Transition states were isolated using a nudged elastic band technique [29]. The resulting intrinsic kinetic information including binding energies, overall reaction energies and activation barriers for some of the steps over Pt(100) and Rh(100) are given in tables 1 and 2. Additionally, the structural information for all adsorbates at different adsorption sites was also calculated and used in the kinetic Monte Carlo algorithm. An integral part of describing the surface reaction kinetics involves accounting for the effects of the local environment on the active sites. Lateral interactions that take place between two or more adsorbates can be characterized as either through-space interactions, which are those that are due to Coulombic and/or steric effects, or through-surface interactions, which are those that result from charge transfer between the adsorbate and metal atoms. These interactions play an important role in dictating the heats of adsorption and surface kinetics as shown from NO TPD and calorimetric experiments over different metals [25, 30]. In this work, we model the through-space interactions by using van der Waals (vdW) terms from the Merck molecular force eld [31]. The through-surface interactions are calculated using a DFT-scaled BOC method [1719]. 2.3. Rate calculation The reaction rate for each elementary step is calculated using transition state theory   Ei ri i exp ; 1 RT where R is the gas constant, T is the temperature and Ei is the activation energy barrier for the elementary reaction i. i is the pre-exponential factor which can either be calculated quantum mechanically or estimated

Table 2.

Activation barriers for elementary reactions in NO reduction over Pt(100) and Rh(100) surfaces. Efor KJ mol1 Erev KJ mol1 Efor KJ mol1 Erev KJ mol1 vfor (S0) (s1) 263.8 184.4 0.0 0.0 0.0 178.1 69.5 458.6 0.0 0.0 0.7 1 1013 1 1013 1 1013 1 1013 1 1013 1 1013 1 1015 1 1013 1 1013 vrev (S0) (s1) 1 1013 1 1013 1 1013 1 1013 1 1013 1 1013 1 1013 0.1 1 1013 1 1013

Reaction NO (g) *$NO* NO* *!N* O* NO* O*!NO2* * NO* N*!N2O* * NO* NO*!N2O* O* N2O*!N2 (g) O* N* N*!N2 (g) 2* O* O*$O2 (g) 2* NO2*!NO2 (g) * N2O*!N2O (g) * 0.0 107.1 133.0 141.9 157.1 0.0 9.0 0.0 118.7 26.5

Pt(100) 214.3 21.0 49.1 92.7 0.0 102.9 140.8 218.6 0.0 0.0 0.0 76.1 182.0 204.7 96.4 0.0 204.4 0.0 177.4 54.0

Rh(100)

364

D. Mei et al. lateral interaction model and incorporated into the core reaction model in the simulation. After initialization, all of the sites on the surface are stochastically sampled in order to construct a cumulative reaction probability distribution that outlines all of the possible kinetic events that can occur. This includes adsorption, surface reaction, surface diffusion and desorption energies. The events that are considered depend on the specic site of the event and its surrounding reaction environment. For each surface reaction event, we calculate the binding energies of reactants as well as the possible product states in order to determine the overall energetics and the effects of reaction environment. The activation barriers for each step are calculated by combining the DFT results with the lateral interactions from our through-surface and through-space interaction models. At any instant in time, ti, the rates for all possible events are added up in order to determine the total rate. The P total rate, ri , is subsequently used along with the variable time-step equation (equation (3)) to determine the time at which the possible event on the surface happens ln RN tv P ; 3 i ri where tv is the variable time step and RN is a random number between 0 and 1. The specic reaction that occurs within the calculated time-step interval is chosen based on the cumulative reaction probability distribution ri si P : 4 i ri The outcome from the simulation includes the detailed structure and composition of the adlayer as a function of time and processing conditions. This temporal and spatial resolution of the surface enables us to compute a range of properties, including either molecular- or site-explicit turnover frequencies (TOFs), activation barriers, surface coverages and other overall averaged energetic properties. 4. Results and discussion 4.1. Temperature programmed reaction 4.1.1. Pt(100) The desorption and decomposition kinetics of NO over the Pt(100) surface were studied by performing a series of kinetic Monte Carlo simulations for temperature-programmed reduction (TPR) under different initial NO surface coverages ranging from 0.1 monolayer (ML) to saturation coverage. These simulations were initialized by simulating the adsorption uptake and equilibration at low temperature (125 K) at a specic partial pressure of NO. Once the surface had

from traditional statistical mechanics. The activation barriers for all of the elementary steps have been determined by DFT calculations over the Pt(100) surface. The barriers over Rh(100) were estimated using a modied DFT-BOC approach. The details of this approach can be found in previous papers [17, 18, 22]. The pre-exponential factors, i , were simply chosen as the traditional statistical mechanical estimates which calculate the changes from the universal frequency constant of 1013 in terms of changes in the modes of vibrational, rotational or translational modes of freedom between the transition state and the reactant state (see for example [32]). Traditional statistical mechanical estimates for unimolecular surface reactions are on the order of 1013, therefore all of the surface reaction steps in the simulation were assumed to be 1013 s1. The preexponential factor for the molecular desorption of O2 was assumed to be 1015 s1 [33]. The calculation of the adsorption rate is characteristically different from the surface elementary reaction rate calculations. The adsorption rate for species i is dened instead as [19, 20]   Ei rad;i s0 Pi As 2pMWi RT0:5 exp ; 2 RT where s0 is the sticking coefcient, Pi is the partial pressure of species i, As is the area of one surface site and MWi is the molecular weight of species i. The sticking coefcients of NO, N2O and NO2 were taken to be 0.7, 1.0 and 1.0, respectively. The sticking coefcient of O2 is assumed to be 0.1. These values were chosen based on previous experimental and microkinetic modelling results for NO decomposition on Pt and Rh [10, 24] .

3. Kinetic Monte Carlo simulation Kinetic Monte Carlo simulation allows us to track the individual molecular transformations as a function of time and processing conditions. The details of this algorithm have been described in previous papers [17 23]. A brief description of the approach is given below. The surface structure dened here is the (100) surface which is represented by a 32 32 atom grid which contains 4098 sites. Periodic boundary conditions were used in the simulations. Many of the structural, electronic and energetic properties calculated from DFT for the full series of intermediates were used as input into the simulation. This includes the binding energies and molecular structures of the adsorbates at all of the possible adsorption sites, vdW radii (physical size) for all of the reactants, intermediates and products along with the reaction energies and activation barriers for elementary surface kinetic processes. This intrinsic kinetic database was subsequently combined with the

First-principles-based kinetic Monte Carlo simulation of nitric oxide decomposition


(a) 0.6
0.5
Desorption and Dissociation

365

sat = 0.54 ML

T = 360 K

NO (ML)

0.4 0.3 0.2 0.1 0.0 250


Dissociation Only

300

350

400

450

500

550

600

T (K)
(b)
0.7

sat = 0.62 ML
0.6 0.5
T = 390 K

0.4 0.3 0.2 0.1 0.0 250

Desorption and Dissociation

Dissociation Only

300

350

400

450

500

550

600

T (K)

Figure 1. Simulation results for the temperature programmed reaction of NO at different initial coverages over (a) Pt(100) and (b) Rh(100).

equilibrated, the desorption steps were turned on and the TPR simulation was started. The simulations follow the elementary diffusion, reaction and desorption processes that occur as temperature is increased at a heating rate of 5 K s1. These simulations were subsequently repeated for different initial partial pressures of NO. The simulation results for NO TPR over the Pt(100) surface are given in gure 1(a). In the low NO coverage range, from 0.1 to 0.2 ML, NO sits almost exclusively at the bridge sites. The adsorbed NO subsequently dissociates to form nitrogen and oxygen adatoms which prefer to occupy neighbouring bridge sites. The resulting simulated TPR spectrum is shown in gure 1(a). NO dissociation appears to begin at 350 K. The maximum peak for NO dissociation falls within the range of 400425 K. The main products are N2 and O2, with very few N2O species formed. The saturation coverage of NO on clean Pt(100) surface, in principle, depends strongly on the nature and the strength of lateral interaction between coadsorbed NO species. The simulations predict an equilibrium saturation coverage of NO over Pt(100) at 125 K to be 0.54 ML. This compares quite favourably with experimental estimation results of

0.5 ML [25, 34, 35]. The strong repulsive interactions between adsorbed NO species at these higher coverages weaken the binding of NO to the surface. The average NO binding energy decreases from 214.3 to 128.8 kJ mol1 as the NO coverage increases from 0.1 to 0.5 ML. At higher NO coverages, direct NO decomposition is not possible since the sites are predominantly lled. Some of the adsorbed NO molecules must rst desorb from the surface in order to create vacant sites that would then allow other NO species to dissociate. NO coupling, however, can occur to some extent at high coverages to form N2O and atomic oxygen. The strong lateral interactions act to lower the intrinsic activation barrier for NO coupling. The simulations show that the NO surface coverage begins to decrease at 300 K for surfaces with coverages in excess of 0.5 ML. At lower coverages, NO desorption occurs at 350 K. The maximum peak for NO reduction is at about 360 K. The dissociation ratio, which is dened as the percentage of the initial amount of NO on the surface which dissociates, is a key measure for NO decomposition. Our simulation indicates that the dissociation ratio is 62% at saturation coverage. The dissociation ratio decreases with increasing initial NO surface coverage. This agrees with the experimental NO dissociation ratio of 66% by TPD study [36] and 75% by high-resolution X-ray photoelectron spectroscopy [34]. Atomic nitrogen formed by NO dissociation prefers to adsorb at either the bridge site or hollow site on a Pt(100) surface depending on the surface coverage. N2 desorption from the Pt(100) appears to follow a secondorder kinetics. This is consistent with the fact that the reaction involves a recombination (or collision) of two surface nitrogen atoms to form N2 which is spontaneously released from the metal surface. The calculated binding energy for N2 is essentially zero. Our DFT results indicate that the activation barrier for associative desorption of N2 from Pt(100) is 9 kJ mol1 at 0.25 ML coverage. This value is nearly the same as the value of 8.7 kJ mol1 reported by Eichler and Hafner for similar DFT calculations [13, 14]. They suggest that high mobility of N atoms on Pt(100) combined with such a low barrier makes the actual barrier for N2 desorption negligible. The desorption of nitrogen is then governed by the diffusion rate for nitrogen atoms on the Pt(100) surface. The rate will be highly coverage dependent. At very high NO coverages, nitrogen recombination will be much more difcult due to the absence of vacant sites necessary to allow for diffusion. It can be expected that more N2O will be formed when the NO coverage is high. The simulation results indicate that more nitrogen atoms are removed in the form of N2O instead of as N2 as the initial coverage

NO (ML)

366

D. Mei et al. Direct dissociation occurs at coverages of NO < 0:2 ML for Pt(100) and NO < 0:3 ML for Rh(100), respectively. N2 and O2 are the main products forming from recombination pathways. N2O formation can occur over Pt(100) at higher NO surface coverages. At low surface coverage, NO dissociation on both surfaces begins around 300 K and is complete around 500 K. At low surface coverages, lateral interactions are relatively minor. This indicates that both surfaces are intrinsically active for NO dissociation. At higher NO coverages, NO dissociation is no longer complete on either surface. The dissociation of NO is highly coverage dependent. Vacant sites have to be created before NO molecules can dissociate into atomic nitrogen and oxygen. At saturation coverages, NO dissociation does not occur until at least 350 K on both surfaces. Some of the NO molecules desorb from the surface at about 300 K. As NO is removed, subsequent NO molecules can react with atomic nitrogen to form N2O. N2O can desorb from the surface or dissociate to form gas phase N2 and adsorbed atomic oxygen. The simulation results indicate that the NO coupling reaction can be an important pathway for NO reduction at high coverage. Perhaps the most important difference between the NO TPD results on Pt and Rh is the temperature at which O2 is formed. Atomic oxygen recombines and desorbs from Pt(100) at 600700 K. The recombination/ desorption temperature for oxygen on Rh(100), however, is 13001400 K which is nearly two times greater than that on Pt(100). This is the result of the much stronger binding for oxygen on Rh than on Pt. This suggests that the removal of oxygen from Rh will be much more difcult than it is on Pt. 4.2. NO reduction under lean-burn condition In addition to modelling TPD under UHV conditions, the kinetic MC simulations can also be used to track actual catalytic kinetics at more realistic operating conditions. This requires that we add in the background partial pressures, run at higher but constant temperatures and allow for adsorption processes to occur. In order to compare our simulation with experimental results [9, 10], we carried out our kinetic simulations at more realistic lean-burn conditions. The partial pressures of NO and O2 were taken to be 0.468 and 60.8 torr, respectively. The temperatures examined here ranged from 573 to 773 K. 4.2.1. Pt(100) The turnover frequency for molecule i (TOFi) was calculated by counting the number of i molecules which desorb from the surface as a function of time. The resulting turnover frequencies for N2 and NO2 from the simulation are given in gure 2. The TOF for N2

increases. The simulation results have shown that N2 desorption temperature ranges from 400 to 450 K. Similarly, O2 desorbs from the Pt(100) surface around 600 up to 800 K. This is consistent with experiments previously reported [35, 37]. In conclusion, the simulation results for NO TPR are in good agreement with observations from various experimental studies including LEED [38], TPD/TPR spectroscopy [36, 37, 39], vibrational spectroscopy [35, 40, 41] and X-ray photoelectron spectroscopy [34, 42]. 4.1.2. Rh(100) The simulation results for the NO TPR spectra over the Rh(100) surface are shown in gure 1(b). The kinetics for NO decomposition over the Rh(100) were found to be somewhat similar to those over Pt(100). The simulations indicate that the NO saturation coverage over Rh(100) is 0.62 ML. This is in good agreement with the experimental value of 0.65 ML reported by Ho and White [43]. The simulation results show that NO dissociation is self-inhibited by NO desorption which does not occur until around 420 K. The desorption of NO creates empty sites which are necessary for NO dissociation. At saturation coverage, about 53% NO dissociates and the remaining 47% NO desorbs from the Rh(100) surface. NO decomposition upon heating is hindered by the available empty 4-fold hollow and bridge sites on the Rh(100) for dissociated nitrogen and oxygen atoms. This is also in good agreement with experimental TPD and second ion mass spectrometry (SIMS) measurements by Hopstaken and Niemantsverdriet [30]. Their results indicated that NO molecules start to desorb around 400 K before NO dissociation occurs and the dissociation ratio is 53% at NO saturation coverage. It was found from the simulation that the upper limit for total dissociation of NO over the Rh(100) surface is 0.3 ML. Below this coverage, adsorbed NO molecules dissociate completely to nitrogen and oxygen atoms. This compares favourably with the experimental result of 0.28 ML [30]. In the simulation, N2 desorbs in the temperature range of 600 to 800 K. Oxygen, on the other hand is much more strongly bound and does not desorb from Rh(100) until 1200 to 1400 K. 4.1.3. Comparison of NO TPR over Pt(100) and Rh(100) The decomposition and desorption of NO over Pt(100) and Rh(100) under UHV conditions show somewhat similar behaviour but there are very important differences. NO adsorbs molecularly on both surfaces at low temperatures. Complete dissociation of NO occurs at lower coverages on both surfaces. At higher coverages, NO must rst desorb before it can begin to dissociate.

First-principles-based kinetic Monte Carlo simulation of nitric oxide decomposition


9.0 8.5
1000 K 800 K 600 K

367

(a) 1.0
0.9 0.8 0.7

ln(TOF) (Site s )

-1

(ML)

-1

8.0 7.5 7.0 6.5 6.0 0.0008

0.6 0.5 0.4 0.3 0.2 0.1

0.0010

0.0012

0.0014
-1

0.0016

0.0018

0.0020

0.0 550

600

650

700

750

800

1/T (K )

T (K)
(b)
1.0 0.9 0.8 0.7

Figure 2. Simulated turnover frequencies for N2 and NO2 formation over Pt(100) for the formation of N2 and NO2 over Pt(100) under catalytic lean-burn conditions. (f) NO2; (g) N2.
12

(ML)

0.6 0.5 0.4 0.3

10
2

ln rN

0.2
8

0.1 0.0 550

600

650

700

750

800

T (K)
Figure 4. Surface coverages of NO and intermediates during NO decomposition under lean-burn conditions: (a) Pt(100) and (b) Rh(100). (h) O (exp.); (i) NO (exp.); (j) NO2 (exp.); (g) O (sim.); (m) NO (sim.); (n) NO2 (sim.); (^) simulated total coverage.

4 4

2 4 ln pNO or ln pO
2

Figure 3.

Simulated reaction orders for NO decomposition under lean-burn conditions over Pt(100).

increases with temperature as expected. In contrast, the TOF of NO2 dropped off at higher temperatures, due to the thermodynamic limitations which drives the reaction towards NO2 dissociation. The simulation can be used as a virtual experiment whereby temperatures and pressures can be systematically changed to determine the apparent activation barriers and reaction orders in the same manner that an experiment would. The apparent activation energy Ea and pre-exponential factor  were calculated by tting our simulation results (see gure 2) to the corresponding Arrhenius expression   Ea rTOF  exp Px Py : RT NO O2 5

The apparent activation energies for N2 and NO2 formation (extracted from 600 to 800 K) calculated from gure 2 are 26.8 and 37.6 kJ mol1, respectively. Previously reported apparent activation barriers for N2 formation from experiment range from 12.9 to 84.1 kJ mol1 for Pt foil to supported Pt/Al2O3 obtained

at different reaction conditions [44, 45]. Our own experiments indicate the apparent activation barrier is 48 kJ mol1 which is 10 kJ mol1 higher than the simulated barrier. By running the simulation at different partial pressures of NO and O2 at constant temperature, we can also determine the reaction orders. The reaction orders of NO and O2 calculated from gure 3 are 0.9 and 0.4 at 723 K. At 773 K, we found that the reaction orders of NO and O2 are 0.55 and 0.06, respectively. This indicates that oxygen inhibition of NO reduction is smaller at higher reaction temperatures. The simulations provide a detailed description of the dynamic changes in the atomic surface structure which allows for the temporal analysis of all molecular conversions as well as occupancy of a specic surface site. The resulting details can also be averaged over the surface. The averaged surface coverages for NO decomposition over Pt(100) as a function of temperature are shown in gure 4(a). The average coverage of oxygen decreases from 0.62 to 0.55 ML as the temperature is increased. The average NO coverage similarly decreases with increasing temperature. As a result, the

368

D. Mei et al. for NO2 desorption is 177.4 kJ mol1 on the Rh(100) surface. The NO2 surface is inactive except for equilibrium that occurs between the forward and reverse rate for NO2 * , NO* O* where * refers to an adsorbed surface intermediate. For the Pt(100) surface, the barrier for NO2 dissociation is 49.1 kJ mol1 while the barrier for NO2 desorption is 118.7 kJ mol1. More importantly, the associative desorption of O2 from Pt occurs within the operating temperature range. This generates a sufcient number of vacant sites for NO decomposition and NO coupling. This ability to remove oxygen while maintaining the ability to activate the NO bond enables Pt to be active for NO decomposition under lean-burn conditions. In addition, it would suggest that alloying Pt with more noble metals such as Au, Ag or Cu might allow one to dramatically suppress oxygen poisoning and hence aid in NO decomposition. 5. Conclusions We combine rst-principle DFT calculations along with kinetic Monte Carlo simulations in order to monitor the molecular transformations involved in the decomposition of NO over Pt(100) and Rh(100) surfaces at ultrahigh vacuum as well as lean-burn conditions. This approach was used to help identify which steps in the mechanism are most critical and how they change as a function of the metal surface as well as the processing conditions. The simulation results reported here were found to be in good agreement with experimental results over both Pt and Rh. Pt(100) is capable of reducing and oxidizing NO to N2 and NO2 under lean-burn conditions while Rh(100) is not. The primary difference between the two surfaces is that Rh is strongly poisoned by oxygen whereas Pt only undergoes a modest decrease in activity due to poisoning. We kindly acknowledge NovoDynamics Inc. and the Ford Motor Company for their support of this work. We would also like to thank Professor Enrique Iglesia (Berkeley) and Dr Randy Cortright (Wisconsin) for invaluable discussions. We would also like to thank Dr Andreas Eichler (Vienna) and Professor Jurgen Hafner (Vienna) for their collaboration, sharing of their work on NO decomposition and use of their VASP code. References
[1] KASPAR, J., FORNASIERO, P., and HICKEY, N., 2003, Catal. Today, 77, 419. [2] BURCH, R., BREEN, J. P., and MEUNIER, F. C., 2002, Appl. Catal. B, 39, 283. [3] BURCH, R., and WATLING, T. C., 1997, Appl. Catal. B, 11, 207. [4] BURCH, R., and WATLING, T. C., 1997, J. Catal., 169, 45. [5] BURCH, R., and MILLINGTON, P. J., 1995, Catal. Today, 26, 185.

total surface coverage drops from 0.8 to about 0.6 ML. The average coverages reported in gure 4 (a) appear to agree well with the experimental/model results reported by Olsson et al. [9], which are also shown in gure 4 for comparison. The results from Olsson et al. were obtained by tting a microkinetic model to known experimental results. The comparison of the oxygen coverages found in the simulation reported here and those of Olsson et al. are similar but shifted by approximately 0.3 ML. This is simply due to the fact that the results by Olsson et al. were normalized to unity. The oxygen coverage was therefore determined by subtracting all of the other coverages from the value of one [9] and not explicitly analysed for. 4.2.2. Comparison of Rh(100) with Pt(100) The kinetics for NO reduction over Rh(100) under the same lean-burn conditions was also examined here by performing simulations over a range of temperatures between 573 to 773 K. Simulation results show there is no activity for either N2 or NO2 production over the Rh(100) surface. Figure 4 (b) presents the surface coverages found in the simulation as a function of temperature. The results clearly show that the oxygen surface coverage remains constant at 0.6 ML whereas the NO coverage decreases from 0.18 to 0.09 ML when the temperature is increased. The lack of nitrogen formation is the direct result of oxygen poisoning. Atomic oxygen is strongly bound to the surface at 396.0 kJ mol1. Our TPR results showed that O2 desorption does not occur until 12001400 K. Oxygen therefore blocks sites necessary to activate NO and ultimately form N2. The few vacant surface sites that exist can adsorb NO. Adsorbed NO, however, is unlikely to dissociate due to the lack of free surface sites. It can react though with atomic oxygen to form adsorbed NO2. At higher temperatures, NO2 is limited by equilibrium and quickly dissociates back to form NO and surface oxygen as NO2 is more strongly bound to Rh than Pt. It is, therefore, much more likely to remain bound to Rh and subsequently dissociate back to form NO and oxygen than it is to desorb. Parapolymerou and Schmidt studied the decomposition kinetics of NO over Pt and Rh along with the inhibition effects due to O2 in the temperature range between 600 and 1800 K [42]. They found that NO decomposition is more strongly inhibited by oxygen on Rh than it is on Pt. At 1300 K, with 0.2 torr of NO and 0.02 torr of O2, the rate is inhibited by a factor of 11 on Rh whereas on Pt it is only inhibited by a factor of 2.3. Our simulations were performed at even higher ratios of inlet oxygen (PO2 =PNO % 15). It is therefore expected that oxygen inhibition over Rh would be stronger. The barrier for NO2 dissociation is essentially zero while the barrier

First-principles-based kinetic Monte Carlo simulation of nitric oxide decomposition


[6] MACLEOD, N., and LAMBERT, R. M., 2002, Appl. Catal. B, 35, 269. [7] SEDLMAIR, C., SESHAN, K., JENTYS, A., and LERCHER, J. A., 2002, Catal. Today, 75, 413. [8] SEDLMAIR, C., SESHAN, K., JENTYS, A., and LERCHER, J. A., 2003, J. Catal., 214, 308. [9] OLSSON, L., WESTERBERG, B., PERSSON, H., FRIDELL, E., SKOGLUNDH, M., and ANDERSSON, B., 1999, J. phys. Chem. B, 103, 10433. [10] OLSSON, L., PERSSON, H., FRIDELL, E., SKOGLUNDH, M., and ANDERSSON, B., 2001, J. phys. Chem. B, 105, 6895. [11] OLSSON, L., and FRIDELL, E., 2002, J. Catal., 210, 340. [12] BOGICEVIC, A., and HASS, K. C., 2002, Surf. Sci., 506, L237. [13] EICHLER, A., and HAFNER, J., 2001, J. Catal., 204, 118. [14] EICHLER, A., and HAFNER, J., 2001, Chem. Phys. Lett., 343, 383. [15] LOFFREDA, D., DELBECQ, F., SIMON, D., and SAUTET, P., 2001, J. chem. Phys., 115, 8101. [16] ZHDANOV, V. P., 1991, Elementary Physicochemical Processes on Solid Surfaces (New York: Plenum Press). [17] MEI, D. H., HANSEN, E. W., and NEUROCK, M., 2003, J. phys. Chem. B, 107, 798. [18] HANSEN, E., and NEUROCK, M., 2001, J. phys. Chem. B, 105, 9218. [19] HANSEN, E. W., and NEUROCK, M., 2000, J. Catal., 196, 241. [20] HANSEN, E. W., and NEUROCK, M., 1999, Chem. Eng. Sci., 54, 3411. [21] HANSEN, E., and NEUROCK, M., 1999, Surf. Sci., 441, 410. [22] HANSEN, E. W., and NEUROCK, M., 2000, Surf. Sci., 464, 91. [23] NEUROCK, M., and MEI, D. H., 2002, Topics Catal., 20, 5. [24] BROWN, W. A., and KING, D. A., 2000, J. phys. Chem. B, 104, 2578. [25] YEO, Y. Y., VATTUONE, L., and KING, D. A., 1996, J. chem. Phys., 104, 3810. [26] KRESSE, G., and FURTHMULLER, J., 1996, Phys. Rev. B, 54, 11169. [27] KRESSE, G., and HAFNER, J., 1994, J. phys. condens. Matter, 6, 8245.

369

[28] PERDEW, J. P., CHEVARY, J. A., VOSKO, S. H., JACKSON, K. A., PEDERSON, M. R., SINGH, D. J., and FIOLHAIS, C., 1992, Phys. Rev. B, 46, 6671. [29] MILLS, G., JONSSON, H., and SCHENTER, G. K., 1995, Surf. Sci., 324, 305. [30] HOPSTAKEN, M. J. P., and NIEMANTSVERDRIET, J. W., 2000, J. phys. Chem. B, 104, 3058. [31] HALGREN, T. A., 1996, J. comput. Chem., 17, 520. [32] DUMESIC, J. A., RUDD, D. F., APARICIO, L. M., REKOSKE, J. E., and TREVINO, A. A., 1993, The Microkinetics of Heterogeneous Catalysis (Washington, DC: American Chemical Society), pp. 3942. [33] CAMPBELL, C. T., 1985, Surf. Sci., 157, 43. [34] RIENKS, E. D. L., BAKKER, J. W., BARALDI, A., CARABINEIRO, S. A. C., LIZZIT, S., WESTSTRATE, C. J., and NIEUWENHUYS, B. E., 2002, Surf. Sci., 516, 109. [35] GARDNER, P., TUSHAUS, M., MARTIN, R., and BRADSHAW, A. M., 1990, Surf. Sci., 240, 112. [36] GOHNDRONE, J. M., and MASEL, R. I., 1989, Surf. Sci., 209, 44. [37] PANJA, C., and KOEL, B. E., 2000, J. phys. Chem. A, 104, 2486. [38] BONZEL, H. P., BRODEN, G., and PIRUG, G., 1978, J. Catal., 53, 96. [39] FINK, T., DATH, J. P., BASSETT, M. R., IMBIHL, R., and ERTL, G., 1991, Surf. Sci., 245, 96. [40] BANHOLZER, W. F., GOHNDRONE, J. M., HATZIKOS, G. H., LANG, J. F., MASEL, R. I., PARK, Y. O., and STOLT, K., 1985, J. vac. Sci. Technol. A, 3, 1559. [41] PIRUG, G., BONZEL, H. P., HOPSTER, H., and IBACH, H., 1979, J. chem. Phys., 71, 593. [42] RIENKS, E. D. L., BAKKER, J. W., BARALDI, A., CARABINEIRO, S. A. C., LIZZIT, S., WESTSTRATE, C. J., and NIEUWENHUYS, B. E., 2003, Surf. Sci., 532, 120. [43] HO, P., and WHITE, J. M., 1984, Surf. Sci., 137, 103. [44] AMIRNAZMI, A., BENSON, J. E., and BOUDART, M., 1973, J. Catal., 30, 55. [45] AMIRNAZMI, A., and BOUDART, M., 1975, J. Catal., 39, 383. [46] PAPAPOLYMEROU, G. A., and SCHMIDT, L. D., 1985, Langmuir, 1, 488.

Das könnte Ihnen auch gefallen