Sunteți pe pagina 1din 19

Surface Science 544 (2003) 5–23

www.elsevier.com/locate/susc

A microkinetic model of the methanol oxidation over silver


A. Andreasen 1, H. Lynggaard, C. Stegelmann, P. Stoltze *

Department of Chemistry and Applied Engineering Science, Aalborg University, Niels Bohrs Vej 8, 6700 Esbjerg, Denmark
Received 14 May 2003; accepted for publication 5 August 2003

Abstract
A simple microkinetic model for the oxidation of methanol on silver based on surface science studies at UHV and
low temperatures has been formulated. The reaction mechanism is a simple Langmuir–Hinshelwood mechanism, with
one type of active oxygen and one route to formaldehyde and carbon dioxide, respectively. The model explains ob-
served reaction orders, selectivity, apparent activation enthalpies and the choice of industrial reaction conditions. More
interesting the model disproves the notion that the mechanism deduced from surface science in UHV cannot be
responsible for formaldehyde synthesis at industrial steady-state conditions. The present work therefore seriously
questions the prevailing models of formaldehyde synthesis in the literature. One of the reasons for this controversy is
that many of the models in the literature are derived from transient experiments exhibiting dynamic effects that are not
present at steady state under industrial conditions.
 2003 Elsevier B.V. All rights reserved.

Keywords: Computer simulations; Models of surface kinetics; Equilibrium thermodynamics and statistical mechanics; Oxidation;
Catalysis; Silver; Alcohols; Oxygen

1. Introduction more water than hydrogen is produced [1,9].


Formaldehyde is only formed in the presence of
The partial oxidation of methanol to formal- oxygen [2]. Steam is added to increase selectivity
dehyde is an important industrial process due to and heat transport [1,3,4]. Traditionally the overall
the versatility of formaldehyde as an intermediate process is regarded as two parallel reactions: an
in chemical synthesis [1]. BASFÕs silver based oxidation (Eq. (1)) and a dehydrogenation (Eq.
process is carried out at 900 K and atmospheric (2)) [3–5].
pressure, the feed consist of a fuel-rich mixture of
methanol and air. At typical reaction conditions 1
CH3 OH þ O2
H2 CO þ H2 O ð1Þ
the selectivity is approximately 90% and the con- 2
version of oxygen approach 100%, and slightly
CH3 OH
H2 CO þ H2 ð2Þ

*
It has been proposed [1,3,4,6] that the selectivity
Corresponding author. Tel.: +45-79-12-76-63 (Office); fax:
+45-75-45-36-43 (Department).
towards formaldehyde is limited by the following
E-mail address: stoltze@aue.auc.dk (P. Stoltze). reactions:
1
Present address: Department of Materials Research, Risoe
National Laboratory, DK-4000 Roskilde, Denmark. 2CH3 OH þ 3O2
4H2 O þ 2CO2 ð3Þ

0039-6028/$ - see front matter  2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.susc.2003.08.007
6 A. Andreasen et al. / Surface Science 544 (2003) 5–23

H2 CO þ O2
CO2 þ H2 O ð4Þ Langmuir–Hinshelwood kinetic expression based
on adsorption on top of an oxide layer was pro-
H2 CO
CO þ H2 ð5Þ posed. The selectivity was explained by total oxi-
dation of both methanol and formaldehyde.
The formation of carbon dioxide is favored by low During the last decade Schl€ ogl and coworkers
temperatures [7–9]. The formation of carbon have studied the oxygen–silver system at high
monoxide is favored by high temperatures temperatures (500–1000 K) and atmospheric pres-
(T > 900 K) [8], and is viewed as a pyrolytic gas sures in great detail (TPD, TPR, ISS, XPS, UPS,
phase reaction [4,6]. in situ-XRD, STM, SEM, RHEED, etc.) [22–27].
Despite the fact that the formaldehyde synthesis The authors have identified three different atomic
has existed for more than a century and a sub- oxygen species above room temperature, denoted
stantial research effort has been devoted to the Oa , Ob , and Oc which desorbs in UHV at 600, 600–
reaction, the mechanism remains controversial. 850, and 900 K, respectively. Oa (XPS (O 1s) ¼
Based on UHV studies on Ag(1 1 0) at 200–300 528.2 eV) is formed by dissociation of molecular
K Wachs and Madix [10] proposed a simple oxygen and is the well known chemisorbed surface
mechanism for methanol oxidation with only one bound oxygen, also termed nucleophilic or ionic
kind of active oxygen. According to this mecha- oxygen in the literature [28–31]. Oa is the active
nism the oxidation of methanol to formaldehyde oxygen species in the mechanism proposed by
goes through a methoxy intermediate. The meth- Wachs and Madix [10]. Ob (XPS (O 1s) ¼ 530.3 eV)
oxy intermediate is formed by a reaction between [26] is formed by dissolution of Oa in the bulk.
methanol and atomic surface oxygen and decom- This formation is activated and occurs via inter-
pose to formaldehyde and hydrogen. The formal- stitial diffusion at a temperature above 450 K
dehyde may be oxidized to carbon dioxide through [23,27]. Oc (XPS (O 1s) ¼ 529.0 eV) is embedded in
a formate intermediate i.e. a consecutive reaction the uppermost layer of silver and is formed by the
path (Eq. (4)). Wachs and Madix results shows no segregation of Ob from the bulk to the surface via
indications of a direct reaction pathway from interstitial diffusion above 580 K [23]. This for-
methanol to carbon dioxide i.e. a parallel reaction mation only occurs in the terminating closed
pathway (Eq. (3)) as proposed by some investiga- packed surface planes (Ag(1 1 1)) and leads to a
tors [8,11]. The methoxy and formate intermedi- pronounced reconstruction and morphological
ates have been identified and studied with in situ changes in silver catalysts [27]. Water induces
Raman, HREELS, XPS, UPS, and TPR on both morphological changes presumable by formation
single crystals and polycrystalline silver catalysts of subsurface OH by reaction with Oc [32–34]. Due
[10,12–17]. Surface science and density functional to reaction-induced restructuring of silver catalysts
theory (DFT) calculations indicates that the mech- and pronounced hysteresis Schl€ ogl et al. concluded
anistic scheme of Wachs and Madix is general for that formaldehyde synthesis is extremely structure-
the oxidation of alcohols on silver and other metal sensitive [8]. However steady-state kinetics has not
surfaces [12,18–20]. been measured on different surface facets or par-
Bhattacharyya et al. [2] proposed a Mars–van ticle sizes (1–5 nm), hence no conclusion on the
Krevelen mechanism [21] from differential reactor structure sensitivity can be made according to the
data at 264–290 C. They observe positive reaction definition of structure sensitivity by Sajkowski and
orders with respect to both methanol and oxygen Boudard [35]. The conclusion of the measurements
and an inhibiting effect caused by water. Robb and is a strong dependence on the preconditioning of
Harriott [11] have performed kinetic investigations the catalyst in transient experiments. This is con-
in a differential reactor at 420 C. The authors sistent with the work of Wachs and coworkers [36]
observe approximately zero order with respect to who showed that preoxidized silver surfaces ini-
oxygen, 0.8 order with respect to methanol and tially results in more active catalysts, but after sev-
decreasing selectivity with increasing conversion. eral hours the same steady-state is reached with or
On the basis of these experiments a modified without the preoxidation. Schl€ ogl and coworkers
A. Andreasen et al. / Surface Science 544 (2003) 5–23 7

also claim that Oa due to its desorption at 600 K aldehyde was significantly longer in these low
cannot be the dominant surface species involved in temperature experiments. Assuming CO2 is formed
partial oxidation of methanol at 900 K [26]. In- by a consecutive reaction of formaldehyde a lower
stead they suggest that Oc is active at industrial selectivity is expected due to the longer lifetime of
conditions. However this conjecture is only valid formaldehyde. Decreasing the coverage of Oa led
in UHV and Oa could easily be the active species at to a significant increase in selectivity which seems
industrial conditions. Indeed it will be shown in logical since more oxygen is needed to obtain total
this work that Oa is present at industrial condi- oxidation products. Furthermore Oa also seems to
tions and most likely the active oxygen at all re- react more rapidly and more completely with
action conditions. This is in agreement with the methanol than Oc and besides that the experiments
work of Wachs and coworkers [36] who showed were transient and conducted without oxygen at-
that only one type of atomic surface oxygen (TPD mosphere. The experiments are therefore not
peak at 600 K) is active in the formaldehyde conclusive. It is very likely that the formation of
synthesis. Oc is too slow to keep up with Oa at steady-state
In recent investigations Nagy et al. [8] and Nagy conditions. A slow formation rate of Oc (slow
and Mestl [9] observe an increasing selectivity to- diffusion of Ob and Oc [27]) and hence a slow de-
wards formaldehyde with increasing temperatures. sorption rate due to microscopic reversibility
At low temperatures the reaction is inactive; it seems to be the most reasonable explanation for
ignites above 450 K. These observations are con- the high thermal stability of Oc . Schl€ ogl and co-
sistent with the investigations of Gavrilin and workers have shown that the amount of Oc de-
Popov [7], Leffert et al. [37], Qian et al. [38] and creased drastically in the presence of methanol
Obraztsov et al. [39]. With this in mind, Gavrilin even in oxygen rich mixtures [26].
and Popov [7] and Nagy et al. [8] suggest different It is evident that there is a conflict between the
kinetics at high and low temperatures. Nagy et al. proposed mechanism deduced from UHV investi-
also claim that different kinds of oxygen are re- gations and the mechanisms deduced from studies
quired in order to explain their experiments [8]. At near industrial conditions. There can be two rea-
low temperatures Oa is active in oxi-dehydroge- sons for this discrepancy. Either the reactions
nation of methanol towards formaldehyde. How- proceeding at room temperature and UHV are
ever due to the strong nucleophility of Oa this different from the reactions at industrial conditions
reaction tends to go to complete oxidation prod- or the interpretation of the complex in situ ex-
ucts. At higher temperatures Oc is believed to periments at industrial conditions is flawed. We
catalyze the selective dehydrogenation of methanol believe the latter is the case because these experi-
to formaldehyde. Thus explaining increasing se- ments includes transient effects that are not present
lectivity with temperature. However the experi- at industrial steady-state conditions. In this work
ments of Nagy et al. and Gavrilin and Popov was we will show that it is indeed possible to interpret
carried out with 100% oxygen consumption and kinetic experiments performed at elevated pres-
kinetics can therefore not be deduced from these sures and temperatures with a simple microkinetic
experiments yet alone reaction mechanisms. model consistent with the mechanism derived by
In our view the strongest evidence for the Wachs and Madix [10] and various surface science
participation of Oc in formaldehyde synthesis experiments. Hence it is not necessary to invoke Oc
was recently published by Muhler and coworkers in order to explain formaldehyde synthesis. This
[40]. By using a temporal-analysis-of-products ap- does not prove that Oc is not participating in the
proach (TAP) Muhler et al. showed that Oc does reaction, but the effect is probably insignificant in
react with methanol to form formaldehyde with steady-state kinetics.
high selectivity. Oa also formed formaldehyde but The approach of microkinetic modeling has
with a lower selectivity. However the experiments successfully been applied for a number of im-
with Oa was conducted at lower temperatures than portant industrial reactions [41–50], which proves
the experiments with Oc and the lifetime of form- a close connection between surface science and
8 A. Andreasen et al. / Surface Science 544 (2003) 5–23

industrial catalysis. An important feature of a description is used for all gas phase molecules and
microkinetic model is that thermodynamic and adsorbates, giving a correct description of the de-
kinetic parameters are physical meaningful and grees of freedom for each species. Furthermore
consistent with experiment and/or theoretical statistical thermodynamics ensures a correct de-
methods such as DFT and transition state theory scription of the gas phase thermodynamics within
(TST). From a microkinetic model it is possible to the ideal gas approximation. The pivot in statisti-
estimate surface coverages, reaction orders and cal mechanics is the partition function from which
activation enthalpy during reaction conditions. all thermodynamic information can be extracted
Hence the applicability of the model is not re- e.g. the equilibrium constants for each step in the
stricted to a particular set of conditions, but can be mechanism. Table 1 summarize the important
used under various conditions where simplified formulas used in the thermodynamic modeling.
models e.g. Power-Law expressions may break
down [44]. In short a microkinetic model should be 2.1. Reaction mechanism
able to bridge the pressure-, temperature-, struc-
ture-, and reactor gaps between experimental and Our model is based on the Langmuir–Hinshel-
theoretical surface science and industrial catalysis. wood mechanism presented in Table 2. All the
To the best of our knowledge this is the first elementary steps in Table 2 have have been ex-
time microkinetic modeling has been applied to the tracted from the UHV work of Wachs and Madix
partial oxidation of methanol to formaldehyde on on Ag(1 1 0) [10]. It should be emphasized that O
silver catalysts. participating in the mechanism in Table 2 corre-
sponds to the surface atomic oxygen denoted Oa
by Schl€ogl and coworkers. The reaction mecha-
2. Methods nism has been kept as simple as possible by ig-
noring known elementary reactions involving
The starting point of a microkinetic model is a carbonate formation [51–53], OH formation [10],
detailed reaction mechanism. The principle of HCOOH formation [38] and oxidative decompo-
microscopic reversibility is applied to each ele- sition of formate [54]. These elementary reactions
mentary step, and the kinetics is described by has been ignored for several reasons. First of all
Arrhenius expressions. A statistical mechanical the purpose of this work is not to develop a de-

Table 1
Statistical mechanical and thermodynamic functions applied in the microkinetic model
Type Partition function Enthalpy
  0   1
Vibration mj hc mj hc
exp exp
Q 2k T P B1 2k T C
zvib ¼ j B  Hvib ¼ jB
@2 hcmj þ B C
mj hc mj hc A
1  exp 1  exp
kB T kB T
 3
2pmkB T 2 kB T 5
Translation ztrans ¼ Htrans ¼ kB T
h2 p 2
kB T
Rotation (2D) zrot ¼ Hrot ¼ kB T
rhcB
 3
1 kB T 2  p 12 3
Rotation (3D) zrot ¼ Hrot ¼ kB T
r hc ABC 2
 
Eg
Electronic zelec ¼ exp Helec ¼ Eg
kB T
Total z ¼ zvib ztrans zrot zelec H ¼ Hvib þ Htrans þ Hrot þ Helec
Eg is the ground state energy, h is PlanckÕs constant, kB is BoltzmannÕs constant, m is the molecular mass, p is the reference pressure
and c is the speed of light in vacuum. A, B and C are rotational constants, r is the symmetry number, mj are the vibrational frequencies.
A. Andreasen et al. / Surface Science 544 (2003) 5–23 9

Table 2 limiting on copper [59]. Pepley et al. have pro-


Reaction mechanism used for the microkinetic model posed a kinetic model for methanol-steam re-
CH3 OH(g) + 
CH3 OH (step 1) forming on copper and concluded that in order to
O2 (g) + 
O2 (step 2) explain kinetic data with their model the decom-
O2 + 
2O (step 3)
2CH3 OH + O
2CH3 O + H2 O (step 4)
position of methoxy must be rate limiting [60].
CH3 O + 
H2 CO + H (slow) (step 5) DFT calculations of methoxy decomposition on
H2 CO
H2 CO(g) +  (step 6) copper confirm a high activation barrier for this
2H
H2 (g) + 2 (step 7) process. It is therefore reasonable to assume that
H2 O
H2 O(g) +  (step 8) methoxy decomposition is a slow process on silver.
H2 CO + O
H + HCOO (step 9)
HCOO + 
H + CO2 (slow) (step 10)
TPR studies of the decomposition of formic acid,
CO2
CO2 (g) +  (step 11) formaldehyde, methanol and methyl formate have
The  signifies a surface site and X  is an adsorbed specie.
shown that CO2 and H2 desorbs simultaneously
around 400 K [10,17,61,62]. The decompositions
of all the above species are believed to go through
tailed model in all respects but to demonstrate that the same formate intermediate [10,17]. This sug-
the reaction mechanism established at UHV can gests that the further oxidation of formaldehyde to
explain kinetic experiments near industrial condi- carbon dioxide is rate limited by the decomposi-
tions. Secondly, the experimental knowledge of the tion of formate (step 10). This conclusion is con-
ignored elementary reactions is too limited to ex- sistent with kinetic measurements of HCOOH
tract reliable kinetic and thermodynamic para- decomposition on different crystal facets of silver
meters. Thirdly, by ignoring these elementary [63–66]. Therefore its assumed in the following
reactions the numerics of the model is greatly that the decomposition of methoxy and formate is
simplified and an analytical solution can be ob- rate limiting while all the other elementary reac-
tained. The consequences of excluding these ele- tions are equilibrated. This assumption will be
mentary reactions will be discussed later in this justified later when the model successfully explain
section. But for now it should be stressed that in- kinetic data. It might seem strange that the uptake
cluding these elementary reactions will not change of oxygen is not rate limiting as its dissociative
the results and conclusions of this work. The sticking probability is very low (107 –105 ) [67,68].
model would merely contain more parameters and However the kinetic data of Robb and Harriott
maybe slightly different parameter values for those [11] indicates a zero order reaction in oxygen,
fitted to kinetic experiments. hence the uptake of oxygen cannot be rate limit-
An important feature of the mechanism by ing. This can be explained by the high oxygen
Wachs and Madix is that the formaldehyde syn- pressure used in kinetic experiments which com-
thesis cannot be divided into a dehydrogenation pensates for the low sticking.
and oxidation. Furthermore CO2 is only formed The following possible elementary reactions
by the further oxidation of formaldehyde through involving OH formation has been excluded from
a formate intermediate i.e. direct methanol com- the model:
bustion (Eq. (3)) is absent.
Wachs and Madix observe simultaneous de- O þ H
OHþ ð6Þ
sorption of different products in TPR experiments
2OH
H2 O þ O ð7Þ
when methanol is adsorbed on preoxidized
Ag(1 1 0) [10] and this desorption occur at higher
OH þ H
H2 Oþ ð8Þ
temperatures than the characteristic desorption
spectra of these products [55–58]. This indicates Therefore the simplified model does not describe
that the products evolves from the same rate lim- the selectivity towards water perfectly nor the
iting step namely the decomposition of the meth- existence of hydroxy species. At high selectivity
oxy intermediate (step 5). It has been shown that toward formaldehyde the mechanism predicts
the decomposition of methoxy (step 5) is rate the production of a 1:1 H2 :H2 O ratio. Including
10 A. Andreasen et al. / Surface Science 544 (2003) 5–23

Eqs. (6)–(8) will increase the formation of water by Surface science experiments have indicated that
decreasing the hydrogen production. Industrially the decomposition of formate could be promoted
only slightly more water than hydrogen is pro- by oxygen [54]:
duced which justifies the neglection of these reac- HCOO þ O
CO2 þ OH ð10Þ
tions at industrial conditions [1,3]. However at
other reaction conditions especially transient ki- This formate decomposition pathway could be
netic experiments this approximation is poor as significant at low temperatures where a large
will become apparent in Section 3. Furthermore coverage of O exists but will be of minor impor-
the formation of OH could serve as an oxygen tance at industrial conditions.
scavenger. Since the stability of OH is low (de- Applying the quasi-equilibrium approximation
sorbs at about room temperature) [57] and the [69] to the Langmuir–Hinshelwood mechanism in
oxygen uptake is assumed to be equilibrated this Table 2 assuming steps 5 and 10 as rate limiting,
scavenger effect would be negligible for steady- the rate and equilibrium equations can be derived
state kinetics above room temperature. According cf. Table 3.
to Wachs and Madix OH may react in a similar
manner as O [10]. 2.2. Model catalyst
The production of HCOOH and HCOOCH3
observed at some experimental conditions has When possible we have derived the model pa-
been ignored in the model [10,37,38]. These reac- rameters from experiments on Ag(1 1 1) because it
tions seems to be unimportant in steady-state ki-
netics. However Qian et al. [38] found that the
production of HCOOH increased to a significant Table 3
Rate and equilibrium equations for the kinetic model based on
level on an aged catalyst. The reason for this aging
reaction steps 1–11
effect is not clear. pCH3 OH
In this paper there will be no attempts to hCH3 OH ¼ K1 h
p
model the decomposition of formaldehyde to pO2
hO2 ¼ K2  h
carbonmonooxide (Eq. (5)) since this reaction is p
1

important only at very high temperatures [6,8, hO ¼ ðK3 hO2 h Þ2 


1 21
37]. Furthermore it is believed to be a pyrolytic hCH3 O ¼ hCH3 OH K4 hO
hH2 O
gas phase reaction [6] and should therefore not k5
r5 ¼ k5 hCH3 O h  hH2 CO hH
lead to site blockage of any kind on the catalyst. K5
1 pH2 CO
CO desorbs at low temperatures and reacts rap- hH2 CO ¼ h
K6 p
idly with Oa to form CO2 on silver hence CO 1 pH 2
1

should only be present in the absence of oxygen hH ¼ K7 2 2 h


p
[51]. 1 pH2 O
hH2 O ¼
 h
When adsorbed atomic oxygen and carbon di- K8 p 
oxide interacts on a silver surface, carbonate is hH CO hO
hHCOO ¼ K9 2
known to form [51–53]: hH
k10
r10 ¼ k10 hHCOO h 
 hCO2 hH
CO2 þ O
COþ
3 ð9Þ K10
1 pCO2
hCO2 ¼ h
This is considered to be a dead end in the mech- K11 p
anism and the most important effect is stabiliza- h ¼ 1  hCH3 OH  hO2  hO  hCH3 O
hH2 CO  hH  hH2 O  hHCOO  hCO2
tion of carbon dioxide. The carbonate will
compete with the other adsorbates in the con- Ki are the equilibrium constants calculated from the molecular
sumption of active sites, but we consider this effect partition functions of the intermediates, ki are the rate constants
assumed to be of the Arrhenius form, ri are the reaction rates of
to be negligible at high temperatures, and the
the rate limiting steps, hX  is the coverage of species X  , pi is the
phenomena is not included in the further model- partial pressure and p is the thermodynamic reference pres-
ing. sure.
A. Andreasen et al. / Surface Science 544 (2003) 5–23 11

is the thermodynamically stable facet and there- that a simple mechanism deduced from surface
fore supposed to be the most abundant facet on science is able to explain experiments at industrial
the industrial catalyst. Nevertheless in some cases conditions.
the Ag(1 1 0) surface has been used due to the
absence of appropriate experiments on Ag(1 1 1). 2.3. Model parameters
The density of sites (d) is estimated from oxygen
chemisorption on Ag(1 1 1) [28] which results in a The equilibrium constants appearing in the
saturation coverage of approx. 0.5 ML oxygen microkinetic model can be calculated using statis-
atoms. This, combined with the fact that a tical thermodynamics [72] from parameters for gas
Ag(1 1 1) facet has 1.38 · 1015 surface atoms per phase molecules and adsorbates. The central pa-
cm2 gives a site density of 6.9 · 1018 sites/m2 . rameters are vibrational frequencies and ground
In general, microkinetic modeling is not sensi- state energies. All parameters for gas phase mole-
tive to the density or type of active sites as long as cules can be extracted from the NIST database [73]
the experiments on the model catalyst used to de- or equivalent. For the adsorbates vibrational fre-
rive the model parameters reflects what is hap- quencies are determined from spectroscopic mea-
pening on an industrial catalyst. This is due to an surements e.g. EELS, IR, Raman. Ground state
inverse proportionality between the site density energies can be fitted from TPD experiments and
and the prefactors of all the elementary reactions. measurements of sticking coefficients. The method
The best way to establish if a model catalyst used for parameter estimation will be explained
reflects an industrial catalyst is to compare steady- below and the results are summarized in Tables 4–
state kinetics of the two systems. Unfortunately 6. For more details on the method we refer to
such measurements has not been performed on other papers [74,75].
Ag(1 1 1) for methanol oxidation. Furthermore the A first order desorption process in UHV for a
structure sensitivity of the reaction is unknown. generic gas phase molecule A is described by the
Formate decomposition has been studied on dif- following rate equation:
ferent Ag facets showing a significant change in
activation barrier with surface structure [63–66]. dhA k
¼  hA  ð11Þ
On the other hand the effect on activities was dt K
canceled by a large compensation effect. Silver
where k is the rate constant of adsorption, K is the
reconstruct dramatically in the presence of oxygen
equilibrium constant and hA is the coverage of A.
and/or water during the formation of Oc and
The equilibrium constant can be calculated from
subsurface OH leading to faceting [27,32–34]. The
the partition function cf. Table 1:
reconstruction seems to be decreased in the pres-
ence of methanol. It is possible that a more accu- zA
K¼ ð12Þ
rate model could be produced by including a zA
dependence of reaction conditions on the site
density as in the case of methanol synthesis on where zA is the partition function of the gas phase
copper [70,71]. However, as a first approximation and zA is the partition function of the adsorbate.
we assume that the site density is constant. As it The equilibrium constants for adsorption and
will be shown in Section 2.3 it has been necessary surface reactions can be calculated in a similar
to fit some parameters to steady-state kinetics. fashion.
These parameters may therefore not reflect the The rate constant k is found by equating the
values on a particular facet but may be some kind initial adsorption rate with the initial sticking rate:
of average value of all the facets/steps/defects r0 p
present at industrial conditions. It should be em- k ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð13Þ
d 2pmkB T
phasized that the purpose of this work is not to
show which facet is the active one or that the re- where r0 is the initial sticking coefficient, p is the
action is structure insensitive but to demonstrate thermodynamic reference pressure, m is the mass
12 A. Andreasen et al. / Surface Science 544 (2003) 5–23

Table 4
Model parameters for gas phase molecules to the statistical thermodynamical treatment
Species Parameters
H2 B ¼ 60:848 cm1 , r ¼ 2, m1 ¼ 4405:3 cm1 , H ¼ 0 @ T ¼ 298 K [75]
O2 B ¼ 1:45 cm1 , r ¼ 2, m1 ¼ 1580 cm1 , H ¼ 0 @ T ¼ 298 K [73]

H2 O A ¼ 27:8847 cm1 , B ¼ 14:51181 cm1 , C ¼ 9:2806 cm1 , r ¼ 2, m1 ¼ 1594:7 cm1 , m2 ¼ 3651:1 cm1 ,
m3 ¼ 3755:9 cm1 , H ¼ 241:818 kJ/mol @ T ¼ 298 K [73].
CH3 OH A ¼ 4:2554 cm1 , B ¼ 0:823 cm1 , C ¼ 0:7928 cm1 , r ¼ 1, m1 ¼ 270 cm1 , m2 ¼ 1033 cm1 , m3 ¼ 1060 cm1 ,
m4 ¼ 1165 cm1 , m5 ¼ 1345 cm1 , m6 ¼ 1477ð2Þ cm1 , m7 ¼ 1455 cm1 , m8 ¼ 2844 cm1 , m9 ¼ 2960 cm1 ,
m10 ¼ 3000 cm1 , m11 ¼ 3681 cm1 , H ¼ 201:2 kJ/mol @ T ¼ 298 K [45,73].
H2 CO A ¼ 9:074 cm1 , B ¼ 1:2899 cm1 , C ¼ 1:129 cm1 , r ¼ 2, m1 ¼ 1163:5 cm1 , m2 ¼ 1247:4 cm1 ,
m3 ¼ 1500:6 cm1 , m4 ¼ 1746:07 cm1 , m5 ¼ 2766:4 cm1 , m6 ¼ 2843:4 cm1 , H ¼ 108:57 kJ/mol @ 298 K [73].
CO2 B ¼ 0:39038 cm1 , r ¼ 2, m1 ¼ 667:3ð2Þ cm1 , m2 ¼ 1384:26 cm1 , m3 ¼ 2349:49 cm1 ,
H ¼ 393:15 kJ/mol @ T ¼ 298 K [73].
A, B and C are the rotational constants, r is the symmetry number, mi are the vibrational frequencies and the degeneracy of a frequency
is enclosed in parentheses. H is the enthalpy of formation.

Table 5
Model parameters for adsorbates to the statistical thermodynamical treatment
Species Parameters
H m? ¼ 1121:0 cm1 , mk ¼ 927:5ð2Þ cm1 , H ¼ 25 kJ/mol [75].
O m1 ¼ 350 cm1 , m2 ¼ 508ð2Þ cm1 , H ¼ 63 kJ/mol @ T ¼ 298 K [81].
O2 m1 ¼ 50ð2Þ cm1 , m2 ¼ 300ð2Þ cm1 , m3 ¼ 675 cm1 , m4 ¼ 220 cm1 ,
H ¼ 44:5 kJ/mol @ T ¼ 298 K [81].
CH3 OH m? ¼ 29 cm1 , mk ¼ 35:51ð2Þ cm1 , mr ¼ 36:0ð3Þ cm1 , m1 ¼ 750 cm1 , m2 ¼ 820 cm1 ,
m3 ¼ 1030 cm1 , m4 ¼ 1150ð2Þ cm1 , m5 ¼ 1470ð3Þ cm1 , m6 ¼ 2860 cm1 , m7 ¼ 2970ð2Þ cm1 ,
m8 ¼ 3320 cm1 , H ¼ 241:6 kJ/mol @ T ¼ 298:15 K [12,45].
CH3 O m? ¼ 330 cm1 , mk ¼ 36:49ð2Þ cm1 , mr ¼ 360ð3Þ cm1 , m1 ¼ 1040 cm1 , m2 ¼ 1150ð2Þ cm1 ,
m3 ¼ 1450ð3Þ cm1 , m4 ¼ 2840 cm1 , m5 ¼ 2940ð2Þ cm1 , H ¼ 104 kJ/mol @ T ¼ 298:15 K [12,45].
H2 CO m? ¼ 400 cm1 , mk ¼ 37ð2Þ cm1 , mr ¼ 23ð3Þ cm1 , m1 ¼ 1476 cm1 , m2 ¼ 1694 cm1 , m3 ¼ 2839 cm1 ,
m4 ¼ 1218:15 cm1 , m5 ¼ 2839 cm1 , H ¼ 135 kJ/mol @ T ¼ 298:15 K [56,77].
H2 O m? ¼ 27:9 cm1 , mk ¼ 27:9ð2Þ cm1 , mr ¼ 740ð3Þ cm1 , m1 ¼ 1660 cm1 , m2 ¼ 3410ð2Þ cm1 ,
H ¼ 293:2 kJ/mol @ T ¼ 298:15 K [57,75].
CO2 m? ¼ 410 cm1 , mk ¼ 30:6ð2Þ cm1 , mr ¼ 12:71ð3Þ cm1 , m1 ¼ 605 cm1 , m2 ¼ 1365 cm1 , m3 ¼ 2350 cm1 ,
H ¼ 414:607 kJ/mol @ T ¼ 298:15 K [77,90].
HCOO m? ¼ 322 cm1 , mk ¼ 35:51ð2Þ cm1 , mr ¼ 400ð3Þ cm1 , m1 ¼ 770 cm1 , m2 ¼ 1340 cm1 , m3 ¼ 1640 cm1 ,
m4 ¼ 2900 cm1 , m5 ¼ 1050 cm1 , m6 ¼ 1377 cm1 , H ¼ 197 kJ/mol @ T ¼ 298:15 K [13,45,61].
mi are the vibrational frequencies and the degeneracy of a frequency is enclosed in parentheses. m? , mk and mr are the frustrated
translational orthogonal frequency, the frustrated translational parallel frequency and the frustrated rotational frequency, respectively.
H is the enthalpy of formation.

of A, kB is BoltzmannÕs constant and T is the


temperature. Substituting this into Eq. (11) the dhA r0 p
¼  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi hA ð14Þ
following equation is obtained: dt d 2pmkB T K
A. Andreasen et al. / Surface Science 544 (2003) 5–23 13

Table 6 2.3.1. Adsorbed oxygen


Arrhenius parameters for the two slow steps determined from The stability of dissociative adsorbed oxygen on
optimization
silver is controversial even though the oxygen/sil-
Reaction, i Ai Hiz ver system has been studied extensively. On
1
5 10
4.2 · 10 s 60 kJ mol1 Ag(1 1 1) the value referenced to by most people is
10 9.8 · 109 s1 77 kJ mol1
the one obtained by Campbell [67]. Campbell used
A is the prefactor and H z is the activation enthalpy. the first order Redhead equation [78], a desorption
preexponential of 1015 s1 and deduced a disso-
ciative adsorption heat of 170.5 kJ/mol O2 from
Experimental TPD spectra are simulated by in- the TPD of dissociatively adsorbed oxygen. As
tegrating equation (14) using numerical tech- stated by Campbell [67] the Redhead analysis is a
niques. The ground state energy comes into play simplification. The desorption is recombinative
through the equilibrium constant and is opti- (second order) and an attractive interaction exists
mized until the experimental peak temperature is between dissociatively adsorbed oxygen. The at-
reproduced. tractive interaction is observed by island forma-
It seems likely that the sticking coefficient of tion at low coverage below 490 K in UHV and by
methanol, formaldehyde and water is near unity a very narrow TPD peak [30,31,67,68,79,80].
[56–58]. Assuming a sticking coefficient of unity, We have made a microkinetic model of oxygen
implies that the activation barrier for the sticking chemisorption on Ag(1 1 1) that (within experi-
process is zero. These assumptions results in the mental uncertainties) describes molecular and
parameters presented in Table 5 which results in dissociative sticking, and molecular and recomb-
TPD peaks at 167 K [58,76], 106 K [56], and 170 K inative desorption [81]. The attractive interaction
[57] for methanol, formaldehyde and water, re- between dissociatively adsorbed oxygen has been
spectively which is consistent with experimental modeled using the mean-field approximation. This
values. The exact stability of these species are not model resulted in a heat of adsorption of 126 kJ/
critical to the model, cf. Section 4. mol O2 and the attraction was found to be 15 kJ/
The stability of hydrogen is deduced from an mol O2 .
experimental D2 desorption spectra by Wachs In the following we will use the heat of ad-
and Madix [10] and with vibrational parameters sorption in the limit of low coverage assuming that
from Cu [77]. By assuming a desorption expo- the attractive interactions should be ignored on
nential of 1013 and an activation barrier of 55.6 polycrystalline or real silver catalysts due to the
kJ/mol the peak temperature of 228 K is repro- lack of long range order on such catalysts [82] and
duced by simulation. Assuming a non-activated the presence of other adsorbates.
sticking of hydrogen we come up with an H
enthalpy of )27.8 kJ/mol. The sticking of hy- 2.3.2. Methoxy and formate
drogen on silver is probably activated as for Cu The methoxy and formate species are special in
[77], which results in a too high predicted value of the sense that there are no corresponding gas
the enthalpy of H. We have also neglected the phase molecules. The only information available
isotope effect, which also leads to discrepancy for these adsorbates are vibrational frequencies
between the stability of H and D. It turns out and decomposition temperatures in TPR on
that the microkinetic model is rather insensitive Ag(1 1 0). From the TPR experiments it is only
to the choice of H parameters. All parameters for possible to establish the rate constants of methoxy
carbon dioxide are taken from Cu [77], because and formate decomposition. The parameters that
experiments on Ag are insufficient. Both for hy- remains to be determined are ground state energies
drogen and carbon dioxide the stabilities will be and the specific decomposition Arrhenius para-
over predicted, but this is considered only to have meters of methoxy and formate. Fig. 1 shows the
minor importance due to the small coverage of simulation of methoxy and formate decomposition
these species cf. Fig. 7. for different Arrhenius parameters but identical
14 A. Andreasen et al. / Surface Science 544 (2003) 5–23

0.06 Leffert et al. [37], Nagy et al. [8] and Nagy and
10
10

11
Mestl [9] (see Figs. 2–5). The kinetic model is im-
0.05 10
10
12 plemented in a plug-flow reactor model and kinetic
experiments are simulated by integrating the de-
Reaction rate [s ]

13
-1

10
0.04
sign equations with an adaptive step size 4 order
0.03 Runge–Kutta algorithm. The optimization occurs
with a least-squares simulated annealing. The
0.02 calculation ensures that the correct TPR decom-
position temperatures of methoxy and formate are
0.01
reproduced, cf. Fig. 1.
0 These parameters are therefore not deduced
250 300 350 400 450 explicitly from surface science but they reproduce
Temperature [K]
all known surface science experiments and have
Fig. 1. Simulation of TPR experiments for methoxy (low physically realistic values according to DFT and
temperature peak) and formate decomposition (high tempera- TST. Other choices of parameters could give simi-
ture peak) using different sets of Arrhenius parameters. b ¼ 15
lar results.
K/s and initial methoxy and formate coverage is 10%.
The prefactors of methoxy and formate de-
composition is in the region 1010 s1 cf. Table 6.
rate constants at the peak temperatures. Its is clear TST predicts a prefactor of about 1013 s1 for
from the figure that the specific Arrhenius pa- surface reactions varying few orders of magnitude
rameters cannot be establish from such TPR ex- dependent of the nature of the transition state
periments. (mobile/immobile) [41,82]. Experimentally preex-
The ground state energy of methoxy has been ponentials vary from 1010 –1016 s1 [82]. The values
determined by ensuring that the models heat of of the prefactors in the model seems low compared
reaction for decomposition of gas phase methanol to TST, however within the range usually reported
into adsorbed methoxy and H is consistent with in the literature and hence physically realistic.
DFT calculations. The total energy calculation Besides the mobility of the transition state the low
code DACAPO [83] is used for the DFT calcula- preexponentials could be a result of overestimating
tions. A four layer 2 · 2 slab representing Ag(1 1 1),
periodically repeated in a super cell geometry with 0.04
six equivalent layers of vacuum between successive
slabs. The bottom layer is fixed while the other
Methanol Conversion [mol]

0.9 % Methanol
layers are allowed to relax. Ionic cores are de- 0.03 1.8 % Methanol
scribed by ultra-soft pseudo-potentials [84], and 2.8 % Methanol
5.5 % Methanol
the Kohn–Sham one-electron valence states are 8.8 % Methanol
19.1 % Methanol
0.02
expanded in a basis of plane waves with a kinetic Simulations
energy below 25 Ry. The surface Brillouin zone is
sampled at 18 special k points. The exchange-
0.01
correlation energy and potential are described by
the generalized gradient approximation (GGA-
PW91) [85–87]. Methoxy and H is placed in the fcc 0
0 0.05 0.1 0.15
site found to be the best site for these species on Oxygen mole fraction
copper [20]. It was not our intention to conduct a
thorough DFT investigation of this reaction but Fig. 2. Simulation of the experiments by Robb and Harriott at
merely to get an estimate of the methoxy stability 695 K and 1.12 atm. Points are experimental measurements
from Ref. [11] and lines have been calculated with the micro-
on Ag(1 1 1). kinetic model. The figure shows the conversion of methanol as a
The remaining parameters are fitted to the ki- function of the partial pressure of oxygen at different partial
netic measurements of Robb and Harriott [11], pressures of methanol.
A. Andreasen et al. / Surface Science 544 (2003) 5–23 15

1
Simulations
CH2O Set I H2CO
0.5
CO2 Set I CO2
0.8
CH2O Set II
0.4 CO2 Set II
0.6
Yield

Predicted yield
0.3

0.4
0.2

0.1 0.2

0
0 0.2 0.4 0.6 0.8 0
Conversion of CH3OH 0 0.2 0.4 0.6 0.8 1
Measured yield
Fig. 3. Simulation of the experiments by Robb and Harriott at Fig. 5. Simulation of 25 experiments by Leffert et al. [37]
695 K and 1.12 atm. Points are experimental measurements plotted on a predicted versus experimental plot.
from Ref. [11] and lines have been calculated with the micr-
okinetic model. The figure shows the yield of H2 CO and CO2 ,
respectively, as a function of the conversion of methanol. Set I
is data obtained with an aged catalyst of silver particles of
about 120 A on alumina (sintered). Set II is data obtained with values reported on Ag(1 1 1) facets (67 kJ/mol) and
 on alumina.
a fresh catalyst of silver particles of about 55 A Ag(1 1 0) facets (127 kJ/mol) [63].

1 3. Results

0.8 In this section our simulations of the kinetic


Selectivety / Conversion

CH2O
experiments of Robb and Harriott [11], Nagy et al.
0.6 CO2 [8], Nagy and Mestl [9], and Leffert et al. [37] are
CH3OH presented using the parameters obtained in the
CH2O (mes.)
previous section.
0.4 CO2 (mes.)
CH3OH (mes.) Fig. 2 shows the simulation of the total rate of
methanol oxidation as a function of oxygen and
0.2 methanol pressures at 420 C as measured by
Robb and Harriott. Robb and Harriott only re-
0
500 600 700 800 900
ported a relative rate and analyzed their data ac-
Temperature [K] cording to the differential reactor approximation
even though the conversion in some cases reached
Fig. 4. Simulation of the experiments by Nagy and Mestl.
30%. We have chosen to analyze their data ac-
Points are experimental measurements from Ref. [9] and lines
have been calculated with the microkinetic model. The figure cording to the integral reactor approximation and
shows the selectivity towards H2 CO and CO2 , respectively, and converting the relative rates to absolute rates. This
the conversion of methanol as a function of temperature. transformation was done by calibrating the rela-
tive rates to a reported standard measurement and
the site density in the fitting process, inaccuracies approximating the density of sites from measure-
of the estimated vibrational frequencies for meth- ments of particle sizes. However it turns out that
oxy and formate or perhaps a need for larger site the trends deduced from the integral analysis is
ensembles for methoxy and formate decomposi- identical to those obtained in the differential re-
tion. actor analysis by Robb and Harriott. Fig. 2 shows
Interestingly the fitted activation barrier of that our model correctly explains the reaction or-
formate decomposition (77 kJ/mol) is between the ders of oxygen and methanol.
16 A. Andreasen et al. / Surface Science 544 (2003) 5–23

Fig. 3 shows the simulation of Robb and Har- 4. Discussion


riottÕs measurements of yield versus methanol
conversion at 420 C. The simulation clearly cap- In the previous section we have demonstrated
tures the correct trends, i.e. the selectivity towards that a microkinetic model based on a mechanism
formaldehyde decreases with increasing conver- deduced from surface science is able to explain a
sions of methanol. However the experimental se- broad range of kinetic experiments. Even more
lectivity decreases much more rapidly with interesting some of these experiments has been
methanol conversion than predicted by the model. used as proof in the literature for a mechanism
The reason for this discrepancy could be the use of different form the one observed in UHV. In this
alumina as a support in the experiments of Robb section we will proceed with a more detailed
and Harriott. Experimental evidence shows that analysis in order to obtain a more thorough un-
alumina decomposes formaldehyde [14]. This hy- derstanding of the model.
pothesis has been confirmed by simulations in-
cluding formaldehyde decomposition on alumina 4.1. Coverages
in the microkinetic model.
Nagy et al. investigated the influence of tem- As mentioned in Section 1 Schl€ ogl and co-
perature on selectivity by applying a low linear workers concluded that Oa cannot be active at
heating rate (1 K/min) on a steady-state flow re- industrial conditions of 900 K since it desorbs at
actor while running the reaction. The oxygen 600 K in UHV. However in the previous section
conversion was 100% at all conditions except at our microkinetic model explained experiments at
very low temperatures (500 K) [8,9]. Fig. 4 shows high temperatures using parameters of O repro-
the simulation of the experimental conversion of ducing the experimental TPD peak temperature at
methanol and selectivity towards formaldehyde 600 K in UHV. Fig. 6 shows the simulated cov-
and CO2 , respectively. The model clearly captures erage of O as a function of temperature for dif-
the trend that formaldehyde selectivity increases ferent oxygen pressures, which clearly indicates
with temperature and saturates at high tempera- that O can be present at industrial conditions. At
ture. The model also captures the trend that CO2 UHV the coverage of O approach zero at 600 K
selectivity is high at low temperature but decrease while a significant coverage is present at 900 K
with increasing temperature. In order to fit the at oxygen pressures above 1 kPa. Interestingly,
experiment of Nagy et al. the inlet oxygen con- Scheffler and coworkers [88] found that surface
centration was reduced with 15% otherwise total
methanol consumption was achieved at high tem-
peratures which is not observed experimentally. 1
The reason for this problem is the exclusion of 0.9 P = 0 Pa
P = 1 Pa
elementary reactions forming water from O and 0.8 P = 10 Pa
H through OH in the model. Therefore the ox- P = 100 Pa
P = 1000 Pa
0.7
ygen consumption is too low for the model. P = 10000 Pa
0.6 P = 100000 Pa
Leffert et al. [37] investigated the effect of tem-
θO*

perature on methanol oxidation in a similar fash- 0.5

ion as Nagy et al. [8] and Nagy and Mestl [9]. 0.4
Furthermore the authors investigated the effect of 0.3
methanol and oxygen concentration on the yield 0.2
towards formaldehyde and CO2 , respectively. Fig. 0.1
5 shows the predicted yield of formaldehyde and
0
CO2 by the model versus experimental yields. As 400 500 600 700 800 900 1000 1100 1200
above it was necessary to reduce the inlet oxygen Temperature [K]
concentration by approx. 20% to explain the ex- Fig. 6. Simulated coverage of atomic oxygen on silver versus
periments. temperature at different oxygen pressures.
A. Andreasen et al. / Surface Science 544 (2003) 5–23 17

atomic oxygen is stable up to 830 K at atmo- at a gas composition of 32% CH3 OH, 4.5% O2 ,
spheric pressure by using the approach of 11.6% H2 CO, 1.3% CO2 , 7.4% H2 O and 6.5% H2 .
atomistic ab initio (DFT) thermodynamics. Con- This composition corresponds to an average gas
sidering the systematic errors of DFT [88] the re- composition between the inlet and outlet of an
sult of Scheffler and coworkers is very close to the industrial reactor (without additional water in the
result obtained in this paper by the microkinetic feed) and we will refer to this as typical reaction
model. The conclusion of Schl€ ogl et al. that Oa conditions. The figure shows that O is the most
should be absent at 600 K is therefore unfounded. dominating adsorbate. The number of free sites
Oa may be present at industrial conditions due to increases with the decrease of the coverage of O .
its thermal stability, however it could be absent for The coverage of all intermediates are small except
other reasons. For example, if a stable layer of Oc formate that becomes significant at low tempera-
is formed, it could block the sites for Oa uptake. tures.
This might be the case in some of the transient
experiments conducted by Schl€ ogl and coworkers. 4.2. Apparent activation enthalpies
At steady-state industrial reaction conditions
blocking of Oa uptake is unlikely for several rea- The apparent activation enthalpy Ez for steps 5
sons. First of all the formation of Oc is slow and and 10 can be derived analytically from the defi-
activated and can never exceed the rate of Oa up- nition [43]:
take since its formed from Oa . Secondly, the cov-
d lnrþ
erage of Oc decreases significantly in the presence E z ¼ kB T 2 ð15Þ
of methanol and can therefore not block Oa up- dT
take in a reactive atmosphere. From the above where rþ is the forward reaction rate. This gives
discussion it is clear that it is impossible to predict the results shown in Table 7 and it is observed that
the presence of intermediates from TPD tempera- a simple relation between Ez and the coverage of
tures in UHV alone. Furthermore it is very im- surface species exist. Ez is a sum of the activation
portant to note that an intermediate does not enthalpy for the rate limiting step and a weighted
necessarily need a large coverage in order to be average of the desorption enthalpies for the in-
important. termediates. The sum of the activation enthalpy of
Fig. 7 shows the calculated coverages at differ- the rate limiting step is the activation barrier of the
ent temperatures for the microkinetic model. The elementary step itself plus the heat of formation/
coverages of all intermediates has been calculated adsorption of the reactants in the rate limiting
step. The average desorption enthalpy is formed

1 Table 7
* Apparent activation enthalpies Eiz for the oxidation of methanol
0.1 CH3OH* to formaldehyde (5) and further from formaldehyde to carbon
O2 * dioxide (10)
0.01 O*
CH3O* E5z ¼ H5z þ H1 þ 1=2H4 þ 1=2H8 þ 1=4H2 þ 1=4H3
H2O*
0.001 2H1 hCH3 OH  2H2 hO2  ðH2 þ H3 ÞhO
Coverage

H2CO*
H*
ð2H1 þ H4 þ H8 þ 1=2H2 þ 1=2H3 ÞhCH3 O
0.0001 HCOO* þ2H6 hH2 CO þ H7 hH þ 2H8 hH2 O þ 2H11 hCO2
CO2*
ð2H9  2H6 þ H2 þ H3 þ H7 ÞhHCOO
1e-05
z z
E10 ¼ H10 þ H9  H6 þ 1=2H2 þ 1=2H3 þ 1=2H7
1e-06
2H1 hCH3 OH  2H2 hO2  ðH2 þ H3 ÞhO
1e-07 ð2H1 þ H4 þ H8 þ 1=2H2 þ 1=2H3 ÞhCH3 O
400 500 600 700 800 900 þ2H6 hH2 CO þ H7 hH þ 2H8 hH2 O þ 2H11 hCO2
Temperature [K] ð2H9  2H6 þ H2 þ H3 þ H7 ÞhHCOO
Fig. 7. Initial coverages as a function of temperature at 1 atm, Hiz is the activation enthalpy for the two slow steps in the
calculated at typical reaction conditions. mechanism (5 and 10) and Hi is the reaction enthalpy for step i.
18 A. Andreasen et al. / Surface Science 544 (2003) 5–23

by multiplying the coverage of the intermediates increase in free sites cf. Fig. 7. Hence it is easy to
by twice the enthalpy of desorption of the inter- create two free sites for the rate limiting steps at
mediates through equilibrium steps. The factor of high temperatures. The activation enthalpy for the
two enters because the rate limiting step requires oxidation of methanol to formaldehyde is larger
two free sites. Ez is thus the sum of the activation than the activation enthalpy for CO2 formation at
enthalpy of the rate limiting step plus the average all temperatures even though the activation barrier
cost of creating two free sites on the surface. In the of step 5 is smaller than step 10 cf. Table 6. The
literature most people seem to be aware of the fact difference in apparent activation enthalpy of steps
that the apparent activation energy contain con- 5 and 10 is almost constant in the entire temper-
tributions from the activation barrier of rate lim- ature interval and correspond to the difference of
iting elementary step plus adsorption heats of the the sum of the activation enthalpy of the rate
adsorbates participating in the forward rate limi- limiting steps (steps 5 and 10) and the adsorption
ting step. However the coverage dependent terms heats.
are usually ignored in the literature. This is a se- In Fig. 9 calculated apparent activation enthal-
rious error in all systems that proceeds with a pies for the oxidation of methanol to formaldehyde
significant coverage of intermediates on the active are compared with experimentally observed values
sites. Coverage dependent terms are responsible from the work of Obraztsov et al. [39], Gavrilin
for the variations in apparent activation energies and Popov [7] and Bhattacharyya et al. [2]. The
with reaction conditions. apparent activation enthalpies of Qian et al. [38]
The expressions in Table 7 have been used to have been extracted from methanol conversions
calculate apparent activation enthalpies at typical versus temperature. Calculated values are obtained
reaction conditions as a function of temperature. from the differential reactor approximation ac-
Results of these calculations are shown in Fig. 8. cording to Eq. (15) (solid line) and according to
Large variations in the activation enthalpies are the integral reactor approach (dashed line) where
observed, as a consequence of changes in cover- rþ in Eq. (15) is substituted with the conversion of
ages with temperature. This is an illustrative ex-
ample of the difficulty in comparing activation
enthalpies obtained at different reaction condi- 200000
Bhattacharyya et al.
Apparent activation enthalpy [J/mol]

tions. Gavrillin et al.


Both activation enthalpies decrease as a func- 150000
Obraztsov et al.
Qian et al.
tion of temperature, which is primarily due to the Simulated (Differential)
Simulated (Integral)

100000
200000
Apparent activation enthalpy [J/mol]

Methanol to formaldehyde
Formaldehyde to carbondioxide 50000
150000

0
400 500 600 700 800 900
100000
Temperature [K]

Fig. 9. Apparent activation enthalpy for the oxidation of


50000 methanol to formaldehyde calculated at differential and integral
conditions and experimental observations from Refs. [2,7,
38,39]. Apparent activation enthalpies for differential condi-
0 tions are calculated at typical reaction conditions. Apparent
400 500 600 700 800 900
Temperature [K] activation enthalpies for integral conditions are calculated from
conversion of methanol in a plug reactor with a feed compo-
Fig. 8. Apparent activation enthalpies for the oxidation of sition of 9.82% CH3 OH and 14.3% O2 . The amount of catalyst
methanol to formaldehyde and further to carbon dioxide. and flow rate is chosen to give nearly complete conversion of
Calculations has been performed at typical reaction conditions. methanol at 900 K.
A. Andreasen et al. / Surface Science 544 (2003) 5–23 19

methanol. As shown the differential model quali-


1
tatively explains the trend in experimental obser-
CH3OH
vations i.e. a decrease in activation enthalpy as a O2
function of temperature. However the values seem

Reaction order α i
H2CO
0.5
to be over estimated. The discrepancy between CO2
O2
calculated and measured values can be accounted
for as deviations from differential conditions i.e. a 0
substantial conversion in experimental observa-
tions. This is clearly shown by the activation en-
thalpies calculated for integral conditions giving a -0.5

much better description of the experimental ob-


servations. 400 500 600 700 800 900
Temperature [K]
In this context it is very important to mention,
that none of the experimental activation enthalpies Fig. 10. Calculated reaction orders for the oxidation of meth-
presented in Fig. 9 have been used in the optimi- anol to formaldehyde as a function of temperature at typical
reaction conditions.
zation of parameters mentioned earlier.

4.3. Reaction orders methanol. Experimentally an inhibiting effect from


water [2] is observed. This observation is repro-
The reaction orders can be derived from the duced by the model, cf. Eq. (21). Calculated re-
following definition [43]. action orders cf. Eqs. (17)–(21) at typical reaction
conditions are plotted in Fig. 10 as a function of
d lnðrþ Þ temperature. The reaction order of methanol,
ai ¼   ð16Þ
d ln ppi formaldehyde, water and carbon dioxide are in-
dependent of temperature and denotes values of 1,
For the formation of formaldehyde the following 0, )0.5 and 0, respectively. The reaction order of
reaction orders can be derived. oxygen varies between )0.75 at low temperatures
(high coverage of O ) to 0.25 at high temperatures
aCH3 OH ¼ 1  2hCH3 OH  2hCH3 O ð17Þ (low coverage of O ). Generally the quantitative
values of the reaction orders presented in Fig. 10
aO2 ¼ 1=4  2hO2  hO  1=2hCH3 O  hH2 CO reproduce the qualitative features discussed above.
ð18Þ
4.4. Selectivity and industrial conditions
aH2 CO ¼ 2hH2 CO  2hHCOO ð19Þ
The observed selectivity versus conversion trend
aCO2 ¼ 2hCO2 ð20Þ on Fig. 3 is due to the increasing amount of
formaldehyde formed at higher conversions. In-
aH2 O ¼ 1=2 þ hCH3 O  2hH2 O ð21Þ
creasing the amount of formaldehyde leads to an

As the coverages of all intermediates except O are increased combustion of formaldehyde. This is a
low according to Fig. 7, the model explains the classical effect of a consecutive reaction path.
positive reaction order with respect to methanol Robb and Harriott [11] postulated that two routes
observed experimentally [2,11]. With respect to to CO2 formation was necessary to explain these
oxygen Eq. (18) shows, that oxygen will have a data. However our model shows that one oxida-
positive reaction order, if the coverage of oxygen, tion route for formaldehyde is enough to explain
methoxy and methanol are moderate. It can be the experiment. The reason for the different con-
seen from Fig. 2 that the model captures the ob- clusion reached by Robb and Harriott is that they
served trends in reaction order at 695 K very well used an aged catalyst for low conversion experi-
i.e. zero order in oxygen and nearly first order in ments and a fresh catalyst for high conversions.
20 A. Andreasen et al. / Surface Science 544 (2003) 5–23

They state in their paper that a fresh catalyst has a 500 K due to the blockage of free sites by O . This
lower selectivity than an aged catalyst. This means site blockage therefore inhibits the reactions below
that they amplified the effect of conversion on se- 400–500 K dependent on the partial pressures of
lectivity by using these different catalysts. The oxygen and methanol.
difference between fresh and aged catalysts may be It is informative to compare our modelÕs ex-
explained by the use of alumina support. planation of the experiment depicted in Fig. 4 and
The most important selectivity trend observed Nagy and MestlÕs own interpretation [8,9]. It
experimentally and predicted by the model is an should be noted that Nagy and MestlÕs interpre-
increased selectivity with temperature. This effect tation was not supported by simulations. Our
can be explained very simple by the microkinetic model explains the trends in Fig. 4 by a high
model. As demonstrated in Fig. 8 the apparent production of CO2 at low temperatures due to a
activation energy of formaldehyde synthesis is al- low apparent activation energy of CO2 formation
ways larger than the apparent activation energy of leading to a low selectivity. Below 500 K the sta-
formaldehyde combustion. Hence the rate of bility of O leads to site blockage and all methanol
formaldehyde formation increases more rapidly oxidation is inhibited as mentioned above. At high
than formaldehyde combustion at increasing tem- temperatures the selectivity increases due to a
perature. In order to obtain a high selectivity the higher apparent activation barrier for formalde-
industrial process should be conducted at high hyde formation compared to CO2 formation.
temperatures. However the temperatures should However Nagy and Mestl explain the selectiv-
not be high enough to promote the pyrolytic gas ity trends by the different temperature stability
phase decomposition of formaldehyde. Further- of various oxygen species (Oa and Oc ) formed,
more a methanol rich feed should be used in the exhibiting different chemical reactivities. Oa par-
process to ensure high selectivity by achieving ticipates in an oxi-dehydrogenation path which
complete oxygen consumption. Otherwise the re- dominates at low temperatures and result in both
maining oxygen would combust the formed formaldehyde and complete oxidation products.
formaldehyde. This is consistent with the work by Oc participates in a direct dehydrogenation path
Wachs and coworkers [36] who show that the se- at high temperatures which exclusively leads to
lectivity towards formaldehyde decreases from formaldehyde. As evidence for this the authors
92.3% to 69.1% when CH3 OH/O2 molar ratio de- claim that Oc is dominating at high temperatures
creases from 3.08 to 0.95 in a fixed bed reactor. and is able to react with methanol. However as
The rate of formaldehyde combustion does not mentioned earlier the uptake of Oc is probable too
decrease with temperature it just does not increase slow to keep up with the uptake of Oa . Further-
as rapidly as the rate for formaldehyde formation. more the necessary reconstruction to form Oc
Another way to increase selectivity would be to could be absent in the presence of methanol rich
reduce the contact time in the reactor. However mixtures [26]. Nagy and Mestl mention that the
the rate of these reactions are so fast that total selectivity increases dramatically above Oa de-
conversion is achieved very rapidly and this option sorption temperature (580–620 K) indicating that
is therefore not feasible. Hence the microkinetic Oc plays a significant role at high temperatures.
model gives a very simple explanation for the This argument is questionable. First of all 620 K is
choice of industrial conditions. the desorption temperature of Oa in UHV. Sec-
Nagy and Mestl [9] observes that the reaction ondly, the selectivity increases more rapidly below
ignites at approx. 450 K with a partial pressure of the desorption temperature cf. Fig. 4. Nagy and
methanol of 0.088 atm in the feed. Gavrilin and Mestl also postulate that a reaction-in-series model
Popov [7] has shown that the reaction ignites at in which formaldehyde is an intermediate product
approx. 550 K with a partial pressure of methanol for CO2 formation cannot explain the observed
of 0.435 atm. From Fig. 8 it is clear that the ap- increasing selectivity with temperature due to the
parent activation energy of both formaldehyde higher thermodynamic stability of CO2 . The au-
and CO2 formation becomes very high below thors therefore conclude that the selectivity trends
A. Andreasen et al. / Surface Science 544 (2003) 5–23 21

is due to changing between two reaction paths i.e. that only seven parameters are critical: the four
oxi-dehydrogenation and direct dehydrogenation. Arrhenius parameters of steps 5 and 10 and the
However our simulations clearly demonstrate that enthalpies of O , CH3 O and HCOO . It has not
a reaction-in-series model is able to explain this been possible to determine the Arrhenius para-
selectivity behavior (this conclusion is valid re- meters and the enthalpies of formate explicitly. In
gardless if one accept our model or not). The so- order to do so surface science experiments and/or
lution to this apparent paradox is very simple: as DFT calculations are needed probing the stability
for all other partial oxidation reactions the selec- of formate. In order to make a more complete
tivity is determined by kinetics not thermody- model including Eqs. (6)–(10) more critical pa-
namics. All the oxygen is simple consumed rapidly rameters will have to be determined. We believe
in the formation of formaldehyde leaving nothing that the experimental and theoretical knowledge of
to oxidize formaldehyde further. Furthermore if this reaction does not justify such a model at
two very different reaction pathways operate at present. The present work has shown that the
low and high temperatures, respectively one would surface mechanism proposed by Wachs and Madix
intuitively expect some kind of evident bend in the can explain industrial formaldehyde synthesis.
data at a temperature corresponding to the change However this does not explicitly prove that this is
in mechanism. This is in contrast to the nice in fact the case. In order to validate the model or
smooth data reported in Fig. 4. With regard to the discard it high quality steady-state kinetics pref-
above discussion we find our interpretation of erable close to industrial conditions are needed
Nagy and MestlÕs experiments much more consis- badly. Such kinetic experiments are lacking at
tent, simple and validated. It is well known that present due to the difficulties in avoiding mass
it is impossible to deduce reaction mechanisms transfer limitation, non-isothermity, deactivation,
explicitly from steady-state kinetics. It must be and aging effects. Furthermore it would be very
considered even more dangerous to deduce mech- interesting to investigate formaldehyde synthesis
anisms from kinetic experiments with total con- on different single crystals especially Ag(1 1 1) to
version. The experiments of Nagy and Mestl led to determine the structure sensitivity of the reaction.
100% conversion of oxygen and can therefore not
be used to deduce the intrinsic kinetics of metha-
nol oxidation. It should be stressed that our model 5. Conclusion
is not deduced from Nagy and Mestl experiments
but their experiments are explained as a conse- We have formulated a simple microkinetic
quence of our model. model for the oxidation of methanol on silver
It is well known that formaldehyde synthesis on based on surface science at UHV and low tem-
silver can be poisoned with electronegative sub- peratures. The reaction mechanism used in the
stances such as halides and sulphur, and consid- model is a simple Langmuir–Hinshelwood mech-
erable attention has been paid to their removal, anism, with a statistical mechanical treatment of
particular from process air [4]. It has been estab- gases and intermediates, having one type of active
lished that electronegative substances such as oxygen and one route to formaldehyde and carbon
chloride blocks the uptake of O and significantly dioxide, respectively. The mechanism contains two
lowers the rate of O chemisorption (Ref. [89] and slow steps: methoxy decomposition to formalde-
references therein). This poisoning is consistent hyde (step 5) and formate decomposition (step 10).
with our model because O is essential in our All other steps are assumed to be equilibrated.
model to the production of formaldehyde. Step 5 controls the rate of methanol oxidation to
formaldehyde, whereas steps 10 controls the
4.5. Critical parameters combustion of formaldehyde (selectivity). The
model has been used to simulate different ki-
The microkinetic model contains more than 100 netic experiments and it explains observed reaction
parameters (see Tables 3–5). However it turns out orders, selectivity, apparent activation enthalpies,
22 A. Andreasen et al. / Surface Science 544 (2003) 5–23

and the choice of industrial reaction conditions. [11] D.A. Robb, P. Harriott, J. Catal. 35 (1974) 176–183.
Interestingly the model explains experiments usu- [12] Q. Dai, A.J. Gellmann, Surf. Sci. 257 (1991) 103–112.
[13] J. Wang, X. Xu, J. Deng, Y. Liao, B. Hong, Appl. Surf.
ally used to disprove that the mechanism deduced Sci. 120 (1997) 99–105.
at UHV is responsible for industrial formaldehyde [14] C.-F. Mao, M.A. Vannice, J. Catal. 154 (1995) 230–244.
synthesis. The most important conclusions of this [15] B. Xinhe, D. Jingfa, J. Catal. 99 (1986) 391–399.
work is that the mechanism of Wachs and Madix [16] T.E. Felter, W.H. Weinberg, G.Y. Lastushkina, P.A.
can explain industrial formaldehyde synthesis and Zhdan, G.K. Boreskov, J. Hrbek, Appl. Surf. Sci. 16
(1983) 351–364.
that only one kind of oxygen (Oa ) needs to be [17] M.A. Barteau, M. Bowker, R.J. Madix, Surf. Sci. 94 (1980)
active in order to do so. This does not prove that 303–322.
the mechanism of Wachs and Madix is correct at [18] I.E. Wachs, R.J. Madix, Appl. Surf. Sci. 1 (1978) 303–
industrial conditions but it is a likely and simple 328.
explanation. Especially considering the recent ad- [19] M. Mavrikakis, M.A. Barteau, J. Mol. Catal. A 131 (1998)
135–147.
vances in DFT and microkinetic modeling proving [20] J. Greeley, M. Mavrikakis, J. Catal. 208 (2002) 291–300.
a close connection between surface science and [21] P. Mars, D. van Krevelen, Chem. Eng. Sci. Spec. Suppl. 3
industrial catalysis. We believe that the reason (1954) 41.
other researchers find mechanisms involving sev- [22] T. Schedel-Niedrig, X. Bao, M. Muhler, R. Schl€ ogl, Ber.
eral kinds of atomic oxygen is due to the fact that Bunsenges. Phys. Chem. 101 (7) (1997) 994–1006.
[23] X. Bao, M. Muhler, Th. Schedel-Niedrig, R. Schl€ ogl, Phys.
they deduce their mechanisms from transient ex- Rev. B 54 (3) (1996) 2249–2262.
periments where these other oxygen species be- [24] X. Bao, B. Pettinger, G. Ertl, R. Schl€ ogl, Ber. Bunsenges.
comes important due to dynamic effects, at these Phys. Chem. 97 (3) (1993) 322–325.
conditions. It is plausible that subsurface oxygen [25] X. Bao, J.V. Barth, G. Lehmpfuhl, R. Schuster, Y. Uchida,
can participate in a sort of Mars–van Krevelen R. Schl€ogl, G. Ertl, Surf. Sci. 284 (1993) 14–22.
[26] X. Bao, M. Muhler, B. Pettinger, R. Schl€ ogl, G. Ertl,
mechanism replenishing the surface oxygen. How- Catal. Lett. 22 (1993) 215–225.
ever we emphasize that we believe that these effects [27] A.J. Nagy, G. Mestl, D. Herein, G. Weinberg, E. Kitzel-
are unimportant at industrial steady-state condi- mann, R. Schlogl, J. Catal. 182 (1999) 417–429.
tions. The microkinetic model is simple and in- [28] C.T. Campbell, J. Catal. 94 (1985) 436–444.
complete and only serves to demonstrate some of [29] V.I. Bukhtiyarov, V.V. Kaichev, I.P. Prosvirin, J. Chem.
Phys. 111 (5) (1999) 2169–2175.
the misunderstandings regarding formaldehyde [30] C.I. Carlisle, T. Fujimoto, W.S. Sim, D.A. King, Surf. Sci.
synthesis in the literature. 470 (2000) 15–31.
[31] C.I. Carlisle, D.A. King, M.-L. Bocquet, J. Cerda, P.
Sautet, Phys. Rev. Lett. 84 (2000) 3899.
References [32] X. Bao, M. Muhler, B. Pettinger, Y. Uchida, G. Lehmp-
fuhl, R. Schl€ ogl, G. Ertl, Catal. Lett. 32 (1995) 171–
[1] C.N. Satterfield, Heterogeneous Catalysis, second ed., 183.
McGraw-Hill, 1991. [33] L. Leffert, J.G. van Ommen, J.R.H. Ross, J. Chem. Soc.
[2] S.K. Bhattacharyya, N.K. Nag, N.D. Ganguly, J. Catal. 23 Faraday Trans. I 83 (10) (1987) 3161–3165.
(1971) 158–167. [34] L. Leffert, J.G. van Ommen, J.R.H. Ross, Appl. Catal. 34
[3] UllmannÕs Encyclopedia of Industrial Chemistry, fifth ed., (1987) 329–339.
vol. A11, VCH Verlagsgesellschaft, 1988. [35] D.J. Sajkowski, M. Boudart, Catal. Rev. Sci. Eng. 29 (4)
[4] M.V. Twigg, Catalyst Handbook, Manson Publishing, (1987) 325–360.
1996. [36] C.-B. Wang, G. Deo, I.E. Wachs, J. Phys. Chem. B 103
[5] Kirk-Othmer Encyclopedia of Chemical Technology, vol. (1999) 5645–5656.
11, John Wiley & Sons, 1994. [37] L. Leffert, J.G. van Ommen, J.R.H. Ross, Appl. Catal. 23
[6] J.F. Griffiths, G. Skirrow, Oxidat. Combust. Rev. 3 (1968) (1986) 385–402.
47–96. [38] M. Qian, M. Liauw, G. Emig, Appl. Catal. A 238 (2003)
[7] V.N. Gavrilin, B.I. Popov, Kinet. Catal. (USSR) 6 (1965) 211–222.
799. [39] A.E. Obraztsov, B.I. Popov, M.P. Shashalevich, L.P.
[8] A. Nagy, G. Mestl, T. R€ uhle, G. Weinberg, R. Sch€ ogl, Taraban, E.A. Pronina, A.G. Shevchenko, Kinet. Catal.
J. Catal. 179 (1998) 548–559. (USSR) 12 (1971) 970.
[9] A. Nagy, G. Mestl, Appl. Catal. A 188 (1999) 337–353. [40] A.C. van Veen, O. Hinrichsen, M. Muhler, J. Catal. 210
[10] I.E. Wachs, R.J. Madix, Surf. Sci. 76 (1978) 531–558. (2002) 53–66.
A. Andreasen et al. / Surface Science 544 (2003) 5–23 23

[41] J. Dumesic, D. Rudd, L. Aparicio, J. Rekoske, A. Trevi~ no, [67] C. Campbell, Surf. Sci. 157 (1985) 43–60.
The Microkinetics of Heterogeneous Catalysis, ACS, 1993. [68] R.B. Grant, R.M. Lambert, Surf. Sci. 146 (1984) 256–
[42] R.D. Cortright, J.A. Dumesic, Adv. Catal. 46 (2001) 161– 268.
264. [69] P. Stoltze, Prog. Surf. Sci. 65 (2000) 65–100.
[43] P. Stoltze, Phys. Scr. 36 (1987) 824–864. [70] C.V. Ovesen, B.S. Clausen, J. Schiøtz, P. Stoltze, H.
[44] C.V. Ovesen, B.S. Clausen, B.S. Hammershøi, G. Steffen- Topsøe, J.K. Nørskov, J. Catal. 168 (1997) 7133.
sen, T. Askgaard, I. Chorkendorff, J.K. Nørskov, P.B. [71] H. Topsøe, C.V. Ovesen, B.S. Clausen, N.-Y. Topsøe,
Rasmussen, P. Stoltze, P. Taylor, J. Catal. 158 (1996) 170– Stud. Surf. Sci. Catal. 109 (1997) 121.
180. [72] F. Reif, Fundamentals of Statistical and Thermal Physics,
[45] T. Askgaard, J.K. Nørskov, C.V. Ovesen, P. Stoltze, McGraw-Hill, 1985.
J. Catal. 156 (1995) 229–242. [73] M.W. Chase Jr. (Ed.), NIST-JANAF Thermochemical
[46] C.T. Campbell, Topics Catal. 1 (1995) 353. Tables, fourth ed., J. Phys. Chem. Ref. Data, Monograph
[47] S.H. Oh, G.B. Fisher, J.E. Carpenter, D.W. Goodman, No. 9, Parts I and II.
J. Catal. 100 (1986) 360. [74] P.A. Taylor, P.B. Rasmussen, C.V. Ovesen, P. Stoltze, I.
[48] V.P. Zhdanov, B. Kasemo, Surf. Sci. Rep. 20 (1986) 111. Chorkendorff, Surf. Sci. 261 (1992) 191–206.
[49] N. Waletzko, L.D. Schmidt, AIChE J. 34 (1988) 1146. [75] C.V. Ovesen, P. Stoltze, J.K. Nørskov, C.T. Campbell,
[50] D.A. Hickman, L.D. Schmidt, AIChE J. 39 (1993) 1164. J. Catal. 134 (1992) 445–446.
[51] C.T. Campbell, M.T. Paffett, Surf. Sci. 143 (1984) 517–535. [76] R. Zhang, A.J. Gellman, J. Phys. Chem. 95 (1991) 7433–
[52] M.A. Barteau, R.J. Madix, J. Chem. Phys. 74 (7) (1981) 7437.
4144–4149. [77] T. Askgaard, Catalytic reaction mechanisms, Ph.D. Thesis,
[53] I. Stensgaard, E. Laegsgaard, F. Besenbacher, J. Chem. Physics Department, Technical University of Denmark,
Phys. 103 (22) (1995) 9825–9831. 1994.
[54] A.G. Sault, R.J. Madix, Surf. Sci. 176 (1986) 415–424. [78] P.A. Redhead, Vacuum 12 (1962) 203.
[55] M. Bowker, M.A. Barteau, R.J. Madix, Surf. Sci. 92 (1980) [79] A. Raukema, D.A. Butler, F. Box, A.W. Kleyn, Surf. Sci.
528–548. 347 (1996) 151–168.
[56] L.E. Fleck, Z.C. Ying, M. Feehery, H.L. Dai, Surf. Sci. 296 [80] S.R. Bare, K. Griffiths, W.N. Lennard, H.T. Tang, Surf.
(1993) 400–409. Sci. 342 (1995) 185–198.
[57] E.M. Stuve, R.J. Madix, B.A. Sexton, Surf. Sci. 111 (1981) [81] C. Stegelmann, P. Stoltze, in preparation.
11–25. [82] R.I. Masel, Principles of Adsorption and Reaction on Solid
[58] H.G. Jenniskens, P.W.F. Dorlandt, M.F. Kadodwala, Surfaces, John Wiley & Sons, 1996.
A.W. Kleyn, Surf. Sci. 357–358 (1996) 624–628. [83] B. Hammer, L.B. Hansen, J.K. Nørskov, Phys. Rev. B 59
[59] I.E. Wachs, R.J. Madix, J. Catal. 53 (2) (1978) 208–227. (11) (1999) 7413–7421.
[60] B.A. Pepley, J.C. Amphlett, L.M. Kearns, R.F. Mann, [84] D.H. Vanderbilt, Phys. Rev. B. 41 (1990) 7892.
Appl. Catal. A 179 (1999) 31–49. [85] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77
[61] B.A. Sexton, R.J. Madix, Surf. Sci. 105 (1981) 177–195. (18) (1996) 3865–3868.
[62] E.M. Stuve, R.J. Madix, B.A. Sexton, Surf. Sci. 119 (1982) [86] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson,
279–290. M.R. Pederson, D.J. Singh, C. Fiolhais, Phys. Rev. B
[63] H.M.C. Sosnovsky, J. Chem. Phys. 23 (8) (1955) 1486– (1992) 6671.
1490. [87] J.A. White, D.M. Bird, Phys. Rev. B 50 (1994) 4954.
[64] H.M.C. Sosnovsky, J. Phys. Chem. Solids 10 (1959) 304– [88] W.-X. Li, C. Stampfl, M. Scheffler, Phys. Rev. Lett. 90
310. (2003) 256102.
[65] J. Bagg, H. Jaeger, J.V. Sanders, J. Catal. 2 (1963) 449–464. [89] C.T. Campbell, J. Catal. 99 (1986) 28–38.
[66] H. Jaeger, J. Catal. 9 (1967) 237–250. [90] E.L. Force, A.T. Bell, J. Catal. 38 (1975) 440–460.

S-ar putea să vă placă și