You are on page 1of 6

Journal of Molecular Catalysis A: Chemical 393 (2014) 296301

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


journal homepage: www.elsevier.com/locate/molcata

Effects of strain on PdZn(1 0 0) for methoxide decomposition:


A DFT study
Xiang Li a , Rern Jern Lim a , Jong Min Lee a , Xin Wang a , Kok Hwa Lim a,b,
a
b

School of Chemical and Biomedical Engineering, Nanyang Technological University, 62 Nanyang Drive, Singapore 637459, Singapore
Singapore Institute of Technology, 10 Dover Drive, Singapore 138693, Singapore

a r t i c l e

i n f o

Article history:
Received 7 November 2013
Received in revised form 19 June 2014
Accepted 21 June 2014
Available online 28 June 2014
Keywords:
PdZn
DFT
Methanol
Decomposition
Strain

a b s t r a c t
Density functional theory has been used to systematically investigate the adsorption of methoxide and
its decomposed intermediates (i.e. H, O, CH3 and CH2 O) as well as methoxide decomposition reaction
on strained PdZn(1 0 0) surfaces. The reaction and activation energies of methoxide C H and C O bond
breaking process showed that expansive strain increased the activity of the PdZn(1 0 0) surface toward
methoxide decomposition, but reduced its selectivity.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Rising greenhouse gas concentration has stimulated interest on the research for new clean technologies, such as power
generation from hydrogen fuel cells. Hydrogen becomes an
attractive option for its cleanness [1]. Among the chemical processes for hydrogen production, methanol steam reforming (MSR),
CH3 OH + H2 O CO2 + H2 [1], is a promising method for its potential application for on board hydrogen generation (i.e. fuel cell). The
Pd/ZnO catalyst was shown to be a viable alternative catalyst to Cu
as it possesses better thermal stability with comparable reactivity.
The classical catalyst, Cu, used in MSR will become unstable and sinter at high temperatures [2,3]. PdZn (1:1) alloy has been identied
as the active component of the Pd/ZnO catalysts [2].
According to previous reports, MSR begins from methoxide
adsorption onto the catalytic surfaces [3]. Further reaction results
in the decomposition of CH3 O into two pathways: (i) the C H bond
cleavage produces CH2 O and H, which eventually leads to the formation of CO/CO2 and H2 ; and (ii) the C O bond scission forms CH3
and O, which in turn leads to the production of C that poisons the
catalysts [47]. Previous study suggested that the C H bond scission is the rate determining step of methanol decomposition [5]

Corresponding author at:Singapore Institute of Technology, 10 Dover Drive,


Singapore 138693. Tel.: +65 65922168; fax: +65 65921190.
E-mail address: kokhwa.lim@singaporetech.edu.sg (K.H. Lim).
http://dx.doi.org/10.1016/j.molcata.2014.06.029
1381-1169/ 2014 Elsevier B.V. All rights reserved.

hence is the focus point of our current study on the activity and
selectivity of methoxide decomposition.
Recently, some studies [813] have shown that strain can signicantly alter the reactivity of transition metal catalysts [8,9,12].
Mavrikakis et al. [9] studied the effect of strain on the adsorption
of O and CO on Ru(0 0 0 1) surfaces, and reported that surface reactivity increases with expansive strain. Mavrikakis et al. also studied
the adsorption O and CO on at and stepped Au(1 1 1) surfaces
and reported that an expansive strain can increase the reactivity
of Au catalysts [10]. Other studies by Zhuang et al. [13] reported
that compressive strain increases the adsorption strength of OH
on Pd(1 1 1). In addition, Nrskov et al. [11] also reported that
expansively strained Fe(1 1 1) and Fe(1 1 0) surfaces favored N2 dissociation. All of these studies have concluded that strain on catalytic
materials can play an important role in the reactivity and selectivity of chemical reactions. Strain can be introduced experimentally
through their support materials. Hence, this theoretical study aims
to demonstrate the potential applications for strain on catalytic
reactions while delineating the effect from other electronic effects
by the supporting substrate.
Relevant methoxide decomposition process on Cu, Pd,
PdZn(1 1 1), PdZn(1 0 0) and (2 2 1) surfaces have been previously studied by Lim et al. [5,14,15]. They found that the reaction
proles on Cu(1 1 1) and PdZn(1 1 1) surfaces are very similar [5].
CH3 O adsorbs more weakly on Pd(1 1 1) compared to the Cu(1 1 1)
surface. Consequently, energy barriers for both C H and C O
bond cleavage on Pd(1 1 1) were much lower than on Cu(1 1 1). In
addition, it was found that the activation energy of C H and C O

X. Li et al. / Journal of Molecular Catalysis A: Chemical 393 (2014) 296301

bond breaking of methoxide on PdZn(1 0 0) and (2 2 1) surfaces


[14] are lower than on PdZn(1 1 1).
To the best of our knowledge, there are no previous theoretical
studies on the effect of strain for PdZn catalysts. Since PdZn was
reported to have comparable activity and selectivity to Cu in the
MSR process, it is valuable to study the effect of strain on PdZn for
the MSR reaction. In this paper, we aim to investigate the effect of
strain on the adsorption of methoxide and its decomposed intermediates as well as the reactivity of methoxide C H and C O bond
breaking on PdZn(1 0 0) surface.
This paper is organized as follows. In Section 2, we present the
computational methods and the models employed. In Section 3,
we provide the comparative discussions on the adsorption characteristics, reaction energetics of methoxide and its decomposition
intermediates (i.e. H, CH2 O, O, CH3 ) on strained PdZn(1 0 0) surfaces and how further methoxide decomposition is affected by the
applied strain. In Section 4, we discuss the reaction pathways and
activation energies of methoxide decomposition. Finally, we conclude our results in Section 5.
2. Computational details and surface models
All calculations were performed using the plane-wave based
Vienna ab initio simulation package (VASP) [1618] using the PBE
generalized-gradient approximation for the exchange-correlation
functional [19]. The interaction between atomic cores and electrons
were described by the projector augmented wave method (PAW)
[20,21]. For integrations over the Brillouin zone, we combined a
(3 3 1) MonkhorstPack grid [22]. We adopted an energy cut-off
of 400 eV throughout our study. Previous studies showed that these
parameters are sufcient for our current purposes [5,2330]. The
coordinates of adsorbates and substrate atoms included in the optimization procedure were optimized until the forces acting on them
were less than 0.01 eV/. The nudged elastic band (NEB) method
was used to locate the transition states [31].
All studies carried out in this paper were on the four layer slab
model of strained PdZn(1 0 0) surfaces. We employed a 4 4 surface
unit cell, comprising of 16 metal atoms per layer. This enables us to
consider a surface coverage as low as 1/16. We would like to refer
our reader to Ref. [22] for the various adsorptions studied. Adsorbates were positioned on one side of the slab with vacuum spacing
of 1 nm. For calculations of surface optimization, the top two layers were relaxed while the bottom two were xed. Adsorption sites
studied are shown in Fig. 1.
Binding energies were calculated as shown in Eq. (1) while the
relative binding energies were calculated as shown in Eq. (2)
Eb = Ead/sub Ead Esub

(1)

297

Fig. 1. Adsorption sites on PdZn(1 0 0) surface: (1) 4HollowZn2 , (2) 4HollowPd2 , (3)
TBTPdZn (4) TopPd and (5) 3HollowZn2Pd . First PdZn layer: dark blue sphere, Pd; orange
sphere, Zn. Second PdZn layer: light blue sphere, Pd; yellow sphere, Zn. (For interpretation of the references to color in this gure legend, the reader is referred to the
web version of the article.)

3. Adsorptions of H, O, CH2 O and CH3 O on PdZn(1 0 0)


surface
3.1. Optimized and reconstructed PdZn(1 0 0) surface
The calculated lattice constant (LC) of the PdZn bulk structure
was 4.14 A [22], which is in good agreement with the reported
experimental value of 4.11 A [4] and previously reported theoretical value of 4.15 A using PW91 functional [32]. Compressive and
to +5% (4.35 A)
were introduced
expansive strains from 5% (3.93 A)
parallel to the surface by changing the optimized bulk LC in 2%
increments. This has been shown to be an effective way to study
the effects of the strain on transition metal catalysts in previous
studies [12,3335]. We have also studied the effects of strain on the
substrates surface optimization and they are tabulated in Table 1.
Our calculations showed that the most stable structure is the
slightly compressed PdZn surface (1%) followed by the surface
with 3% strain. The surfaces have stabilization energies of 0.019
and 0.016 eV/2 with respect to the bulk truncated structure (see
Table 1), respectively. We have studied all possible adsorption sites
and found that the most stable sites for all strained surfaces remain
unchanged with respect to the unstrained surface. All calculated
adsorption energies are tabulated in the Supplementary material.
Optimized geometries for all species are shown in Fig. 2.

where Ead/sub is the total energy of the adsorption complexes, Ead


and Esub are the total energy of the adsorbate in the gas phase and
of the clean substrate, respectively.

3.2. Atomic H adsorption

Erelative = Eb + Esub

Atomic H usually prefers to bind on the highest coordination site


(4Hollow) of metal surfaces such as Pd [5,14,27,35,36]. Similarly, H
binds to the highest coordination site on the PdZn(1 0 0) surface

(2)

where Esub is the energy difference between optimized strained


surface and unstrained unoptimized surface (LC = 0%).
Er =



(Eb )P +

(Ead )P

 

(Eb )R +

(Ead )R

(3)

Reaction energies were calculated as shown in Eq. (3) where, (Eb )p


and (Eb )R are the summation of binding energy of all products and
reactants, respectively. (Ead )p and (Ead )R refers to the summation of total energy of all the products and reactants in the gas phase,
respectively. Negative values for all three equations denote an
exothermic reaction or favorable adsorption. All adsorbed species
in this study were calculated with respect to their respective gas
phase species except for H and O.

Table 1
Energy differences (in eV/2 ) per surface area between optimized bulk truncated
strained PdZn(1 0 0) surfaces with respect to unstrained bulk truncated surface.
LC (%)

LC ()

Esub -optimized

5
3
1
0
+1
+3
+5

3.93
4.01
4.10
4.14
4.18
4.26
4.35

0.016
0.016
0.019
0.011
0.003
0.039
0.086

298

X. Li et al. / Journal of Molecular Catalysis A: Chemical 393 (2014) 296301

Fig. 2. Pertinent geometries of most stable adsorption complexes for all species (in ) on relaxed PdZn(1 0 0) surface. Only rst layer is shown for clarity.

(4HollowPd2 ) with a binding energy 2.66 eV. On strained surfaces


(see Fig. 3), the adsorption energy for surfaces increased slightly
with the increasing expansive strain. This observation is in line with
the weakened PdPd interaction due to the increase in the LC, hence
allowing for a stronger H Pd bond.
Compared to Cu(1 0 0) surfaces, where expansive strain makes
the complex unstable due to longer H Cu bond length [35], H Pd
bond length reduces with expansive strain on the PdZn(1 0 0) surface. This can be rationalized by the covalent radius:the covalent
radius for H Pd is 1.76 A [37]. From Table 2, H Pd distance reduces
from 1.84 to 1.81 A when expansive strain was introduced through
the change of LC from 5% to +5%, hence making the interaction
more favorable.

3.3. Atomic O adsorption


Similar to previous study on PdZn(1 1 1) surface [14], atomic O
preferably occupies the 4HollowZn2 sites. Our calculations show
that atomic O binds much more strongly on PdZn(1 0 0) compared
to PdZn(1 1 1) with a binding energy of 4.94 and 4.75 eV, respectively, on the unstrained surface. This is due to the much shorter
O Zn bond length of 1.94 A on the PdZn(1 0 0) surface compared
to 1.98 A on the PdZn(1 1 1) surface. The introduction of expansive strain resulted in the increase of binding energy from 4.64
to 4.99 eV. However, the O Zn bond length remained relatively
unchanged at 1.94 A with respect to expansive strain. Thus, the
O Zn bonding is most stable at +5% LC. A similar trend is reported
on the Cu(1 0 0) surface [35], where atomic O binds more strongly
from 5.30 to 5.51 eV as expansive strain is introduced. No calculated bond length was reported in the Cu(1 0 0) study. This
phenomena can be rationalized in terms of charge repulsion where
the negatively charged adsorbed O is repelled by surface electrons
hence is stabilized at larger LC where the metalmetal bonds are
weaker.

3.4. Methyl adsorption


Methyl binds to the PdZn(1 0 0) surface via the C atom, similar
to previous studies on Cu(1 0 0) and PdZn(1 1 1) [14,35]. Our calculations show that the adsorption of methyl is most stable on the
TopPd site with the CPd distance of 2.15 A and a binding energy of
1.62 eV. Previous studies showed that atomic carbon binds much
more favorably on Pd site compared to Zn thus methyl should
behave similarly as it binds via its carbon atom [5]. Like atomic
H and O, the binding strength of methyl increases with expansive
strain (see Fig. 3). The C Pd bond length remains unchanged with
respect to increasing expansive strain at 2.15 A (see Table 2). This
is due to the binding location of methyl (TopPd ) where changes in
LC have no effect with respect to the binding site.
In contrast, methyl favors the bridge site on the Cu(1 0 0) surface [35]. This is due to the different metal atoms present on the
PdZn(1 0 0) surface but not on the Cu(1 0 0) surface. Methyl interacts with the 3s orbital Cu atoms via its 2 molecular orbital (MO)
[5]. On the Cu(1 1 1) surface, methyl binds at the 4Hollow site as the
surface is much more compact compared to the Cu(1 0 0) surface
where the shortest CuCu distance is the bridge site instead of 4Hollow [5]. Although there is difference in binding sites on Cu(1 0 0),
the trend of binding strength of methyl is similar to PdZn(1 0 0)
when there is increase in LC. This is due to the weakening of Cu Cu
bond which in turn enhances the stability of adsorbed methyl [35].
3.5. Formaldehyde adsorption
Formaldehyde binds to the surface in a top-bridge-top (TBT)
conguration. Specically, bonds are formed with surface Pd and
Zn via the C and O atoms of formaldehyde, respectively (TBTPdZn ).
The binding site and very weak binding of CH2 O is similar to previous studies on PdZn(1 1 1) and Cu(1 0 0) [14,35]. The weak binding
is due to the fact that CH2 O is a close shell molecule. Similar to
other adsorbates, increasing the LC will result in an increase in the
binding energy of CH2 O, while compressive strain will destabilize

X. Li et al. / Journal of Molecular Catalysis A: Chemical 393 (2014) 296301

299

Fig. 3. Change in binding energy with respect to change in LC: (a) hydrogen at 4HollowPd2 , (b) oxygen at 4HollowZn2 , (c) methyl at TopPd , (d) formaldehyde at TBTPdZn , and
(e) methoxide at 4HollowZn2 .

the CH2 O adsorption (see Fig. 3). The change in LC will affect the
binding energy as the TBT structure is dependent on the PdZn
length. As the surface expands, it allows for better 2 interactions
between CH2 O and the PdZn(1 0 0) substrate as observed by the
reduction of C Pd bond length from 2.20 to 2.18 A while the O Zn
bond length remains unchanged at 2.05 A (see Table 2). From the
perspective of covalent radii, the summation of covalent radius of
Pd and C is 2.15 A (0.76 A and 1.39 A for C and Pd, respectively [37]).
Similarly, O Zn has a stable bond length of 1.94 A (Section 3.3).
An additional possible reason for the weak binding of CH2 O on

compressed PdZn(1 0 0) surface is the strained bond angles of


O Zn and C Pd. On the Cu(1 0 0) surface, the Cu O C bond angle
increased from 111.6 to 119.4 as the LC is increased to +5%. On
the other hand, the Cu C O bond angle increased much less from
101.1 to 102.9 [35].
3.6. Methoxide adsorption
Our calculations showed that methoxide binds to the PdZn(1 0 0)
surface via its O atom, consistent with previous studies on Cu(1 0 0)

Table 2
Bond lengths (in ) for all species at their respective binding sites and binding atoms.
Binding site

H
4HollowPd2

O
4HollowZn2

CH3
TopPd

CH2 O
TBTPdZn

CH3 O
4HollowZn2

LC (%)

H Pd

O Zn

C Pd

C Pd

O Zn

O Zn

5
3
1
0
+1
+3
+5

1.84
1.83
1.82
1.82
1.82
1.81
1.81

1.95
1.95
1.94
1.94
1.94
1.94
1.94

2.15
2.15
2.15
2.15
2.14
2.14
2.14

2.2
2.2
2.19
2.19
2.19
2.18
2.18

2.05
2.05
2.05
2.05
2.05
2.05
2.05

2.13
2.11
2.1
2.09
2.08
2.07
2.06

300

X. Li et al. / Journal of Molecular Catalysis A: Chemical 393 (2014) 296301


Table 3
Reaction and activation (in brackets) energy (in eV) for methoxide C H and C O
bond breaking on strained PdZn(1 0 0).a
LC (%)

CH3 O CH2 O + H

CH3 O CH3 + O

Er (Ea )

5
3
1
0
1
3
5

1.03 (1.44)
0.95 (1.32)
0.87 (1.25)
0.85 (1.13)
0.80 (1.08)
0.73 (0.95)
0.67 (0.82)

0.98 (1.93)
0.90 (1.83)
0.83 (1.76)
0.80 (1.70)
0.78 (1.51)
0.74 (1.16)
0.68 (0.94)

0.04 (0.49)
0.05 (0.51)
0.04 (0.51)
0.05 (0.57)
0.02 (0.43)
0.01 (0.21)
0.01 (0.12)

the most thermodynamically favorable adsorption for all species


studied [35].

Fig. 4. Relative binding energies for all species on various LC.

and PdZn(1 1 1) surfaces [14,35]. The most stable complex is at the


Zn dominated 4HollowZn2 sites which is in line with the study on
PdZn(1 1 1) surface [5]. Like other species in this study, the binding
energy of methoxide increases with increasing LC. Binding energy
increases from 2.20 to 2.53 eV while the O Zn bond length
decreases from 2.13 to 2.06 A (see Fig. 3). On Cu(1 0 0) surface, the
binding energy follows the same trend but showed a slight decrease
at +5% LC. This was attributed to the increase in Cu Cu bond length
that elongates the O Cu bond, hence reducing the effectiveness of
binding at the hollow site.
3.7. Relative binding energy
The effects of strain and binding energy are summarized in Fig. 4.
From Fig. 4, we can see that the most thermodynamically favorable adsorption for all species studied occurs when the substrate is
slightly compressed at 1% LC. This is most likely due to the higher
stability of the PdZn(1 0 0) surface when it is slightly compressed
(See Table 1) where more energy is released during optimization.
The previous study using Cu(1 0 0) arrived at the same conclusion
where the slightly compressed Cu(1 0 0) surface (1%) provides

4. Activation energy and reaction pathway of CH3 O


decomposition on PdZn(1 0 0) surface
The most stable geometry and adsorption site for all species was
used in this study of C O and C H bond breaking processes. The initial state (IS) selected was CH3 O while the remaining species (CH2 O,
CH3 , H and O) were used for the nal state (FS) of the reaction pathways. For the ease of discussion, geometries for IS, transition states
(TS) and FS are shown in Fig. 5. All activation and reaction energy
are shown in Table 3 for comparison.
4.1. C H bond scission
We nd that the TS for the decomposition of CH3 O into CH2 O
and H were similar for all investigated surfaces. The transition state
illustrates that one C H bond is stretched by 0.5 A (Fig. 5). Subsequently, this leads to the formation of a new stable H Pd bond thus
breaking the C H bond. The remaining CH2 O fragment binds at the
TBTPdZn site with C and O binding to Pd and Zn, respectively. These
results are in good agreement with previous study [14] on methoxide decomposition on unstrained PdZn(1 0 0) surface. From Table 3,
the activation and reaction energies for the C H bond scission reaction reduces with the increasing expansive strain. This shows that
expansive strain can enhance the reactivity of C H bond breaking
by lowering the energy barriers (by 0.05 eV, 0.18 eV and 0.31 eV for
+1%, +3% and +5% LC, respectively) with respect to the unstrained

Fig. 5. Pertinent geometries of initial, transition and nal state structures (in ) for C H bond scission (Top) and C O bond scission. Only rst layer is shown for clarity.

X. Li et al. / Journal of Molecular Catalysis A: Chemical 393 (2014) 296301

surface. This is consistent with some previous studies on Au, Cu, Ru


and Ni surfaces [9,10]. It can be elucidated that the expansive strain
leads to a more active Pd and Zn centers for the catalysis process
which promotes C H bond scission.

301

hence potentially competing with the C H bond scission process.


Therefore, expansive strain can enhance the activity but reduces
the selectivity of the PdZn(1 0 0) surface.
Acknowledgments

4.2. C O bond scission


Similar to C H bond scission, transition states have similar
geometries for all investigated surfaces. From Fig. 5, CH3 shifts
toward the nearest Pd atom and the C O bond is signicantly elongated by 0.6 A in the TS. Subsequently, CH3 is adsorbed on top
of the Pd atom while the atomic O shifts from the 4HollowZn2
to 3HollowZn2Pd site in the FS. Like the C H bond scission process, these results are in line with previous study on unstrained
PdZn(1 0 0) surface [14]. The activation and reaction energies show
a similar trend to the C H bond scission process where both energies decrease with increasing expansive strain.
The C H bond scission process is generally preferred due to its
lower activation and reaction energies compared to the C O bond
scission. On unstrained and compressed surfaces, there is little difference in the selectivity where E for both energies remains fairly
constant thus favoring the C H bond scission process (see Table 3).
When expansive strain is introduced, both energies for C O bond
scission decrease at a faster rate than for C H bond scission hence
reducing E. This means that increasing the LC could potentially
cause a reduction in selectivity toward formaldehyde despite being
able to increase the activity of the PdZn(1 0 0) surface.
Lastly, due to the weak binding energies of CH2 O, we have calculated the van der Waals interaction of our systems and found that
the binding energies of CH3 O, CH2 O and CH3 changed by about
0.35 eV and atomic H and O by 0.15 eV for all surfaces studied.
This results in the reaction energy to be reduced by approximately
0.13 eV for both pathways on all surfaces. However, these do not
alter our conclusion above.
5. Conclusions
In the current work, we have studied the effects of strain on
the stability of the PdZn(1 0 0) surface. We then investigated the
effect of compressive and expansive strains on the adsorption of
atomic H, atomic O, methyl, formaldehyde and methoxide on the
PdZn(1 0 0) surface. Generally, atomic H and O favors highly coordinated hollow sites (H at Pd dominated sites while O at Zn dominated
sites). Methyl favors the TopPd site and formaldehyde forms a stable structure in a TBTPdZn conguration. All of the species studied
followed the same trend of increased stabilization through the
introduction of expansive strain. The slightly compressed surface
(LC = 1%) was found to provide the most thermodynamically
favorable adsorption for all species studied.
The introduction of expansive strain enhances both C H and
C O scission process by effectively lowering activation and reaction energies. The PdZn(1 0 0) surface is selective toward the C H
bond scission process as this pathway requires much lower energy.
The selectivity remains unchanged when compressive strained is
introduced but becomes much more competitive when expansive
strained is introduced. This is due to the more rapid reduction in
activation and reaction energy for the C O bond scission process

X.L. acknowledges funding from China Scholarship Community


(No. [2008]3019) and NTU. R.J.L. is grateful for research scholarship
from NTU. This project is supported by NTU grant M58120000 and
RG 110/06.
Appendix A. Supplementary data
Supplementary data associated with this article can be
found, in the online version, at http://dx.doi.org/10.1016/j.molcata.
2014.06.029.
References
[1] D.L. Trimm, Z.I. nsan, Catal. Rev. 43 (2001) 3184.
[2] N. Iwasa, T. Mayanagi, S. Masuda, N. Takezawa, React. Kinet. Catal. Lett. 69
(2000) 355360.
[3] N. Iwasa, N. Takezawa, Top. Catal. 22 (2003) 215224.
[4] Z.-X. Chen, K.M. Neyman, A.B. Gordienko, N. Rsch, Phys. Rev. B 68 (2003)
075417.
[5] Z.-X. Chen, K.M. Neyman, K.H. Lim, N. Rsch, Langmuir 20 (2004) 80688077.
[6] Z.-X. Chen, K.M. Neyman, N. Rsch, Surf. Sci. 548 (2004) 291300.
[7] K.M. Neyman, R. Sahnoun, C. Inntam, S. Hengrasmee, N. Rsch, J. Phys. Chem. B
108 (2004) 54245430.
[8] M. Gsell, P. Jakob, D. Menzel, Science 280 (1998) 717720.
[9] M. Mavrikakis, B. Hammer, J.K. Nrskov, Phys. Rev. Lett. 81 (1998) 28192822.
[10] M. Mavrikakis, P. Stoltze, J.K. Nrskov, Catal. Lett. 64 (2000) 101106.
[11] A. Logadottir, J.K. Nrskov, Surf. Sci. 489 (2001) 135143.
[12] L. Grabow, Y. Xu, M. Mavrikakis, Phys. Chem. Chem. Phys. 8 (2006) 33693374.
[13] Y.G. Chen, L. Zhuang, J.T. Lu, Chin. Chem. Lett. 18 (2007) 13011304.
[14] Z.-X. Chen, K.H. Lim, K.M. Neyman, N. Rosch, Phys. Chem. Chem. Phys. 6 (2004)
44994504.
[15] Z.-X. Chen, K.H. Lim, K.M. Neyman, N. Rsch, J. Phys. Chem. B 109 (2005)
45684574.
[16] G. Kresse, J. Hafner, Phys. Rev. B 47 (1993) 558561.
[17] G. Kresse, J. Furthmller, Comput. Mater. Sci. 6 (1996) 1550.
[18] G. Kresse, J. Furthmller, Phys. Rev. B 54 (1996) 1116911186.
[19] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 38653868.
[20] P.E. Blchl, Phys. Rev. B 50 (1994) 1795317979.
[21] G. Kresse, D. Joubert, Phys. Rev. B 59 (1999) 17581775.
[22] H.J. Monkhorst, J.D. Pack, Phys. Rev. B 13 (1976) 51885192.
[23] K.H. Lim, L.V. Moskaleva, N. Rsch, ChemPhysChem 7 (2006) 18021812.
[24] K.H. Lim, Z.-X. Chen, K.M. Neyman, N. Rsch, J. Phys. Chem. B 110 (2006)
1489014897.
[25] K.H. Lim, Z.-X. Chen, K.M. Neyman, N. Rsch, Chem. Phys. Lett. 420 (2006)
6064.
[26] K.H. Lim, K.M. Neyman, N. Rsch, Chem. Phys. Lett. 432 (2006) 184189.
[27] A.B. Mohammad, I.V. Yudanov, K.H. Lim, K.M. Neyman, N. Rosch, J. Phys. Chem.
C 112 (2008) 16281635.
[28] C.Y. Yue, K. Lim, Catal. Lett. 128 (2009) 221226.
[29] J.-Y. Bo, S. Zhang, K. Lim, Catal. Lett. 129 (2009) 444448.
[30] X. Li, M.M. Wong, K. Lim, Theor. Chem. Acc. 127 (2010) 401409.
[31] G. Mills, H. Jnsson, G.K. Schenter, Surf. Sci. 324 (1995) 305337.
[32] K.M. Neyman, K.H. Lim, Z.-X. Chen, L.V. Moskaleva, A. Bayer, A. Reindl, D.
Borgmann, R. Denecke, H.-P. Steinruck, N. Rosch, Phys. Chem. Chem. Phys. 9
(2007) 34703482.
[33] Y. Xu, M. Mavrikakis, Surf. Sci. 494 (2001) 131144.
[34] J. Wintterlin, T. Zambelli, J. Trost, J. Greeley, M. Mavrikakis, Angew. Chem. Int.
Ed. 42 (2003) 28502853.
[35] A.Y. Foo, K. Lim, Catal. Lett. 127 (2009) 113118.
[36] A.B. Mohammad, K. Hwa Lim, I.V. Yudanov, K.M. Neyman, N. Rosch, Phys. Chem.
Chem. Phys. 9 (2007) 12471254.
[37] F.H. Allen, O. Kennard, D.G. Watson, L. Brammer, A.G. Orpen, R. Taylor, J. Chem.
Soc. Perkin Trans. 2 (1987) S1S19.

You might also like