Você está na página 1de 14

J Clust Sci (2008) 19:601614

DOI 10.1007/s10876-008-0210-9
ORIGINAL PAPER

Modeling the Adsorption of CO on Small Pt, Fe and Co


Clusters for the FischerTropsch Synthesis
Rafael de Souza Monteiro Llian W. C. Paes
Jose Walkimar de M. Carneiro Donato A. G. Aranda

Received: 16 June 2008 / Published online: 7 November 2008


Springer Science+Business Media, LLC 2008

Abstract The adsorption of CO on Fe, Pt and Co clusters was modeled by the


DFT approach using the B3LYP and the BPW91 functionals together with the
LANL2DZ and the 6-31G(d) basis set. These calculations show that although CO
adsorbs more strongly on Pt than on either Fe or Co, the dissociation energy on Fe
and on Co is lower than the corresponding dissociation energy on Pt. Therefore, the
activation energy for dissociation is not determined by the adsorption energy.
Additionaly, the CO bond distances also do not show any correlation to the
adsorption energy.
Keywords

CO adsorption  Metallic clusters  FischerTropsch Synthesis

Introduction
Adsorption of molecular species on metallic surfaces usually represents the first step
in their activation on catalyst sites. Therefore, understanding the adsorption forms
and reaction steps on surfaces may help one design catalysts that have improved
qualities, such as higher selectivity and lower costs. The chemisorption of molecules
such as carbon monoxide has been thoroughly studied [15], because there is an
increasing interest in understanding the reaction mechanism that occurs in the gas/
metal interface. This attention is mainly motivated by the role that metals play as
catalysts on reactions leading to high-quality transportation fuels [6].
R. de Souza Monteiro  D. A. G. Aranda
Laboratorio de Tecnologias Verdes (GreenTec)Escola de Qumica, Centro de Tecnologia,
Universidade Federal do Rio de Janeiro, Room E-211, Rio de Janeiro, RJ, Brazil
L. W. C. Paes  J. W. de M. Carneiro (&)
Departamento de Qumica Inorganica, Instituto de Qumica, Universidade Federal Fluminense,
Outeiro de Sao Joao Batista, s/n, 24020-141 Niteroi, RJ, Brazil
e-mail: walk@vm.uff.br

123

602

R. de Souza Monteiro et al.

In this work, theoretical studies of carbon monoxide adsorption on small metallic


clusters were performed to help understanding the first step of the FischerTropsch
reaction. An important step in the FischerTropsch synthesis consists of hydrogenation of CO on a metallic catalyst, leading to many liquid products, such as
gasoline, diesel and olefins [7].
Iron and cobalt are metals on which catalysts for the FischerTropsch process are
commonly based [813]. On the other hand, platinum clusters were studied due to
the high number of previous works where the PtCO interaction is investigated, and
several papers show different forms of CO adsorption on platinum clusters [1418].
Indeed, adsorption energies and preferential adsorption forms for CO adsorption on
platinum are strongly method and basis set dependent [19].
The chemisorption of carbon monoxide on iron was studied in details in some
works [20, 21]. These studies show that CO adsorbs mainly on a hollow position,
although atop adsorption can also occur, depending on the coverage of CO on the
surface [22]. The adsorption of CO on cobalt surfaces has also been examined and
most of these studies have focused on the (0001) surface of Co [23, 24]. The main
adsorption site on Co(0001) surface is at the top of a single cobalt atom [25].

Methodology
The model used for simulation of adsorption on catalytic surfaces is always a matter
of debate. Simulations using periodic boundary conditions usually reproduce
extended surfaces where impurities or defects may be included. On the other hand,
simulations of small clusters with some tenth or less atoms, although having the
intrinsic deficiency of producing models where atoms with highly unsaturated
coordination sphere are present, are also able to generate situations, which seems to
more appropriately reproduce the experimental environment. It is well established
that the catalytic ability of an active surface site is a local phenomenon [26, 27] and
may be described reasonably well with a small cluster. The local properties, such as
geometries and vibrational frequencies are well described by these models. They
can also be used as models for strongly dispersed catalysts. Based on these facts, we
decided to study the behavior of small clusters of each metal, and their ability to
adsorb and dissociate the CO molecule [28].
The platinum clusters were simulated using their (111) face. As a first approach
this might also be used to simulate the iron surface. However, the Fe(111) surface is
inactive for the FischerTropsch synthesis [29]. For this reason, the (110) face was
used in our studies. For the cobalt cluster the Co(0001) face was simulated. The
cobalt and the iron clusters used in this study have ten and eleven atoms,
respectively, arranged into two layers (Co7 ? 3 and Fe7 ? 4). The CoCo distances
and the shortest FeFe distances are 2.47 A
. The Pt clusters
in the cluster are 2.51 A
.
also have ten atoms arranged in two layers (Pt7 ? 3) with PtPt distances of 2.77 A
All calculations were done using the Density Functional approach. The hybrid
B3LYP functional as proposed and parameterized by Becke [30] was applied for
calculations of the platinum clusters. B3LYP is a mixture of HartreeFock and DFT
exchange terms with the gradient-corrected correlation functional of Lee et al. [31].

123

Modeling the Adsorption of CO on Small Pt, Fe and Co Clusters

603

For the iron and cobalt clusters the BPW91 functional was employed. The BPW91
functional uses the pure-exchange electron gas formula as the local density
approximation (LDA) with Becke gradient correction [30], and the correlation
correction by PerdewWang [32, 33]. Hays effective core potential [34] LanL2DZ
as denoted in Gaussian 03W [35], which include relativistic effects for the heavy
metals, was used for the transition metals, while the D95V basis set was used for the
carbon and oxygen atoms. In order to improve the quality of the calculated energy,
single-point energy calculations with the 6-31G(d) basis set replacing the D95V
basis set on the carbon and oxygen atoms were done for selected adsorption forms.
Additionally, adsorption energies were corrected for basis set superposition error
using the counterpoise correction method [36, 37]. Calculations were done using
both the restricted (for singlet spin states) as well as the unrestricted (for higher spin
states) formalism. The Gaussian 03W software was employed in all calculations.
In a recent publication [38] the lowest energy electronic state found for a Pt10
cluster was that with multiplicity 9 (S = 4). For the cobalt and iron clusters single
point calculations with several electronic states for different spin multiplicities were
carried out in order to determine the lowest energy electronic state. The calculations
showed a state with multiplicity 35 (S = 17) and 9 (S = 4) as the most stable states
for Fe11 and Co10, respectively.
The adsorption energy was used as the parameter to quantify the interaction
between the carbon monoxide molecule and the metallic surfaces. It was calculated
according to Equation 1:
EADS ECO=Cluster  ECO ECluster

Where EADS is the adsorption energy, ECO/Cluster is the energy of the CO ? cluster
ensemble and ECO and ECluster are the energies of the CO molecule and the cluster,
respectively.

Results and Discussion


Adsorption Forms and Adsorption Energy
PtCO System
For adsorption of CO on the Pt10 cluster we calculated the 7 adsorption forms shown
in Fig. 1. The corresponding adsorption energies and geometrical parameters are
given in Table 1.
The lowest energy adsorption forms were found for arrangements 1, 2 and 6,
which represent adsorption on bridge (1 and 2) and atop (6) positions, respectively.
The energy difference for adsorption on bridged and on atop positions is small, on
the order of 0.3 eV. This indicates that these arrangements may interconvert into
one another with low energy demand. From Table 1, it may also be seen that
adsorption on the central atom is not favored. This may be rationalized in terms of
the higher coordination number of the central atom, what reduces its ability to
interact with the carbon monoxide molecule.

123

604

R. de Souza Monteiro et al.

Fig. 1 Adsorption forms of CO on the Pt10 cluster

Table 1 Adsorption energies


(B3LYP/LANL2DZ) and
geometrical parameters for the
species shown in Fig. 1

Adsorption
form

Adsorption
Energy (eV)

PtC bond
)
length (A

CO bond
)
lengt (A

-3.11

1.998

1.208

-3.01

1.995

1.209

-2.13

2.041/2.092

1.225

-2.17

2.009

1.224

-1.65

1.887

1.172

-2.81

1.835

1.183

-1.82

2.010/1.997

1.207

Table 2 Adsorption Energies (eV) for the most stable adsorption forms on Pt10
Adsorption
form

B3LYP/ LANL2DZ
(S = 0)

UB3LYP/ LANL2DZ
(S = 4)

UB3LYP/ 6-31G(d)a
(S = 4)

BSSEb

-3.11

-2.56

-2.52

-2.06

-3.01

-2.17

-2.23

-1.75

-2.81

-2.66

-2.74

-2.12

6-31G(d) for carbon and oxygen, LANL2DZ for the Pt atoms

Adsorption energy after correction for basis set superposition error

For the adsorption forms 1, 2 and 6 single point energy calculations were done
with the multiplicity 9 (S = 4) and the 6-31G(d) basis set replacing the D95V basis
set on the carbon and oxygen atoms. The results are show in Table 2.

123

Modeling the Adsorption of CO on Small Pt, Fe and Co Clusters

605

A calculation with the unrestricted formalism reduces the adsorption energy in all
cases (because it stabilizes more strongly the bare cluster), the reduction being
higher for the bridged forms. Increasing the size of the basis set on the carbon and
oxygen atoms does not significantly changes the adsorption energy.
When the BSSE correction is applied the adsorption energy is further reduced to
a value substantially lower than that calculated without the BSSE correction. On
average the BSSE reduces the adsorption energy by about 0.5 eV. The new lowest
energy adsorption form is the atop form (6).
The experimental value for the adsorption energy of CO on platinum was
obtained by single crystal calorimetry and is about 1.90 eV [39]. Theoretical values
found in the literature show high dependence on the CO adsorption site. For linear
adsorption on atop sites, values from 1.55 to 1.95 eV were reported [39, 40]. For
adsorption on the bridge site adsorption energy of 1.82 eV was calculated [39],
while for 3-hollow adsorption sites values between 1.82 and 2.34 eV were obtained
[39, 40]. The results of the present work clearly show that it is crucial to consider
the electronic spin state in the calculation in order to obtain results comparable with
the experimental observation. This is a behavior similar to that found before for
adsorption of benzene on Pt10 cluster [38].
Most of the reported works on adsorption of CO on platinum were done using
periodic boundary conditions (PBC) with large CO coverage. In these simulations
the metal atoms have higher saturation and there are no acidic borders, with the
surface being much more stabilized than in the present study. Therefore, the values
given in Table 1, which in general are more negative than those, reported in the
literature, may be attributed to the much less saturated conditions of the atoms in the
clusters we simulated. In the small clusters the atoms in the border have high acidity
and therefore bind much more strongly to the carbon monoxide molecule than in the
case of more saturated atoms. However, the energy difference is clearly reduced
when correcting for both basis set superposition error (BSSE) and including a large
basis set on the carbon and oxygen atoms. This shows that these corrections are
necessary for correctly reproduce the experimental adsorption energy.
The preferential form for adsorption of CO on platinum has been a matter of
debate [41]. Some authors [18] show that the preferred adsorption forms of CO on
Pt surfaces are, in the order, atop, bridge and hollow. On low CO coverage,
however, the preferential site is the atop one [42]. In the present study the atop form
is the most stable, although with low energy difference for bridge forms. This trend
was also found by Gil et al [40], who observed that the B3LYP functional predicts
two adsorption sites (atop and fcc hollow) almost isoenergetic.
The CO distances for all adsorption forms (Table 1) show a little stretching
in the gas phase [43]. The largest CO
compared to the experimental value of 1.13 A
distances are calculated for the arrangements where the CO molecule is on a bridge
or hollow position, interacting with more than one platinum atom. The
corresponding C-metal distances are also longer for these arrangements than they
are for atop arrangements. This may be rationalized in terms of the electronic
interactions between the carbonyl group and the metal surface. A clear rehybridization occurs for adsorption in the bridge or hollow sites, leading the carbon atom
to adopt a nearly sp2 hybridization. On the other hand, for adsorption on atop site the

123

606

R. de Souza Monteiro et al.

hybridization of the carbon atom is more of a sp nature. As a consequence the


carbon-metal distances in the last case are smaller. The CO bond distance after
adsorption depends not only on carbon rehybridization, as discussed above, but also
on the electron back donation from the metal. It seems that when interacting with
two or more metal atoms back donation becomes more effective, leading to a more
pronounced increase in the CO bond distance. Therefore, while energies are more
strongly determined by the degree of saturation of the model atoms to which the CO
molecule interacts, the geometric parameters are more strongly influenced by
electronic reorganization.
FeCO System
The different adsorption forms of CO on the Fe11 cluster are given in Fig. 2. The
corresponding adsorption energies are given in Table 3.
For the Fe11CO system the lowest energy adsorption forms were found for
arrangements (7), (8) and (9), which represent adsorption on bridge, 3-hollow and
4-hollow sites, respectively. Indeed, in the case of adsorption of CO on the Fe11
clusters the adsorption energies are all very similar. Therefore, the CO mobility on
the Fe11 surface is still easier than on the Pt10 cluster.
Single point energy calculations with the 6-31G(d) basis set replacing the D95V
on the carbon and oxygen atoms and using the unrestricted formalism (S = 17) for
adsorption forms (7), (8), and (9) are given in Table 4, which also includes the
effect of the basis set superposition error on these forms. Increasing the basis set
size in the present case systematically reduces the adsorption energy by about 0.2
0.3 eV, making the different adsorption forms even more similar in terms of
adsorption energy. The BSSE in this case is negligible.

Fig. 2 Geometrical arrangements for adsorption of CO on the Fe11 cluster

123

Modeling the Adsorption of CO on Small Pt, Fe and Co Clusters


Table 3 BPW91/LANL2DZ
adsorption energies (eV) and
geometrical parameters for the
arrangements shown in Fig. 3

Adsorption
forms

Adsorption
Energy (eV)

607

FeC bond
)
length (A

CO bond
)
length (A

-1.61

1.803

1.207

-1.63

1.816

1.207

-1.73

1.796

1.208

-1.44

1.999

1.236

-1.56

2.002

1.221

-1.57

1.969

1.235

-1.74

1.975

1.247

-1.76

1.966

1.248

-1.77

1.967

1.248

Table 4 Adsorption Energies (eV) for the most stable adsorption forms on Fe11
Adsorption forms

UBPW91/ LANL2DZ

UBPW91/6-31G(d)a

BSSEb

-1.74

-1.50

-1.49

-1.76

-1.49

-1.47

-1.77

-1.49

-1.47

6-31G(d) for carbon and oxygen, LANL2DZ for the Fe atoms

Adsorption energy after correction for basis set superposition error

The experimental adsorption energy for CO on iron surfaces is 1.24 eV [10].


Adsorption energies of 1.95 eV for adsorption on atop position and of 1.701.90 eV
for adsorption on the bridge sites were calculated [10, 22] using periodic boundary
conditions. It seems that border effects are not so important for CO adsorption on
iron, as compared to adsorption on platinum, probably due to the higher acidity of
the iron surface.
The FeC and CO bond lengths are given in Table 3. Both the FeC and CO
distances found in this work are close to those reported in the literature [10, 22].
Regarding the CO and FeC distances the behavior is similar to that discussed
before for adsorption on the platinum clusters. Thus the CO distance is elongated
when compared to the CO distance in the gas phase, with adsorption on the 3- and
4-hollow sites leading to the most pronounced elongation. Similarly, the FeC bond
is longer for the bridge and 3- and 4-hollow site adsorption forms than for
adsorption on atop positions. The explanation for these geometrical deformations
may again follow the same rationalization given in the case of the adsorption on
platinum, with electronic reorganization being the main origin for geometrical
changes.
CoCO System
The several forms for CO adsorption on the Co10 cluster are shown in Fig. 3. The
corresponding adsorption energies are given in Table 5.

123

608

R. de Souza Monteiro et al.

Fig. 3 Geometrical arrangements for adsorption of CO on cobalt clusters (Co10)


Table 5 BPW91/LANL2DZ
adsorption energies and
geometrical parameters for the
arrangements shown in Fig. 3

Adsorption
forms

Adsorption
Energy (eV)

CoC bond
)
length (A

CO bond
)
length (A

-3.28

1.751

1.204

-2.63

1.760

1.205

-3.02

1.890

1.241

-2.80

1.908

1.243

Table 6 Adsorption energies (eV) for the most stable adsorption forms on Co10
Adsorption forms

BPW91/LANL2DZ

BPW91/6-31G(d)a

BSSEb

-3.28

-1.88

-1.59

-3.02

-1.70

-1.56

-2.80

-1.57

-1.20

6-31G(d) for carbon and oxygen, LANL2DZ for the Co atoms

Adsorption energy after correction for basis set superposition error

Our calculation shows that adsorption on a border cobalt atom (adsorption form
1) is the most stable, with adsorption energy of 3.28 eV. Once again single point
energy calculations with the 6-31G(d) basis set replacing the D95V on the carbon
and oxygen atoms and using the unrestricted formalism (S = 4) were done for
adsorption forms 1, 3, and 4 (Table 6). As observed for iron, increasing the basis set
size systematically reduces the adsorption energy, this time, however, by a
considerably higher value (up to 1.4 eV in the case of adsorption on the atop form
1). BSSE correction has also alarger effect than in the previous case, reducing the
adsorption energy on average by 0.3 eV.
Our final BSSE corrected value for Co adsorption on the Co10 cluster is 1.59 eV,
with very similar adsorption energies for adsorption forms 1 and 3. Other works
report values for this adsorption energy from 1.66 to 1.99 eV and from 1.60 to
1.65 eV, depending on the adsorption sites and CO coverage [12, 23]. It should be
added, however, that these adsorption energies were obtained with DFT approaches
using plane wave basis set. The experimental isosteric heat of adsorption is 1.33 eV
[23].
As shown in Table 5 both the CO and the CoC distances gradually increase as
the CO binding site changes from adsorption form 1 to 4. The changes in the CoC

123

Modeling the Adsorption of CO on Small Pt, Fe and Co Clusters

609

distances are somewhat larger than in the CO distances. The calculated CO and
CoC bond distances for the most stable arrangement are in good agreement with
and
experimental and previous theoretical data (CO bond length is 1.17 0.06 A

the CoC distance is 1.78 0.06 A) [23, 25, 44]. Once again it is observed that
both the CoC and CO distances increase with the degree of coordination of the
carbon atom.
CO Dissociation on the Metallic Clusters
Figure 4 illustrates the relationship between the CO bond distances and the
adsorption energies for the most stable Pt, Fe and Co clusters. This picture shows a
converse relationship between the CO bond distance and the adsorption energy. As
a first approach it could be expected that stronger adsorption energies would lead to
higher geometrical changes in the CO structure, although, as shown above, the CO
bond elongation seems to be more dependent on the adsorption forms than on
adsorption energy. Therefore, it seems that there is no direct correlation between the
adsorption energy and the elongation in the CO bond, a result that has already been
found before [45]. Also, no correlation was found between the CO stretching
frequency, which should qualitatively be related to the CO bond length, and the
adsorption energy. Therefore, the very important step of CO dissociation on the
metal surfaces should depend on other factors, in addition to the adsorption energy.
We, therefore, calculated the activation energy for CO dissociation on the Pt, Fe and
Co clusters.
The dissociation of carbon monoxide on the Pt, Fe, and Co clusters was done
following the pathways showed in Fig. 5, as proposed elsewhere [10, 46, 47], and
Fig. 6 showed the structure after CO dissociation.
For the dissociation on the Pt10 cluster we moved the oxygen atom to the next
neighbouring Pt atom on the border of the cluster, as calculations on several forms

2.2
Pt

Adsoprtion Energy (eV)

2.1
2.0
1.9
1.8
1.7
1.6
1.5
1.4

Co
Fe

1.18 1.19 1.20 1.21 1.22 1.23 1.24 1.25 1.26


C-O bond distance (A)

Fig. 4 Correlation between the calculated CO bond length (dC-O) and adsorption energies

123

610

R. de Souza Monteiro et al.

Fig. 5 Pathways for CO dissociation on the a Pt10, b Fe11, and c Co10 clusters

Fig. 6 Structure after CO dissociation on the a Pt10, b Fe11, and c Co10 clusters

123

Modeling the Adsorption of CO on Small Pt, Fe and Co Clusters

611

Fig. 7 Energy profile for CO dissociation on platinum, iron and cobalt clusters

for dissociated CO on the Pt10 cluster revealed that the most stable arrangement is
that where both carbon and oxygen atoms are on bridged positions involving
neighbouring Pt atoms on the border of the cluster. Two different reaction pathways
were proposed for the dissociation of CO on the closed packed Co(0001) surface
[47]. In path I, CO dissociates onto a neighbouring equivalent hollow sites, resulting
in a transition state which occupies two adjacent bridge sites of the close-packed Co
atoms. In path II, CO dissociates over the top of a surface Co atom. In this work we
calculated only path II (Figs. 5 and 6). This choice was based on the calculated
adsorption energy, which is stronger for adsorption on atop position than on bridge
position. The same pathway was followed for dissociation on the Fe11 cluster.
Figure 7 shows the profile for CO dissociation on the platinum, iron and cobalt
clusters. As shown the activation energy for CO dissociation follows the order
Fe \ Co \ Pt, in agreement with the order of elongation in the CO bond length
(Fe [ Co [ Pt, see Fig. 4). This picture shows that dissociation on the lighter
transition metal iron involves lower activation energy. In contrast, the heavier
platinum induces much higher activation energy for CO dissociation.
Figure 8 shows a correlation between the activation energy and the CO bond.
It clearly indicates that larger activation energies are associated to those cases
where the CO bond distance stretches the least. Therefore higher activation energy
are calculated for dissociation on Pt10, where the CO bond distance is closer to the
value for free CO. Similarly, lower activation energy is calculated for dissociation
on Fe11, where the CO bond is more stretched.
Figure 9 shows a plot of the HOMO and LUMO energies for the three clusters
and the CO molecule.
Rationalized in terms of valence bond theory, the mechanism for catalytic CO
dissociation on a metal surface involves two steps after adsorption, M?CO charge
transfer from filled d orbitals on the metal to the 2p* orbital on CO, followed by CO
dissociation with electron transfer back to the metal [11]. Existence of low energy
unoccupied 2p* molecular orbital on CO permits interaction of the adsorbed
molecule as a Lewis p acid with the metal surface. Similarly, the presence of higher
energy filled d orbitals on the metal helps electron transfer to the antibonding 2p*
orbital on CO. The HOMO energies of the iron and cobalt clusters are higher than

123

612

R. de Souza Monteiro et al.

Activation Energy (eV)

5.5

Pt

5.0
4.5
4.0

Co

3.5
3.0

Fe

2.5
1.17 1.18 1.19 1.20 1.21 1.22 1.23 1.24 1.25 1.26
C-O distance (A)
Fig. 8 Correlation between the calculated CO bond length (dC-O) and the activation energy for CO
dissociation

HOMO and LUMO energies (eV)

-0.05
-0.10
-0.15
-0.20
-0.25
-0.30
-0.35
Fe

Co

Pt

CO

Fig. 9 HOMO and LUMO energies for the Pt10, Fe11 and Co10 clusters and the CO molecule

that of the platinum cluster, therefore helping charge transfer from the metal surface
to the p* orbital on CO. This phenomenon may be a consequence of the decrease in
the energies of the metal d orbitals due to its fulfilling. Therefore the lighter
transition metals, which have higher energy occupied orbitals, may more efficiently
activate the CO molecule for dissociation, via electron transfer to the empty
antibonding 2p* orbital on CO.

Conclusions
We simulated the first steps of the FischerTropsch synthesis by means of molecular
modeling procedures calculating the several CO adsorption forms on small metallic

123

Modeling the Adsorption of CO on Small Pt, Fe and Co Clusters

613

Pt, Fe and Co clusters. In our final model the preferential CO adsorption form on the
Pt10 cluster is on atop position, however only 0.06 eV more stable than for
adsorption on a bridge position. For adsorption on the Fe11 cluster we found bridge
and four hollow sites as almost isoenergetic, while for adsorption on the Co10 cluster
atop and three hollow sites were found as preferential and for the Pt11 cluster is only
in a atop position. The CO adsorption energy is higher for adsorption on Pt than on
either Fe or Co.
Calculations of several adsorption arrangements led to estimation of the energy
for the MCO interaction. For platinum and iron clusters, the most favorable
interactions are those on bridge and 4-hollow sites, respectively. Atop position was
the most stable one for the cobalt model studied in this work. The increase CO
bond distance at the transition state indicates that the barrier for CO dissociation is a
late transition state. Late transition state has been demonstrated on Pt surfaces in
recent publication [42].
In addition, the adsorption energy does not affect the CO bond distance, probaly
the higher influence is on the Fermi level. A variation in Eads therefore reflects
changes in the overall stability of the adsorbatesubstrate complex. In contrast to
this the CO, CMetal distances and the CO stretch frequency are sensitive
measures of the local bond characteristics.
Dissociation energies for CO on clusters were also estimated. The values found
for this parameter on iron and cobalt clusters were 1.01.5 eV smaller than those
found for the dissociation of carbon monoxide on platinum.
Acknowledgments We gratefully acknowledge financial support from FAPERJ (grant 26/152.786/
2006), ANP, CNPq (grant 47611/2006-0), and Petrobras (Projeto GTL).

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.

H. Sellers and J. Gilsaon (1999). Surf. Sci. 426, 147.


I. M. Ciobica and R. A. van Santen (2003). J. Phys. Chem. B. 107, 3808.
F. Abild-Pedersen and M. P. Andersson (2007). Surf. Sci. 601, 1747.
W. Liu, Y. F. Zhu, J. S. Lian, and Q. Jiang (2007). J. Phys. Chem. C. 111, 1005.
P. Liu and J. K. Norskov (2001). Phys. Chem. Chem. Phys. 3, 3814.
M. Kubo, T. Kubota, C. Jung, M. Ando, S. Sakahara, K. Yajima, K. Seki, R. Belosludov, A. Endou,
S. Takami, and A. Miyamoto (2004). Catal Today. 89, 479.
C. Wang, L. Xu, and Q. Wang (2003). J. Natural Gas Chem. 12, 10.
J. Gaube and H.-F. Klein (2008). J. Mol. Catal. A: Chem. 283, 60.
H. D. Burtron (2007). Ind. Eng. Chem. Res. 46, 8938.
D. E. Jiang and E. A. Carter (2004). Surf Sci. 570, 167.
A. C. Pavao, T. C. F. Guimaraes, S. K. Lie, C. A. Taft, and W. A. Lester (1999). J. Mol. Struct.
(Teochem). 458, 99.
X.-Q. Gong, R. Raval, and P. Hu (2005). J. Chem. Phys. 122, 24711.
R. V. Belosludov, S. Sakahara, K. Yajima, S. Takami, M. Kubo, and A. Miyamoto (2002). App. Surf.
Sci. 189, 245.
I. Grinberg, Y. Yourdshahyan, and A. M. Rappe (2002). J. Chem. Phys. 117, 2264.
N. A. Besley (2005). J. Chem. Phys. 122, 184706.
D. Geschke, T. Bastug, T. Jacob, S. Fritzsche, W.-D. Sepp, and B. Fricke (2001). Phys Rev B. 64,
235411.
P. J. Feibelman, B. Hammer, J. K. Nrskov, F. Wagner, M. Scheffler, R. Stumpf, R. Watwe, and
J. Dumesic (2001). J. Phys. Chem. B. 105, 4018.

123

614

R. de Souza Monteiro et al.

18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.

K. Doll (2004). Surf. Sci. 573, 464.


S. A. Wasileski, M. T. M. Koper, and M. J. Weaver (2001). J. Phys. Chem. B. 105, 3518.
S. P. Mehandru and A. B. Anderson (1998). Surf. Sci. 201, 345.
Y. Yang, H. Xiang, Y. Xu, L. Bai, and Y. Li (2004). Appl. Catal. A. 266, 181.
A. Stibor, G. Kresse, A. Eichler, and J. Hafner (2002). Surf. Sci. 507, 99.
M. Gajdos, A. Eichler, and J. Hafner (2004). J. Phys. Condens. Matter. 16, 1141.
Gong X-Q, R. Raval, and P. Hu (2004). Surf. Sci. 562, 247.
H. Papp (1983). Surf. Sci. 129, 205.
M. Neurock (2003). J. Catal. 216, 73.
J.-R. Chang, S.-L. Chang, and T.-B. Lin (1997). J. Catal. 169, 338.
D. Curulla, A. Clotet, J. M. Ricart, and F. Illas (1999). J. Phys. Chem. B. 103, 5246.
C. Rong and C. Satoko (1989). Surf. Sci. 223, 101.
A. D. Becke (1992). J. Chem. Phys. 96, 2155.
C. Lee, W. Yang, and R. G. Parr (1988). Phys. Rev. B. 37, 785.
J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh, and C. Fiolhais
(1992). Phys. Rev. B. 46, 6671.
J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh, and C. Fiolhais
(1993). Phys. Rev. B. 48, 4978.
P. J. Hay and W. R. Wadt (1985). J. Chem. Phys. 82, 299.
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A.
Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V.
Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M.
Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai,
M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R.
Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y.
Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A.
D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J.
V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P.
Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J. A.
Pople (2004). Gaussian, Inc., Wallingford, CT.
S. Simon, M. Duran, and J. J. Dannenberg (1996). J. Chem. Phys. 105, 11024.
S. F. Boys and F. Bernardi (1970). Mol. Phys. 19, 553.
M. T. M. Cruz, J. W. M. Carneiro, D. A. G. Aranda, and M. Buhl (2007). J. Phys. Chem. C. 111,
11068.
H. Orita, N. Itoh, and Y. Inada (2004). Surf. Sci. 571, 161.
A. Gil, A. Clotet, J. M. Ricart, G. Kresse, M. G. Hernandez, N. Rosch, and P. Sautet (2003). Surf. Sci.
530, 71.
I. Grinberg, Y. Yourdshahyan, and A. M. Rappe (2002). J. Chem. Phys. 117, 2264.
J. V. Nekrylova, C. French, A. N. Artsyukhovich, V. A. Ukraintsev, and I. Harrison (1993). Surf. Sci.
Lett. 295, L987.
G. Herzberg, Electronic Spectra of Polyatomic Molecules, Van Nostrand, NY (1966).
S. J. Jenkins and D. A. King (2002). Surf. Sci. 504, (2002), 138.
T. E. Shubina and M. T. M. Koper (2002). Electrochim. Acta. 47, 3621.
Y. Morikawa, J. J. Mortensen, B. Hammer, and J. K. Nrskov (1997). Surf. Sci. 386, 67.
Q. Ge and M. Neurock (2006). J. Phys. Chem. B. 110, 15368.

33.
34.
35.

36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.

123

Você também pode gostar