You are on page 1of 5

Catalysis Today 263 (2016) 1115

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Importance of Pd monomer pairs in enhancing the oxygen reduction


reaction activity of the AuPd(1 0 0) surface: A rst principles study
Hyung Chul Ham a,b, , Gyeong S. Hwang a, , Jonghee Han b , Sung Pil Yoon b ,
Suk Woo Nam b , Tae Hoon Lim b
a
b

Department of Chemical Engineering, The University of Texas at Austin, Austin, Texas, 78712, USA
Fuel Cell Research Center, Korea Institute of Science and Technology (KIST), Seoul, Korea

a r t i c l e

i n f o

Article history:
Received 2 May 2015
Received in revised form 15 July 2015
Accepted 30 July 2015
Available online 21 August 2015
Keywords:
AuPd(1 0 0)
Ensembles
ORR
First-principles

a b s t r a c t
Based on density functional theory calculations, we present that pairs of 1st nearest Pd monomers play an
important role in signicantly enhancing the oxygen reduction reaction (ORR) on the AuPd(1 0 0) surface.
While the catalytic ORR activity tends to be sensitive to the surface atomic ordering, we nd that the Pd
monomer pairs lead to a substantial reduction in the activation barrier for O/OH hydrogenation with no
signicant suppression of OO bond scission, thereby considerably lowering the overall activation energy
for the ORR as compared to the case of pure Pd(1 0 0). On the other hand, an isolated Pd monomer tends
to greatly suppress the OO bond cleavage reaction, which in turn slows down the ORR kinetics. Unlike
the monodentate adsorption of O2 on an isolated Pd monomer, the pairing of Pd monomers allows O2
adsorption in a bidentate conguration and consequently facilitating OO bond scission. However, the
barrier for OH hydrogenation at each Pd site shows no signicant change between the isolated and paired
cases, while it is noticeably lower than the pure Pd case.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Bimetallic gold-palladium (AuPd) catalysts have been tested and
used for various chemical reactions such as production of vinyl
acetate monomer [1], low temperature oxidation of carbon monoxide [25], selective oxidation of formic acid [68], and selective
production of hydrogen peroxide (H2 O2 ) from molecular O2 and H2
[6,911]. Recent studies suggest that the surface activity of AuPd
is governed by the various alloying effects such as the so-called
electronic (ligand) effects [4,6,1215] (electronic state change by
metalmetal interactions) and ensemble (geometric) effects (modication of catalytic activity by the unique arrangement of surface
atoms) [1620]. However, the underlying mechanism of AuPd alloy
catalysis still remains unclear.
It has been theoretically and experimentally found that an isolated Pd monomer surrounded by Au atoms is a good site for
increasing the rate of O/OH hydrogenation reaction but signicantly slowing down the kinetics of O2 bond cleavage, thereby
yielding the signicant enhancement of selective H2 O2 formation

Corresponding author. Tel.: +82 2 958 5889; fax: +82 2 958 5199.
Corresponding author. Tel.: +1 512 471 4847; fax: +1 512 471 7060.
E-mail addresses: hchahm@kist.re.kr, ham.hyungchul@gmail.com (H.C. Ham),
gshwang@che.utexas.edu (G.S. Hwang).
http://dx.doi.org/10.1016/j.cattod.2015.07.054
0920-5861/ 2015 Elsevier B.V. All rights reserved.

[2125]. The reactivity of Pd monomer in the AuPd catalyst can be


also altered due to additional geometrical factors. For instance, it
tends to be changed by the presence of low-coordination surface
atoms and the strain imposed on the outer-layer atoms, in association with the size and shape of nanoparticle catalysts and the
lattice parameter mismatch between the substrate and the adlayer
[13,2628]. The different surface facets such as (1 1 1) and (1 0 0)
[29] may also affect the reactivity of isolated Pd monomers, leading
probably to the change in catalysis of the oxygen reduction reaction (ORR). For example, according to a recent density functional
theory (DFT) calculation [23], ORR strongly depends on the surface arrangement of Au and Pd atoms Pd(1 1 1). In particular, on Pd
monomers surrounded by less active Au atoms, H2 O2 is selectively
produced, while on Pd dimer, the reactivity to H2 O is predicted to
be enhanced compared to pure Pd(1 1 1). But there has been little
investigation concerning the interplay between different facet and
Pd ensemble.
In this paper, we investigate the arrangement of Pd atoms
and its effect on ORR on the AuPd(1 0 0) surface, particularly the
role of pairs of 1st nearest Pd, using periodic density functional
theory (DFT) calculations. Here, we only consider small-sized Pd
ensembles including monomer, monomer pair, and dimer which
are experimentally found to be energetically favorable when
the surface coverage of Pd is sufciently low [30,31]. We rst
examine the relative stabilities among those small Pd ensembles

12

H.C. Ham et al. / Catalysis Today 263 (2016) 1115

Fig. 1. Tilted side view of the model PdAu surfaces considered in this work, an isolated Pd monomer (indicated by 1M), a pair of 1st nearest monomers (1st 2MP), a dimer (D),
a trimer (T), a tetramer (TE) and a Pd(1 0 0) slab (P). The green, gold, and gray balls represent surface Pd, surface Au, and subsurface Pd atoms, respectively. (For interpretation
of the color information in this gure legend, the reader is referred to the web version of the article.)

considered and also calculate and compare the reaction energetics


and barriers for O-O bond scission and O/OH hydrogenation on the
different Pd sites. Next, we analyze the surface electronic structure
modications upon O2 adsorption to better understand the reasons underlying the activity enhancement of pairs of 1st nearest Pd
monomers toward O-O scission.

become smaller than 5 102 eV/. The lattice constant for bulk
which is virtually identical to previous
Pd is predicted to be 3.95 A,
DFT-GGA calculations and also in good agreement with the exper The binding energy (Eb ) of O2 is calculated
imental value of 3.89 A.
by Eb = E(M) + E(O2 ) E(O2 /M), where E(O2 /M), E(M), and E(O2 ) represent the total energies of the O2 /slab system, the bare slab, and
an isolated triplet O2 molecule in the gas state, respectively.

2. Computational methods
3. Results and discussion
The calculations reported herein were performed on the basis
of spin polarized DFT within the generalized gradient approximation (GGA-PW91) [32], as implemented in the Vienna Ab-initio
Simulation Package (VASP) [33]. The projector augmented wave
(PAW) method [34] with a plane-wave basis set was employed to
describe the interaction between core and valence electrons. The
valence congurations employed to construct the ionic pseudopotentials are: 5d10 6s1 for Au, 4d9 5s1 for Pd, and 2s2 2p4 for O. An
energy cut-off of 350 eV was applied for the planewave expansion
of the electronic eigenfunctions. For Brillouin zone sampling, we
used a (3 3 1) Monkhorst-Pack mesh of k points to determine
the optimal geometries and total energies of systems examined,
and increased the k-point mesh size up to (8 8 1) to re-evaluate
corresponding electronic structures. Reaction pathways and barriers were determined using the climbing-image nudged elastic
band method (c-NEBM) [35] with eight intermediate images for
each elementary step.
For a model surface, we used a supercell slab that consists of a
rectangular 4 4 surface unit cell with four atomic layers each of
which contains 16 atoms. For each AuPd surface model, the topmost
surface layer that is overlaid on a three-layer Pd (1 0 0) slab contains
selected Pd ensembles including an isolated Pd monomer (indicated by 1M), a pair of 1st nearest monomers (1st 2MP), a dimer
(D), a trimer (T), and a tetramer (TE) (see Fig. 1). A slab is separated from its periodic images in the vertical direction by a vacuum
space corresponding to seven atomic layers. While the bottom two
layers of the four-layer slab were xed at corresponding bulk positions, the upper two layers were fully relaxed using the conjugate
gradient method until residual forces on all the constituent atoms

3.1. Relative stability of small Pd ensembles


Before looking at the reactivity of various Pd ensembles toward
ORR, we rst examined the relative stability of Pd ensembles on the
(1 0 0) surface. Table 1 summarizes the predicted formation energies per Pd atom for the 1M, 1st 2MP, D, T, and TE ensembles,
given by Ef = [EPdAu EAu + NPd (EAu-bulk EPd-bulk )]/NPd , where
EPdAu , EAu , EAu-bulk , and EPd-bulk represent the total energies of
PdAu/Pd(1 0 0) [Pd ensembles are overlaid on a three-layer Pd
(1 0 0) slab], Au/Pd(1 0 0) [Au monolayer are overlaid on a threelayer Pd (1 0 0) slab], bulk Au (per atom), and bulk Pd (per atom),
respectively, and NPd indicates the number of Pd atoms on a given
PdAu surface. Here, the smaller the formation energy, the higher
the stability of Pd ensembles.
The calculation results suggest that 1M, D, T, and TE in order
of decreasing stability; that is, Pd atoms would have a tendency
to remain isolated, rather than form aggregates in the AuPd (1 0 0)
surface, consistent with recent theoretical and experimental results
[30,36,37]. In addition, our calculations show that the Ef of 1st 2MP
is comparable to that of 1M, implying that 1st 2MP can exist stably
for surface catalytic reactions. Hence, we only considered the 1M,
1st 2MP, and D sites for the study of O-O bond scission and O/OH
hydrogenation, with comparison to the pure Pd(1 0 0) case.
3.2. O-O bond scission
In this study, we examined the three elementary steps
(the so-called dissociative mechanism) for the ORR in acidic

H.C. Ham et al. / Catalysis Today 263 (2016) 1115

13

Table 1
Calculated formation energies (in eV) of various Pd ensembles on AuPd(1 0 0) surface. The green and yellow represent the Pd and Au atoms, respectively.
1M

1st 2MP

TE

0.39

0.39

0.42

0.42

0.43

condition: (A) O2 adsorption [O2 (g) O2 ], (B) O2 scission


[O2 O + O], (C) O hydrogenation [O + H OH], and (D) OH hydrogenation [OH + H H2 O(g)]. Although the ORR is complicated and
its detailed mechanism still remains under debate, the comparisons
of these scission and hydrogenation reactions would give important
insight into the activity of different ensemble sites toward the ORR
[23,38,39].
Table 2 summarizes the predicted total energy changes (E) and
activation barriers (Ea ) for O-O bond scission and O/OH hydrogenation reactions. The results clearly show that 1st 2MP signicantly
enhances the ORR as compared to pure Pd(1 0 0), while for the 1M
and D cases the opposite is true. Next, we will discuss each reaction
step in detail.
The O2 binding energies (Eb , with respect to the triplet ground
state of gas phase O2 ) of small Pd ensembles are predicted to
be much smaller as compared to the pure Pd(1 0 0) case, i.e., P
(1.73 eV) > D (1.10 eV) > 1st 2MP (0.60 eV) > 1M (0.48 eV). It is worth
pointing out that 1st 2MP has a noticeably higher Eb (by 0.12 eV)
than 1M. According to our Bader charge analysis, the amount of
charge transferred to the adsorbed O2 from the surface is much
greater in 1st 2MP (0.63e ) than in 1M (0.32e ); consequently, the
O-O bond is more elongated (1.39 A in 1st 2MP and 1.29 A in 1M)
as the transferred charge lls the antibonding O2 2p states.
The O2 scission reaction is predicted to be exothermic by 0.52 eV
in 1st 2MP, whereas endothermic by 0.02 eV in 1M. Our calculations
also predict a substantially lower barrier for O-O bond cleavage in
1st 2MP (0.48 eV) than the 1M case (1.04 eV). Note that the large
barrier for the O2 scission reaction in 1M causes an increase in
the formation of H2 O2 rather than H2 O. According to the reaction
energetics/barriers in the associative mechanism (see Table S1 in
Supporting Information), the selectivity for H2 O2 formation is the
highest in 1M.
The enhanced kinetics of O-O bond cleavage on 1st 2MP over 1M
is related to the unique arrangement of Pd monomers on the (1 0 0)
surface. As illustrated in Fig. 2, O2 on 1st 2MP can be stabilized in the
bridge-hollow-bridge (b-h-b) conguration by being bonded to two
Pd atoms, indicating that both Pd atoms in 1st 2MP are involved in
O2 adsorption. This is possible because the Pd-Pd distance of 3.95 A
in 1st 2MP is sufciently short that O2 adsorbed can be in a bidentate conguration. Note that the distance between two nearest Pd
such that
monomers in the (1 1 1) surface is much longer (4.83 A)
O2 prefers to adsorb onto a single Pd atom in a monodentate conguration with Eb of 0.25 eV. Moreover, the O atoms in the transition
and nal states on 1st 2MP can be signicantly stabilized by being
linked to two Pd atoms, leading to a reduction in the activation
energy barrier compared to the 1M case (Figs. 2 and 3).
The enhanced O2 activation on 1st 2MP can also be demonstrated by the analysis of the local density of states (LDOS) projected
on the 2p orbital of O2 adsorbed on a single Pd monomer (1M), a pair
of 1st nearest monomers (1st 2MP), a dimer (D), a trimer (T) and a
pure Pd(1 0 0) (P) (Here, the LDOS of a tetramer (TE) is not shown
because it has a similar LDOS to a trimer case). As displayed in Fig. 4,

Fig. 2. Predicted molecular congurations of the initial, transition and nal states
in the O-O bond scission and O/OH hydrogenation reactions at the 1st 2MP site. Red,
yellow, green, and white balls indicate O, Au, Pd, and H atoms, respectively. (For
interpretation of the color information in this gure legend, the reader is referred
to the web version of the article.)

Fig. 3. Predicted potential energy diagram on various Pd ensembles for


O2 (g) + 2H2 (g) 2H2 O(g) reaction.

14

H.C. Ham et al. / Catalysis Today 263 (2016) 1115

Table 2
Calculated total energy changes (E) and activation barriers (Ea in parenthesis) for O-O bond scission and O/OH hydrogenation reactions on various Pd ensembles with
respect to fully separated co-adsorbed species. The symbol P indicates the pure Pd(1 0 0) slab. The All energy values are given in eV.

(A) O2 (g) O2
(B) O2 O + O
(C) O + H OH
(D) OH + H H2 O(g)

1M

1st 2MP

Pt(1 1 1)

0.48 ()
0.02 (1.04)
1.40 (0.34)
0.67 (0.68)

0.60 ()
0.52 (0.48)
0.92 (0.48)
0.54 (0.63)

1.10 ()
0.54 (0.89)
0.83 (0.71)
0.05 (1.03)

1.73 ()
0.89 (0.15)
0.55 (0.54)
0.14 (0.95)

0.83 ()
1.77 (0.56)
0.00 (1.06)
0.30 (0.34)

Fig. 4. Local density of states projected onto the 2p orbital of O2 adsorbed on a single
Pd monomer (1M), a pair of 1st nearest monomers (1st 2MP), a dimer (D), a trimer
(T) and a pure Pd(1 0 0) (P). The 2p orbital of gas state O2 is also presented. The Fermi
level (Ef ) is marked by the vertical line.

for the Pd monomer pairs (1st 2MP), trimer (T) and pure Pd(1 0 0)
cases, the spin-up and spin-down states are nearly degenerate with
a magnetic moment of  0.02, indicating the formation of peroxolike O2 by completely lling the antibonding 2p orbital (g* ) of
adsorbed O2 by the charge transfer from the surface. Here, O2 is
preferentially adsorbed in the bridge-hollow-bridge (b-h-b) conguration. On the contrary, for the single Pd monomer and dimer
cases, we predict the signicant magnetic moment of  0.99 (1M)
and  0.50 (D), implying the superoxo-like character of the bound
O2 by partially lling antibonding 2p orbital (g* ) by less charge
transfer from surface than the 1st 2MP, T and P cases. Here, O2
is preferentially adsorbed in the top-bridge-top (t-b-t) conguration. This demonstrates that both Pd atoms in a pair of 1st nearest
monomers (1st 2MP) can play an key role in signicantly enhancing
the O2 activation on AuPd(1 0 0) surface.
3.3. O/OH hydrogenation
Looking at O/OH hydrogenation, our calculations predict very
noticeable enhancement in the kinetics on 1M (0.34/0.68 eV) and
1st 2MP (0.48/0.63 eV) when compared to the pure Pd(1 0 0) surface (0.54/0.95 eV), while the O/OH hydrogenation barriers tend
to increase on D (0.71/1.03 eV). This suggests that Pd monomers
in both 1M and 1st 2MP can contribute to facilitating the O/OH
hydrogenation reactions; that is, the monomer pair can behave
like an isolated monomer for O/OH hydrogenation. It would be also
interesting to note that the rate-limiting step on each ensemble
may change considering the signicant differences in the activation
barriers for the O2 scission and OH hydrogenation reactions.
First, we nd that for the Pd(1 0 0) surface the OH hydrogenation reaction is rate-limiting with a barrier of 0.95 eV, while for the
Pt(1 1 1) surface, the O hydrogenation reaction is rate-determining
step with a barrier of 1.06 eV, indicating the improved ORR kinetics
in Pd(1 0 0) over Pt(1 1 1). This agrees well with a recent experimental result where the ORR activity in Pd(1 0 0) is higher than Pt(1 1 1)
case [40].
Next, looking at Pd ensembles cases, we nd that the OH hydrogenation reaction is rate-limiting on 1st 2MP (Ea = 0.63 eV), and
D (1.03 eV), whereas the O2 scission reaction is on 1M (1.04 eV).

Considering also the activation barriers, we could expect that the


kinetics of ORR can be signicantly enhanced on 1st 2MP compared
to pure Pd(1 0 0), while the opposite is true for the 1M and D cases.
The kinetics O/OH hydrogenation can be correlated to the relative binding strengths of OH/O/H2 O. Our calculations show no
signicant variation in the binding energy of OH at the bridge
sites of 1M (Eb = 2.41 eV) and 1st 2MP (2.40 eV) and H2 O at the
top sites 1M (0.31 eV) and 1st 2MP (0.32 eV). For O adsorption,
the hollow site in 1st 2MP (Eb = 3.99 eV) tends to be energetically
more favorable than that in 1M (Eb = 3.66 eV). As such, the predicted exothermicties for the hydrogenation reactions are smaller
in 1st 2MP than 1M (see Table 2). Overall, our study clearly highlights that the unique arrangement of Pd monomers on the (1 0 0)
surface, especially 1st 2MP, can contribute to signicantly enhancing the ORR by facilitating O-O bond scission without substantial
suppression of O/OH hydrogenation.
4. Conclusion
We present the impact of pairs of 1st nearest Pd monomers on
the catalytic ORR activity of the AuPd(1 0 0) surface based on DFT
calculations. Our study suggests that the ORR strongly depends on
the distribution of Pd atoms in the AuPd(1 0 0) surface. Especially,
the pair of 1st nearest Pd monomers can greatly enhance the kinetics of ORR compared to the pure Pd(1 0 0) surface by substantially
reducing the kinetics of the rate-determining OH hydrogenation
reaction without the signicant suppression of the O-O bond scission. However, for the single Pd monomer case, the barrier for O-O
bond scission is largely increased, which results in the reduction of
ORR kinetics. This activity enhancement of the pair of 1st nearest
Pd monomers over a single Pd monomer toward ORR is due to the
large stabilization of O2 /O by being linked to two Pd atoms in the
course of O-O bond breaking reaction and no signicant variation
of OH binding energies in O/OH hydrogenation reaction. Here, OH
is adsorbed at the bridge site by being bonded with one Pd and one
Au atom rather than at the hollow site by being connected with two
Pd atoms. This study also highlights that, when designing the AuPd
nanocatalysts, the preferential exposure of (1 0 0) facet in surface
can be one of key requirements to enhancing the ORR.
Acknowledgement
This work was supported by the KIST institutional program
of Korea Institute of Science and Technology (2E25412) and the
R.A. Welch Foundation (F-1535). The authors also thank the Texas
Advanced Computing Center for use of their computing resources.
Appendix A. Supplementary data
Supplementary material related to this article can be found,
in the online version, at http://dx.doi.org/10.1016/j.cattod.2015.07.
054
References
[1] S. Venkatachalam, T. Jacob, Phys. Chem. Chem. Phys. 11 (2009) 32633270.
[2] J.A. Boscoboinik, C. Plaisance, M. Neurock, W.T. Tysoe, Phys. Rev. B 77 (2008).

H.C. Ham et al. / Catalysis Today 263 (2016) 1115


[3] H.C. Ham, J.A. Stephens, G.S. Hwang, J. Han, S.W. Nam, T.H. Lim, J. Phys. Chem.
Lett. 3 (2012) 566570.
[4] V. Pallassana, M. Neurock, L.B. Hansen, B. Hammer, J.K. Norskov, Phys. Rev. B
60 (1999) 61466154.
[5] J. Zhang, H.M. Jin, M.B. Sullivan, F. Chiang, H. Lim, P. Wu, Phys. Chem. Chem.
Phys. 11 (2009) 14411446.
[6] N.N. Edwin, J.K. Edwards, A.F. Carley, J.A. Lopez-Sanchez, J.A. Moulijn, A.A.
Herzing, C.J. Kiely, G.J. Hutchings, Green Chem. 10 (2008) 11621169.
[7] D.A. Bulushev, S. Beloshapkin, J.R.H. Ross, Catal. Today 154 (2010) 712.
[8] H. Ohtani, M.A. Vanhove, G.A. Somorjai, Surf. Sci. 187 (1987) 372386.
[9] B.E. Solsona, J.K. Edwards, P. Landon, A.F. Carley, A. Herzing, C.J. Kiely, G.J.
Hutchings, Chem. Mater. 18 (2006) 26892695.
[10] J.K. Edwards, A.F. Carley, A.A. Herzing, C.J. Kiely, G.J. Hutchings, Faraday
Discuss. 138 (2008) 225239.
[11] C. Samanta, Appl. Catal. A-Gen. 350 (2008) 133149.
[12] J. Kitchin, J. Nrskov, M. Barteau, J. Chen, Phys. Rev. Lett. 93 (2004).
[13] H.C. Ham, G.S. Hwang, J. Han, S.W. Nam, T.H. Lim, J. Phys. Chem. C 114 (2010)
14922.
[14] R. Hoffmann, Rev. Modern Phys. 60 (1988) 601628.
[15] J.A. Stephens, H.C. Ham, G.S. Hwang, J. Phys. Chem. C 114 (2010)
2151621523.
[16] J.P. Perdew, Y. Wang, Phys. Rev. B 45 (1992) 1324413249.
[17] P.B. Balbuena, S.R. Calvo, E.J. Lamas, P.F. Salazar, J.M. Seminario, J. Phys. Chem.
B 110 (2006) 1745217459.
[18] J.A. Rodriguez, P. Liu, J. Hrbek, J. Evans, M. Perez, Angew. Chem. Int. Ed. 46
(2007) 13291332.
[19] P. Liu, J.K. Norskov, Phys. Chem. Chem. Phys. 3 (2001) 38143818.
[20] F. Maroun, F. Ozanam, O.M. Magnussen, R.J. Behm, Science 293 (2001)
18111814.

15

[21] F. Pittaway, L.O. Paz-Borbon, R.L. Johnston, H. Arslan, R. Ferrando, C. Mottet, G.


Barcaro, A. Fortunelli, J. Phys. Chem. C 113 (2009) 91419152.
[22] A. van de Walle, G. Ceder, J. Phase Equilibria 23 (2002) 348359.
[23] H.C. Ham, J.A. Stephens, G.S. Hwang, J. Han, S.W. Nam, T.H. Lim, Catal. Today
165 (2011) 138144.
[24] L.K. Ouyang, G.J. Da, P.F. Tian, T.Y. Chen, G.D. Liang, J. Xu, Y.F. Han, J. Catal. 311
(2014) 129136.
[25] H.C. Ham, G.S. Hwang, J. Han, S.W. Nam, T.H. Lim, J. Phys. Chem. C 113 (2009)
1294312945.
[26] P. Vassilev, M.T.M. Koper, J. Phys. Chem. C 111 (2007) 26072613.
[27] J.J. Burton, E.S. Machlin, Phys. Rev. Lett. 37 (1976) 14331436.
[28] B. Hammer, Y. Morikawa, J.K. Norskov, Phys. Rev. Lett. 76 (1996) 21412144.
[29] N.M. Markovic, P.N. Ross, Surf. Sci. Rep. 45 (2002) 121229.
[30] P. Han, S. Axnanda, I. Lyubinetsky, D.W. Goodman, J. Am. Chem. Soc. 129
(2007) 1435514361.
[31] F. Gao, Y.L. Wang, D.W. Goodman, J. Am. Chem. Soc. 131 (2009) 57345735.
[32] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 38653868.
[33] S. Stolbov, M.A. Ortigoza, T.S. Rahman, J. Phys.-Cond. Matter 21 (2009).
[34] P.E. Blochl, Phys. Rev. B 50 (1994) 1795317979.
[35] E.D. German, M. Sheintuch, J. Phys. Chem. C 112 (2008) 1437714384.
[36] Z.J. Li, F. Gao, W.T. Tysoe, J. Phys. Chem. C 114 (2010) 1690916916.
[37] J.A. Stephens, G.S. Hwang, J. Phys. Chem. C 115 (2011) 2120521210.
[38] H.C. Ham, D. Manogaran, K.H. Lee, K. Kwon, S.A. Jin, D.J. You, C. Pak, G.S.
Hwang, J. Chem. Phys. 139 (2013).
[39] J. Greeley, J.K. Norskov, J. Phys. Chem. C 113 (2009) 49324939.
[40] S. Kondo, M. Nakamura, N. Maki, N. Hoshi, J. Phys. Chem. C 113 (2009)
1262512628.

You might also like