Surface Science: Kai Li, Zhongjun Zhou, Ying Wang, Zhijian Wu

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Surface Science 612 (2013) 63–68

Contents lists available at SciVerse ScienceDirect

Surface Science
journal homepage: www.elsevier.com/locate/susc

A theoretical study of CH4 dissociation on NiPd(111) surface


Kai Li a, Zhongjun Zhou a, b, Ying Wang a,⁎, Zhijian Wu a,⁎
a
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, PR China
b
State Key Laboratory of Theoretical and Computational Chemistry, Institute of Theoretical Chemistry, Jilin University, Changchun 130023, PR China

a r t i c l e i n f o a b s t r a c t

Article history: The CH4 dissociation on the NiPd(111) surface is studied by using the density functional theory (DFT). The pos-
Received 4 January 2013 sible adsorption sites are proposed and the favorite adsorption site(s) are determined. The potential energy
Accepted 20 February 2013 curve for CH4 dissociation is presented. Compared with pure Ni(111) and Pd(111) surfaces, the dissociation of
Available online 27 February 2013
CH4 on NiPd(111) surface is more favored, especially on the Ni reaction center of NiPd(111) surface. The intro-
duction of Pd improves the Ni catalytic ability for CH4 dissociation. A synergistic effect exists between Ni and Pd
Keywords:
NiPd(111)
that results in an improved catalytic performance for CH4 disassociation over that of either parent metal. Bime-
CH4 dissociation tallic NiPd is predicted to be a good catalyst for CH4 dissociation, in good agreement with experiment.
Reaction barrier © 2013 Elsevier B.V. All rights reserved.

1. Introduction On the theoretical side, there were also several studies available
by using the density functional theory (DFT). This included the CH4
The methane dissociation is the key to transform the resources of nat- dissociation over a Ni-based bimetallic (111) surface such as Ni–Au
ural gas into useful products like hydrogen. However, the high stability of [18,42,43], Ni–Co [44], Ni–Cu [45], Ni–Sn [19], or Ni–M surface
methane makes it hard to dissociate at room temperature. Thus, catalytic (M = Cu, Ru, Rh, Pd, Ag, Pt, and Au) [46], etc. The results indicated
CH4 dissociation to produce hydrogen has attracted increasing attention that the introduction of Au and Sn to Ni made bimetallic surface less
in recent years [1–9]. Because of the good catalytic activity and low cost, active for CH4 dissociation and weakened C adsorption [18,19,43].
Ni-based catalysts have become one of the most attractive catalysts in The relative high energy barriers and the low C binding energy
CH4 dissociation. However, carbon deposition is a severe problem for explained the experimentally observed coke resistance. To our
Ni-based catalysts and usually deactivates the catalysts [10–17]. Thus, knowledge, theoretical study of CH4 dissociation over Ni–Pd surface
improving the activity of Ni-based catalysts and suppressing coke forma- is still rare. Recently, the physical origin of the synergistic effect
tion are challenging technological problems. over Ni–M surface (M = Cu, Ru, Rh, Pd, Ag, Pt, and Au) with Ni/
In order to improve the stability and activity of Ni-based catalysts, a bi- M = 3 was conducted [46]. Although the electronic structure and ad-
metallic catalyst is often used, instead of pure Ni. Many Ni-based bimetal- sorption properties had been investigated, the detailed reaction
lic catalysts have been proposed by experiments, such as Ni–Cu, Ni–Co, mechanism of CH4 dissociation over NiPd was not provided. In this
Ni–Pt, Ni–Au, Ni–Sn and Ni–Pd etc [18–41]. Among them, Ni–Au and work, we present a study on the possible reaction mechanism for
Ni–Sn bimetallic catalysts show a suppressed coke formation, but at the CH4 dissociation on NiPd(111) surface by using the density functional
cost of losing a catalytic activity [18,19]. Ni–Pt bimetallic catalyst with theory (DFT). The Pd/Ni = 1 is selected based on the previous exper-
the lowest Pt/Ni ratio (Pt0.3Ni10) exhibited a higher catalytic activity and iment, because at this ratio, the hydrogen yield is the highest [23]. The
stability for CH4 dissociation than monometallic Ni [20,21]. In addition, adsorption geometries, energies of CHx (x = 0–4) on NiPd(111) sur-
Ni–Pd bimetallic catalyst has received much attention for CH4 dissociation face and activation energies for the four elementary steps involved in
because of the high activity and stability [22–27]. Takenaka et al. investi- CH4 dissociation are investigated. For comparison, CH4 dissociation
gated the CH4 dissociation over M–Ni/SiO2 (M = Cu, Rh, Pd, Ir, Pt) and on elemental Pd(111) and Ni(111) surfaces is also given.
found that Pd–Ni catalysts had the highest yields of hydrogen with a
mole ratio of Pd / (Ni + Pd) = 0.5 [22,23]. Steinhauer et al. investigated 2. Computational details
the catalytic activity of bimetallic Ni–Pd on different supports and found
that they had a better performance than single metal Ni or Pd-catalysts 2.1. Method
for CH4 dissociation [24]. The superior catalytic performance of Pd pro-
moted Ni catalysts came from the synergistic effect between Ni and Pd The calculations were performed using Vienna ab-initio simula-
[25]. tion package (VASP) [47–50]. The interactions between valence elec-
trons and ion cores were treated by Blöchl's all-electron-like
⁎ Corresponding authors. Tel.: +86 43185262801. projector augmented wave (PAW) method [51,52]. The exchange-
E-mail addresses: ywang_2012@ciac.jl.cn (Y. Wang), zjwu@ciac.jl.cn (Z. Wu). correlation functional was the generalized gradient approximation

0039-6028/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.susc.2013.02.012
64 K. Li et al. / Surface Science 612 (2013) 63–68

with the Perdew–Burke–Ernzerhof, known as GGA-PBE [53]. The bottom are fixed in their bulk positions, and those in the top and sec-
wave functions at each k-point were expanded with a plane wave ond layers are allowed to relax. A vacuum layer as large as 12 Å is
basis set with a kinetic cutoff energy up to 400 eV and electronic occu- used along the c direction normal to the surface to avoid periodic in-
pancies were determined according to a Fermi scheme with an energy teractions. In order to reduce the interaction between adsorbates on
smearing of 0.1 eV. Brillouin zone integration was approximated by a the surface, a (2 × 2) supercell is used to model the coverage of a 1/4
sum over selected special k-points using the Monkhorst–Pack method monolayer.
[54] and they are set to 3 × 3 × 1. Geometries were optimized until The chemisorption energy, Eads, is defined as follows
the energy was converged to 1.0 × 10−6 eV/atom and the force con-
verged to 0.01 eV/Å. Because of the existence of magnetic atom, spin ΔEads ¼ Eadsorbates=slab −ðEslab þ Eadsorbates Þ
polarization was considered in all calculations.
The transition state (TS) structures and the reaction pathway were where Eadsorbates/slab is the total energy of adsorbates on (111) surface,
computed using the climbing image nudged elastic band (CI-NEB) Eslab is the total energy of the bare slab of the surface, and Eadsorbates is
method [55]. The minimum energy path (MEP) was optimized the total energy of free adsorbates. The first two terms are calculated
using a force-based conjugate-gradient method [50] until the maxi- with the same parameters. The third term is calculated by setting the
mum force is less than 0.01 eV/Å. In addition, TS was verified by isolate adsorbate in a box of 12 Å × 12 Å × 12 Å. The negative ΔEads
only one imaginary frequency. indicates exothermic chemisorption, positive value suggests endo-
thermic chemisorptions, while values closer to zero (slightly negative
2.2. Model or slightly positive), are representative of weak physisorption.

The NiPd alloy with Ni/Pd = 1 showed a remarkable performance 3. Results and discussion
in CH4 dissociation and reforming [22–24]. The formation of the Ni–Pd
alloys was observed to include (111) and (200) surfaces by the XRD 3.1. CHx adsorption on NiPd(111)
pattern [26]. Thus, we choose the ground state of Ni (face-centered-
cubic lattice) [56,57] and replace half of the Ni atoms by Pd atoms to There exist nine possible adsorption sites on the NiPd(111) surface
simulate the 1:1 Ni–Pd alloy. The calculations show that the lattice con- (Fig. 1), they are, two different Top sites (TNi and TPd), Hexagonal
stants of alloy are a = 3.841 and c = 3.618 Å, different from the lattice Close Packed sites (HCPNi and HCPPd), and Face Center Cubic sites
constants of Ni bulk (3.518 Å) and Pd bulk (3.920 Å). Since experimen- (FCCNi and FCCPd), as well as three different Bridge sites (BNi–Pd, BNi–Ni
tally, NiPd(111) and NiPd(200) surfaces are observed [26], in order to and BPd–Pd). Here, we have investigated all the possible adsorption con-
determine which surface is more stable, the surface energy (Esurf) is in- figurations and energies of CHx (x = 0–4) and H on NiPd(111) surface
troduced which is defined as the energy of the slab related to its bulk (see Fig. S1 in the supporting information).
reference [58]:
3.1.1. CH4 adsorption on NiPd(111)
E −nEbulk According to previous studies, CH4 adsorption is reported to be a
Esurf ¼ slab
2A physisorption process on pure metal surfaces with an extremely
small adsorption energy [42–45]. Similarly, our results show that
where Eslab is the total energy of the surface slab, Ebulk is the total energy the introduction of Pd also does not increase CH4 adsorption energy
of the bulk NiPd, A is the surface area with a factor of 2 due to each slab significantly. The adsorption energy is calculated to be − 0.02 eV,
containing two surfaces, and n is the number of NiPd formula units in close to the values on pure Ni(111) (− 0.02) and Pd(111) (− 0.03)
the slab. The small Esurf means that the surface is more stable [58]. surfaces, respectively.
Thus, the calculated Esurf values of 0.045 eV for (111) and 0.103 eV for
(200) indicate that NiPd(111) is more stable than (200). This is the rea- 3.1.2. CH3 and CH2 adsorption on NiPd(111)
son why we study NiPd(111) surface in this work. From the adsorption energies of CH3 and CH2 on NiPd(111) (Fig. S1,
The surface is obtained by cutting NiPd (fcc) bulk along (111) di- CH3(a) to CH3(h) and CH2(a) to CH2(g)), we find that CH3 and CH2 pre-
rection, the thickness of surface slab is chosen to be a three-layer slab fer to occupy the hollow site with two Ni atoms, that is, FCCPd site as
which is proved to be a reasonable model and widely employed in shown in Fig. 2b and c. For CH3, all three H atoms point outward, with-
previous literatures [43–45]. In our calculations, the atoms in the out interaction between H and Ni/Pd. For the adsorbed CH2, it is seen

Fig. 1. Top view (a) and side view (b) of the possible adsorption sites on NiPd(111) surface.
K. Li et al. / Surface Science 612 (2013) 63–68 65

that the energy is lowered when one H atom is bonded to the Ni atom, smaller than those for CH (around − 6.0 eV). In addition, the adsorp-
which is favorable for the following dissociation of C–H. In the tion energies of the four H-adsorption configurations are very close
favorite CH3 and CH2 adsorption configurations (Fig. 2b and c), C\Ni and the diffusion barrier of H on NiPd(111) is only 0.15 eV,
bond lengths (2.115 Å in CH3, 1.859 and 1.980 Å in CH2) are shorter suggesting that the migration of H on the surface is relatively easy.
than the C\Pd bond lengths (2.330 Å in CH3, 2.072 Å in CH2). This
might be an indication that the interaction between C and Ni is stronger
than that between C and Pd. 3.1.4. C adsorption on NiPd(111)
Different from the configurations on Ni(111) and Pd(111) sur-
3.1.3. CH and H adsorption on NiPd(111) faces, there are only three stable geometries for C adsorption on
CH prefers to locate at the hollow sites HCPPd and FCCPd, whose ad- NiPd(111) (Fig. S1, C(a) to C(c)). The initial structures FCCPd and
sorption energies are −6.27 eV and −6.24 eV, respectively, slightly HCPPd are all converged to the most stable fourfold site (denoted as
higher than −6.35 eV on the pure Ni(111) surface at the same level FF in Fig. 2e). It is clear that when C is adsorbed on FF, the NiPd
of theory (Fig. S1, CH(a) to CH(d)). This indicates that CH on NiPd (111) surface deforms severely. In this adsorption configuration,
alloy is less stable than on pure Ni surface. Similar situation is observed compared to the isolated NiPd(111), the bond of Ni\Pd is shortened
in the previous studies on NiCo(111) and NiCu(111) [44,45]. from 2.674 to 2.648 Å, while Ni\Ni and Pd\Pd bonds are elongated
Similar to CH adsorption configurations, H also prefers to adsorb at by 14.1% (from 2.723 to 3.106 Å) and 1.0% (from 2.723 to 2.749 Å),
hollow sites with four possible configurations (Fig. S1, H(a) to H(d)). respectively. The surface distortion is consistent with the experiment
The calculated adsorption energies (around − 2.7 eV) are much observation by Takenaka et al. [22,23].

Fig. 2. The most stable adsorption geometries and selected parameters of CHx (x = 0–4) and H on NiPd(111) (only the top two layers of the NiPd slab are displayed in the top view,
and the top layer in the side view). The gray balls denote C atoms and the white balls denote H atoms. The top H–Ni indicates one of the H atoms pointing to the Ni atom.
66 K. Li et al. / Surface Science 612 (2013) 63–68

3.2. CH4 dissociation on NiPd(111) centers are given in Fig. 3 (TS1 and TS1′, with the latter indicating Pd
as a reaction center). In the two configurations, the CH3 fragment is
In order to investigate the mechanism of CH4 dissociation on the adsorbed at the top site and the breaking H atom is located at the
NiPd (111) surface, we choose the most stable geometry as the initial nearest hollow site. The breaking C\H bond increases by 45.7%
state and co-adsorbed CHx − 1 + H species with the lowest Eads are (from 1.097 to 1.598 Å in TS1) or 48.3% (from 1.097 to1.627 Å in
set as the final configurations in the minimum energy path. Each TS1′) compared to the C\H bond length in isolated CH4, and the
step is divided into two possible pathways: one is with Ni as the forming Ni\H bond is elongated by 22.7% (from 1.723 to 2.114 Å in
catalytic reaction center and the other is with Pd as the catalytic reac- TS1) or 27.3% (from 1.723 to 2.194 Å in TS1′) with respect to the
tion center. bond length of H adsorbed on FCCPd site. The elongation of the formed
bond is smaller than that of the breaking bond, indicating that the
3.2.1. CH4 → CH3 + H two TSs are product-like. This means that the both reaction channels
As mentioned above, CH4 on NiPd(111) is a physical adsorption proceed via a “late” transition state. This character in the TSs is consis-
process, and all configurations have almost the same adsorption ener- tent with the calculated result that CH4 decomposition to CH3 and H is
gy. Therefore, we select the CH4 adsorption on the top of Ni or Pd con- an endothermic process with the reaction energies of 0.30 and
figurations as the initial state. As FCCPd is the most stable site for CH3 0.48 eV for the Ni and Pd reaction centers, respectively, in agreement
adsorption, while H adsorption has nearly the same stability for the with Hammond's postulate [59]. It is worth to note that the reaction
studied sites, the configurations of CH3 adsorbed at FCCPd and H locat- energy with the Ni reaction center is less endothermic by 0.18 eV
ed at FCCPd are chosen as the final state in the first step of CH4 disso- compared to that of the Pd reaction center. In addition, the energy
ciation on NiPd(111). Two transition states on two different reaction barrier height of the reaction with the Ni reaction center (0.84 eV,

Fig. 3. The transition state structures for CHx (x = 1–4) dissociation on NiPd(111)(only the top two layers of the NiPd slab are displayed in the top view, and the top layer in the side
view). (a–d) for TS with the Ni reaction center and (e–g) for TS′ with the Pd reaction center. The C–H (rupture) indicates the breaking C\H bond. The vimag indicates the imaginary
frequency.
K. Li et al. / Surface Science 612 (2013) 63–68 67

Fig. 4) is lower than that of the Pd reaction center (1.01 eV). This sug- on the Pd center also confirmed that the Ni reaction center is more
gests that taking Ni as the reaction center is favored both thermody- preferred. Thus, Ni atoms have a higher catalytic activity than Pd for
namically and kinetically. CHx (x = 1–4) dissociation.

3.2.2. CH3 → CH2 + H 3.3. Comparison with Ni(111) and Pd(111) surfaces
CH3 at the FCCPd site is the initial structure for CH3 dissociation,
while both of the CH2 and H co-adsorbed at FCCPd site are the final As discussed above, it is seen that Ni shows a higher catalytic activ-
states for the reaction with the Ni center. For the Pd reaction center, ity than Pd on NiPd(111) surface, so the minimum energy path of the
the configuration of CH2 at FCCPd site and H at HCPPd site is set as reaction should be CH4 dissociation on NiPd(111) with the Ni center.
the final state. The two TS structures for CH3 dissociation on the In order to show the performance of NiPd catalyst, we also investigat-
NiPd(111) surface are shown in Fig. 3b and f. In Fig. 3b. The carbon- ed the mechanism of CH4 dissociation over pure Ni(111) and Pd(111)
containing fragment CH2 is located at the FCCPd site and the detached surfaces by the same level of theory (possible configurations are
H atom is positioned at the adjacent hollow site with dC–H = 1.620 Å. shown in Figs. S2–S4). The potential energy surface of CH4 dissociation
In Fig. 3f, the detached H is located on the top of the Pd atom with the on Ni(111), Pd(111) and NiPd(111) surfaces is shown in Fig. 4. It is
C–H distance of 1.834 Å. It is clear to see that the breaking C\H bond seen that CHx (x = 1–4) dehydrogenation into CHx − 1 (x = 1–4)
length in the former case (Ni reaction center) is shorter than the lat- and H on NiPd(111) is more favored both thermodynamically and ki-
ter (Pd reaction center) by around 0.2 Å, suggesting that CH3 dissoci- netically than that on Ni(111) and Pd(111) surfaces.
ation is more difficult on the Pd reaction center than on the Ni A synergistic effect exists between Ni and Pd that results in the
reaction center. The calculated CH3 dehydrogenation with the Ni re- improved catalytic performance of NiPd for CH4 dissociation over
action center is endothermic only by 0.15 eV and the barrier height that of either parent metal. This might be explained in the following.
is 0.64 eV (Fig. 4), while those for the reaction with Pd are 0.19 and First, the introduction of Pd changes the geometrical structure of
0.90 eV, respectively. pure Ni. The lattice parameter increases from 3.518 Å in pure Ni to
3.841 Å for a and 3.618 Å for c in NiPd alloy. Second, it also makes
3.2.3. CH2 → CH + H the surface charge redistributed. Our calculations indicate that with
CH2 adsorbed at the FCCPd site (Fig. S1, CH2(a) and Fig. 2c) with the increase of x in CHx (x = 1, 2, 3), the average net charge of Ni
two Ni atoms is more stable than CH2 at the hollow site with two is positive and changes from 0.27, 0.28 to 0.30 |e|, while that of Pd
Pd atoms (Fig. S1, CH2(d)–CH2(e)). In addition, in the former config- is negative from − 0.12, − 0.19 to − 0.22 |e|. The positive charge of
uration, one of the H atoms bonded to Ni with the dNi–H = 1.732 Å. Ni and negative charge of Pd are in agreement with the small electro-
Thus, the adsorbed H can be easily removed from CH2. Therefore, negativity of Ni (1.91) and large electronegativity of Pd (2.20). This
we only consider the CH2 dehydrogenation with the Ni reaction cen- suggests that the Ni\Pd bond is polarized. Since the radicals of CHx
ter. In TS3, as shown in Fig. 3c, the distance between C and the (x = 1–3) have redundant electrons, they prefer to be adsorbed to
cleaved H atom is 1.502 Å, which increases by 36.9% (from 1.097 to the positive Ni. Thus, Ni becomes the reaction center.
1.502 Å) compared to the C\H bond length in isolated CH4, while Hanmmer and Nørskov pointed out that the d-band center of sim-
the Ni\H bond is elongated by 79.3% (from 1.723 to 3.090 Å). The ex- ilar transition metal correlates with the general catalytic activity and
tent of the elongation of the Ni\H bond is larger than that of the the reactivity of the transition metal system increases as the d-band
breaking bond, indicating that the TS is reactant-like. Thus, the reac- center is shifted up [60,61]. Thus, to further explain why NiPd alloy
tion channel proceeds via an “early” transition state. This is consistent has a higher catalytic activity and Ni atom is the reaction center for
to the early character of TS with an exothermic process (− 0.27 eV) in CHx dissociation, the d-band centers of the Ni(111), Pd(111) and
Hammond's postulate [59]. It is noted that CH2 decomposition is the NiPd(111) surfaces are examined. Our calculated d-band centers of
only exothermic process in the whole CH4 dissociation processes on pure Ni and Pd are − 1.47 eV and − 1.89 eV, respectively, while the
NiPd(111) surface. Furthermore, the C–H dissociation barrier is the average value for NiPd is − 1.71 eV. Obviously, the average d-band
lowest with an energy of 0.28 eV compared to other dissociation pro- center of NiPd alloy surface is far away from a Fermi level compared
cesses (Fig. 4). with Ni(111). However, the d-band center of Ni on NiPd(111) surface
is −1.14 eV, closer to a Fermi level than pure Ni, demonstrating that
3.2.4. CH → C + H CHx adsorbed on the sites with two Ni atoms is more stable than
For the Ni reaction center, CH at FCCPd is selected as the initial on the other sites, and dissociations at the Ni reaction center are
structure, while for the Pd reaction center, CH at HCPPd is set as the
initial state since the adsorption for CH at HCPPd is slightly more sta-
ble than FCCPd by 0.03 eV (Fig. S2, CH(a)–CH(b)). The configuration of
C at FF and H at FCCPd sites is set as final state. The TS structures are
shown in Fig. 3d and g. In Fig. 3d, the breaking C\H bond is
1.562 Å, and its orientation is almost parallel to the surface. From
the calculated data, the deformation of NiPd(111) surface is obvious.
This is because compared with an isolated NiPd(111) (planar) sur-
face, the two Ni atoms bonded with C are enlarged by 11% (from
2.723 to 3.012 Å) on the xy plane, while the Pd and Ni bonded with
C rise about 0.187 and 0.023 Å above the surface (along z axis). The
energy barrier of CH dehydrogenation into C and H is 1.04 eV
(Fig. 4) and the reaction is endothermic by 0.32 eV, which is the
highest electronic barrier during the CH4 dissociation process. In
Fig. 3g, the C atom locates at the middle between two Ni atoms
with dC–Ni = 1.754 Å, and H moves toward the top of Pd with the
breaking C\H bond of 2.198 Å. However, the deformation of the sur-
face is not so large as that in the Ni reaction center. The activation en-
ergy for CH dissociation with the Pd center is 1.29 eV and the reaction Fig. 4. Potential energy curves for CH4 dissociation on Ni(111), Pd(111) and NiPd(111)
is endothermic by 0.35 eV. The higher barrier and more endothermic surfaces, respectively.
68 K. Li et al. / Surface Science 612 (2013) 63–68

preferred. Thus, we conclude that Ni–Pd synergistic effect enhances [16] H.F. Abbas, W.M.A.W. Daud, Int. J. Hydrog. Energy 35 (2010) 141.
[17] C.H. Bartholomew, Appl. Catal. A 212 (2001) 17.
Ni catalytic activity in alloy [46]. [18] F. Besenbacher, I. Chorkendorff, B.S. Clausen, B. Hammer, A.M. Molenbroek, J.K.
Nørskov, I. Stensgaard, Science 279 (1998) 1913.
4. Conclusions [19] E. Nikolla, J. Schwank, S. Linic, J. Catal. 263 (2009) 220.
[20] D. San-José-Alonso, J. Juan-Juan, M.J. Illán-Gómez, M.C. Román-Martínez, Appl.
Catal. A 371 (2009) 54.
The mechanism of methane dissociation on NiPd(111) surface has [21] S. Özkara-AydInoglu, A.E. Aksoylu, Int. J. Hydrog. Energy 36 (2011) 2950.
been investigated by using the density functional theory. Our results [22] S. Takenaka, Y. Shigeta, E. Tanabe, K. Otsuka, J. Catal. 220 (2003) 468.
[23] S. Takenaka, Y. Shigeta, E. Tanabe, K. Otsuka, J. Phys. Chem. B 108 (2004) 7656.
show that CHx (x = 1–3) prefers to adsorb at the hollow sites with [24] B. Steinhauer, M.R. Kasireddy, J. Radnik, A. Martin, Appl. Catal. A 366 (2009) 333.
two Ni atoms (FCCPd or HCPPd) and CHx (x = 1–4) dissociation on [25] S. Damyanova, B. Pawelec, K. Arishtirova, J.L.G. Fierro, C. Sener, T. Dogu, Appl.
the Ni reaction center is more favored than on the Pd reaction center. Catal. B Environ. 92 (2009) 250.
[26] Y. Nabae, I. Yamanaka, M. Hatano, K. Otsuka, J. Phys. Chem. C 112 (2008) 10308.
The largest electronic energy barrier of CH4 dissociation is CH dissoci-
[27] F. Menegazzoa, M. Signorettoa, F. Pinnaa, P. Cantona, N. Pernicone, Appl. Catal. A
ation to atomic carbon and hydrogen on NiPd(111) surface. By com- 439 (2012) 80.
paring with pure Ni(111) and Pd(111) surfaces, we observed that [28] G. Chen, G.Q. Guan, Y.T. Kasai, H.X. You, A. Abudul, J. Power Sources 196 (2011)
Ni–Pd alloy shows the highest catalytic activity. A synergistic effect 6022.
[29] J.G. Zhang, H. Wang, A.K. Dalai, J. Catal. 249 (2007) 300.
exists between Ni and Pd that results in an improved catalytic perfor- [30] J.C.S. Wu, H.C. Chou, Chem. Eng. J. 148 (2009) 539.
mance for CH4 dissociation over that of either parent metal. We con- [31] A. Horváth, G.O. Stefler, A. Kienneman, A. Pietraszek, L. Guczi, Catal. Today 169
clude that NiPd is a good catalyst to decompose CH4, consistent to the (2011) 102.
[32] T. Huang, W. Huang, J. Huang, P. Ji, J. Fuel Process. Techol. 92 (2011) 1868.
experimental observation. [33] J.G. Zhang, H. Wang, A.K. Dalai, Ind. Eng. Chem. Res. 48 (2009) 677.
[34] M. García-Diéguez, I.S. Pieta, M.C. Herrera, M.A. Larrubia, L.J. Aleman, J. Catal. 270
Acknowledgments (2010) 136.
[35] M. García-Diéguez, E. Finocchio, M.A. Larrubia, L.J. Alemany, G. Busca, J. Catal. 274
(2010) 11.
This work is financially supported by the National Natural Science [36] M. García-Diéguez, I.S. Pieta, M.C. Herrera, M.A. Larrubia, L.J. Alemany, Appl. Catal.
Foundation of China (grant nos: 20921002, 21273219, 21203174). A 377 (2010) 191.
[37] N.C. Triantafyllopoulos, S.G. Neophytides, J. Catal. 239 (2006) 187.
The authors are also thankful for the financial support from the De- [38] S.Y. Foo, C.K. Cheng, T.H. Nguyen, A.A. Adesina, Catal. Today 164 (2011) 221.
partment of Science and Technology of Sichuan Province. The compu- [39] Y. Nabae, I. Yamanaka, M. Hatano, K. Otsuka, J. Electrochem. Soc. 153 (2006)
tational time is supported by the Performance Computing Center of A140.
[40] L.B. Avdeeva, O.V. Goncharova, D.I. Kochubey, V.I. Zaikovskii, L.M. Plyasova, B.N.
Jilin University, China.
Novgorodov, Sh.K. Shaikhutdinov, Appl. Catal. A 141 (1996) 117.
[41] T.V. Reshetenko, L.B. Avdeeva, Z.R. Ismagilov, A.L. Chuvilin, V.A. Ushakov, Appl.
Appendix A. Supplementary data Catal. A 247 (2003) 51.
[42] P. Kratzer, B. Hammer, J.K. Nørskov, J. Chem. Phys. 105 (1996) 5595.
[43] H.Y. Liu, R.X. Yan, R.G. Zhang, B.J. Wang, K.C. Xie, J. Nat. Gas Chem. 20 (2011) 611.
Supplementary data to this article can be found online at http:// [44] H.Y. Liu, R.G. Zhang, R.X. Yan, B.J. Wang, K.C. Xie, Appl. Surf. Sci. 257 (2011) 8955.
dx.doi.org/10.1016/j.susc.2013.02.012. [45] H.Y. Liu, R.G. Zhang, R.X. Yan, R.X. Li, B.J. Wang, K.C. Xie, Appl. Surf. Sci. 258 (2012)
8177.
[46] C. Fan, Y.A. Zhu, Y. Xu, Y. Zhou, X.G. Zhou, J. Chem. Phys. 137 (2012) 014703.
References [47] G. Kresse, J. Furthmüller, Comput. Mater. Sci. 6 (1996) 15.
[48] G. Kresse, J. Hafner, Phys. Rev. B 47 (1993) 558.
[1] K. Tomishige, Y.G. Chen, K. Fujimoto, J. Catal. 181 (1999) 91. [49] G. Kresse, J. Hafner, Phys. Rev. B 49 (1994) 14251.
[2] J.R. Rostrup-Nielsen, J.H.B. Hansen, J. Catal. 144 (1993) 38. [50] G. Kresse, J. Furthmüller, Phys. Rev. B 54 (1996) 11169.
[3] A. Erdöhelyi, J. Cserényi, F. Solymosi, J. Catal. 141 (1993) 287. [51] P.E. Blöchl, Phys. Rev. B 50 (1994) 17953.
[4] A. Malaika, B. Krzyzynska, M. Kozlowski, Int. J. Hydrog. Energy 35 (2010) 7470. [52] G. Kresse, D. Joubert, Phys. Rev. B 59 (1999) 1758.
[5] J.H. Kim, D.J. Suh, T.J. Park, K.L. Kim, Appl. Catal. A 197 (2000) 191. [53] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865.
[6] N.Z. Muradov, Int. J. Hydrog. Energy 18 (1993) 211. [54] H.J. Monkhorst, J.D. Pack, Phys. Rev. B 13 (1976) 5188.
[7] M.C.J. Brandford, M.A. Vannice, Appl. Catal. A 142 (1996) 97. [55] G. Henkelman, B.P. Uberuaga, H. Jónsson, J. Chem. Phys. 113 (2000) 9901.
[8] K. Otsuka, S. Kobayashi, S. Takenaka, Appl. Catal. A 190 (2000) 261. [56] J. Gong, L.L. Wang, Y. Liu, J.H. Yang, Z.G. Zong, J. Alloys Compd. 457 (2008) 6.
[9] S. Tang, L. Ji, J. Lin, H.C. Zeng, K.L. Tan, K. Li, J. Catal. 194 (2000) 424. [57] G.Y. Guo, H.H. Wang, Chin. J. Phys. 38 (2000) 949.
[10] Y.H. Hu, E. Ruckenstein, Adv. Catal. 48 (2004) 297. [58] Y. Duan, Phys. Rev. B 77 (2008) 045332.
[11] S.C. Schouten, O.L.J. Gijzeman, G.A. Bootsma, Bull. Soc. Chim. Belg. 88 (1979) 541. [59] G.S. Hammond, J. Am. Chem. Soc. 77 (1955) 334.
[12] C.H. Bartholomew, Catal. Rev. Sci. Eng. 24 (1982) 67. [60] B. Hammer, J.K. Nørskov, Surf. Sci. 343 (1995) 211.
[13] J.M. Ginsburg, J. Pina, T.E. Solh, H.I. de Lasa, Ind. Eng. Chem. Res. 44 (2005) 4846. [61] B. Hammer, Top. Catal. 37 (2006) 3.
[14] J.R. Rostrup-Nielsen, D.L. Trimm, J. Catal. 48 (1977) 155.
[15] J.R. Rostrup-Nielsen, Catal. Today 37 (1997) 225.

You might also like