CMS 2015 Pt3M CO Oxidation

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Computational Materials Science 96 (2015) 237–245

Contents lists available at ScienceDirect

Computational Materials Science


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

DFT studies on Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh) clusters for CO oxidation
Xin Lian a, Wenlong Guo a, Feila Liu a, Yang Yang a, Peng Xiao b, Yunhuai Zhang a,⇑, WeiQuan Tian c
a
College of Chemical Engineering, Chongqing University, Chongqing 400044, China
b
College of Physics, Chongqing University, Chongqing 400044, China
c
State Key Laboratory of Urban Water Resource and Environment, Institute of Theoretical and Simulational Chemistry, Academy of Fundamental and Interdisciplinary Sciences,
Harbin Institute of Technology, Harbin 150080, China

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

Article history: The reaction mechanism of CO oxidation catalyzed by several Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh) clusters
Received 8 July 2014 has been investigated with density functional theory calculations in the present work. The reaction pre-
Received in revised form 15 September 2014 fers to proceed via Langmuir–Hinshelwood mechanism. The calculated barriers for the reactions medi-
Accepted 20 September 2014
ated by bimetallic clusters are comparable with that catalyzed by monometallic Pt4 cluster. According
Available online 11 October 2014
to thermodynamics and kinetics results, CO oxidation can take place readily without thermal activation
and the O2 scission to form OCOO* is the rate-determining-step. Pt3Mo exhibits superior catalytic activity
Keywords:
for CO oxidation among all those investigated clusters and good adsorption for O2. The different perfor-
CO oxidation
Bimetallic clusters
mance of those bimetallic clusters for CO oxidation is scrutinized with aid of molecular orbital and nat-
Catalysis ural bond orbital population analysis.
Kinetics Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction the O2 concentration increases to 2%, the CO conversion increases


to 99% at 90 °C. A 4:1 ratio was proposed for PtMo alloy catalyst
Catalytic oxidation of CO has been extensively studied due to its [20], with this ratio, the PtMo catalyst displays the lowest onset
importance for both fundamental research and industrial applica- oxidation potential, i.e. the highest activity for CO oxidation.
tions, e.g. tail gas processing [1] and purification of hydrogen for Two possible mechanisms [31–35] have been proposed to
proton-exchange membrane fuel cells (PEMFCs) [2]. Among vari- explain the activities for CO oxidation of Pt-based catalysts. One
ous heterogeneous catalysts, precious metals, Pt [3–6], Ru [7,8], is based on the promotion of CO oxidation on Pt atoms by a second
Pd [9–12], and Au [13–15] are efficient catalysts for oxidation of metal which produces OH-type species at lower potential. The
CO with O2. In comparison with other noble metals, Pt catalysts other one is the ligand mechanism, which is based on modification
show the most promising performance and are considered as ideal of electronic property of the Pt by a second metal in the catalyst.
electrode materials in fuel cells. However, the limited Pt source and The modified electronic property results in the different chemi-
deactivation by trace amounts of CO are two challenging issues for sorption (or lower desorption energy) of the catalyst for CO so that
its large scale application [16]. Therefore, it is well desirable to find the CO coverage on the Pt sites used for H2 oxidation is reduced.
suitable Pt-based catalysts both economically and chemically effi- Theoretical works have been devoted to investigate CO oxidation
cient for CO oxidation. on Pt-based alloys to study the possible mechanism for the
To this end, an alternative strategy to enhance the reactivity of enhanced activity of Pt-based bimetallic catalysts for CO oxidation.
Pt nanoparticles, alloying Pt nanoparticles with a second metal to However, the possible mechanism is not yet clear. For example, Pt/
form a bimetallic Pt catalyst (‘‘nano-alloy’’) has attracted signifi- Ru [36,37] catalysts show superior activity toward methanol oxi-
cant attention. Some Pt-based binary-alloy catalysts, such as dation, because dehydrogenation process occurs on Pt site while
Pt–Ni [17,18], Pt–Mo [19–21], Pt–Sn [22,23], Pt–Co [24,25], Pt–Au Ru site provides oxygenated species for CO oxidation according
[26], and Pt–Ru [27,28], possess better activity and stability than to the bifunctional mechanism. On the contrary, in Pt/Au [38] cat-
the corresponding monometallic catalysts for CO oxidation. Pt/Ru alysts, Au does not participate in reaction directly and its crucial
with atomic ration of 2:1 achieves 90% CO conversion and 90% role is to stabilize Pt by enhancing the oxidation potential of Pt.
CO2 selectivity at 150 °C [29]. The most active catalyst of Pt–Pd/ Extensive experimental studies found that the presence of sec-
CeO2 has a composition of Pt/Pd = 1/7 [30]. On this catalyst, when ond metals (Ni, Mo, Ru, Pd, Rh), either alloyed or co-deposited with
Pt, yields significant improvement on the CO tolerance. However,
⇑ Corresponding author. there are few reports on how these transition metals doped Pt

http://dx.doi.org/10.1016/j.commatsci.2014.09.025
0927-0256/Ó 2014 Elsevier B.V. All rights reserved.
238 X. Lian et al. / Computational Materials Science 96 (2015) 237–245

Table 1
The ground states for the complexes of Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh) with O2 or CO. The superscripts and values in square brackets
denote the electronic states and the binding energies, respectively. Pt, Ni, Mo, Ru, Pd, Rh, C and O atoms are represented by blue, violet,
turquoise, light-sea green, teal, dark-cyan, gray and red spheres.

>(E H9@ >(E H9@ >(E H9@ >(E H9@ >(E H9@ >(E H9@

>(DGV H9@ >(DGV H9@ >(DGV H9@ >(DGV H9@ >(DGV H9@ >(DGV H9@

>(E H9@ >(DGV H9@ >(DGV H9@ >(DGV H9@ >(DGV H9@ >(DGV H9@

Fig. 1. Optimized geometries of intermediates and transition states involved in the CO oxidation in path I and path II on Pt4. Pt, C and O atoms are represented by blue, gray
and red spheres. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

catalysts improve the activity on CO oxidation. The reactions of identified and the enhanced catalytic activity of those bimetallic
Pt-based bimetallic clusters can provide useful information for clusters is scrutinized.
understanding the mechanisms involved in the realistic and
complicated catalytic systems [39,40].
In the present work, the reaction mechanism of CO oxidation on 2. Computational detail and models
Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh) clusters will be studied with den-
sity functional theory (DFT) based calculations. Reaction processes Due to there are usually no more than 4 atoms on the top level
via single-center pathway and two-center pathway approaching of surface participate in reaction of small molecules [41] and tetra-
the Pt and M atoms of Pt3M are considered. The most favorable hedral Pt3M used to be considered as the representative to illus-
reaction pathways of CO oxidation on each bimetallic catalyst are trate catalytic problems [38,42], tetrahedral Pt3M (M = Pt, Ni, Mo,
X. Lian et al. / Computational Materials Science 96 (2015) 237–245 239

3. Results and discussion

3.1. Adsorptions of CO and O2 on Pt3M clusters

Various geometries have been optimized for adsorption of O2


and CO on different adsorption sites of Pt3M (M = Pt, Ni, Mo, Ru,
Pd, Rh) clusters in different spin multiplicities. The most stable
conformations were chosen to study the possible complexes for
CO oxidation.
For CO adsorption, the molecule prefers the ‘‘end-on’’ [49] con-
formations with C atom binding to the metal atoms. As shown in
Table 1, the adsorption energies (Eads) of CO on the Pt sites are
much larger than those on the M sites, thus that CO prefers to bind
Fig. 2. Potential energy curve for CO oxidation promoted by Pt4 along path I (black
to the Pt sites when both Pt and M sites are available. Pt4 cluster
line) and path II (red line). The energies include zero-point energy corrections. (For
interpretation of the references to colour in this figure legend, the reader is referred has the highest adsorption energy 2.38 eV, which is about
to the web version of this article.) 0.37 eV higher than that of Pt3Rh cluster (the weakest adsorption).
With the doping of M atom dope into Pt clusters, the CO adsorption
Ru, Pd, Rh) was used to model the subnanoscale particle catalysts energy is lower than that on pure Pt cluster, and it results in less
in this work. All calculations were performed with the Gaussian03 CO poisoning of catalysts. The CO bond length elongates slightly
package [43]. DFT based method B3LYP [44–46] was used for reac- from initial 1.13 Å to 1.15 Å.
tion pathway location and property analyses. The LANL2DZ [47] However, the O2 molecule prefers the ‘‘side-on’’ configurations
effective core potential and basis set for metal atoms, and the 6- with both oxygen atoms binding to metal atoms. The calculated Eads
311G+(d) [48] basis set for C and O atoms were employed. No sym- of O2 on both Pt and M sites are lower than those of CO. On Pt3Ru,
metric constraints were imposed during geometrical optimizations Pt3Ni and Pt3Rh, O2 binds the M sites while it binds to the Pt sites
and various spin multiplicities were considered for metal clusters. on Pt3Pd and Pt3Mo. O2 has much lower adsorption energy on Pt3Pd
Frequency analysis for the all stationary points was performed at and Pt3Rh than on other investigated clusters. The small O2 adsorp-
the same level of theory to characterize the stationary points and tion energy, in comparison with CO, is attributed to the strong elec-
make the zero-point-energy (ZPE) corrections. Intrinsic reaction tronegativity of O atom [50], which enables O2 to attract electrons
coordinates (IRC) calculations were carried out to verify that each strongly. This is also evidenced by the elongated O2 bond length sig-
saddle point links two desired minima. nificantly from 1.20 Å to 1.30 Å activated upon adsorption.
The chemisorption energy of CO and O2 was calculated using From the above discussion, we demonstrated that both CO and
the following Eq. (1): O2 molecules can adsorb on the various metal clusters which will
  provide the initial stage for the catalytic reaction of CO oxidation.
Eads ¼ ECO=O2 þ EPt3M  ECO=O2 Pt3M ð1Þ Because CO has much stronger adsorption energy than that of O2,

Fig. 3. Optimized geometries of intermediates and transition states involved in the CO oxidation in paths I, II and III on Pt3Ni. Pt, Ni, C and O atoms are represented by blue,
violet, gray and red spheres. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
240 X. Lian et al. / Computational Materials Science 96 (2015) 237–245

Fig. 4. Potential energy curve for CO oxidation promoted by Pt3Ni along path I Fig. 6. Potential energy curve for CO oxidation promoted by Pt3Mo along path I
(black line), path II (red line) and path III (blue line). The energies include zero-point (black line), path II (red line) and path III (blue line). The energies include zero-point
energy corrections. (For interpretation of the references to colour in this figure energy corrections. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.) legend, the reader is referred to the web version of this article.)

Pt4 þ CO þ O2
we can conclude that the active sites in bimetallic clusters would
be first occupied by the coming CO rather than O2, while O2 is more In a previous work, Wang et al. [38] studied PtmAun (m + n = 4)
easily activated by the metal clusters compared to CO. and CO oxidation on Pt4 cluster. Comparative studies on the CO
oxidation were carried out in the present work with those in
3.2. CO oxidation on Pt3M (M=Pt, Ni, Mo, Ru, Pd, Rh) clusters Wang’s work.
Fig. 2 displays the potential energy profiles of CO oxidation on Pt4
Due to the stronger adsorption of CO on those bimetallic clus- and the structures of all stationary points along this potential energy
ters that O2, the Langmuir–Hinshelwood mechanism is considered curve are summarized in Fig. 1. IM1 and IM1* in paths I and II are
in the present work [51,52]: the co-adsorbed CO and O2 firstly form similar to those in a previous work [38], whereas they have lower
a peroxo-type complex intermediate, i.e., CO (gas) + O2 (gas) ? adsorption energies (52.92 and 64.46 kcal/mol). The breaking
OOCO* ? CO2 + O (ads). According to preference of adsorption of of O–O bond goes through transition states of TS1–2 and TS1 —2
CO and O2 on Pt3M, there are three possible pathways for CO oxi- and the energy barriers are 21.25 kcal/mol and 38.74 kcal/mol
dation [38,53]. In path I, both CO and O2 adsorb on the same Pt site. respectively. The conversion from IM2 (IM2*) to IM3 (IM3*) has to
In path II, the adsorption of CO and O2 occurs on the two Pt sites. In overcome TS2–3 (TS2 —3 ), the energy barrier is much lower than
order to probe whether bifunctional mechanism works for CO oxi- the one from IM1 (IM1*) to IM2 (IM2*). The overall potential energy
dation on these clusters, CO and O2 bind to Pt and M sites respec- curve is similar to that in Wang’s work except for slight structure
tively in path III. and energy difference caused by the choice of method.

Fig. 5. Optimized geometries of intermediates and transition states involved in the CO oxidation in paths I, II and III on Pt3Mo. Pt, Mo, C and O atoms are represented by blue,
turquoise, gray and red spheres. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
X. Lian et al. / Computational Materials Science 96 (2015) 237–245 241

Fig. 7. Optimized geometries of intermediates and transition states involved in the CO oxidation in paths I, II and III on Pt3Ru. Pt, Ru, C and O atoms are represented by blue,
light-sea green, gray and red spheres. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

state, the O–O bond breaks to form atomically adsorbed intermedi-


ate IM5. IM5 is a peroxide-like specie [54,55], where the O–O dis-
tance elongates to 1.51 Å and the net change on CO is still small.
This is similar to that on Pt4 while with smaller second reaction
barrier (12.12 kcal/mol). The O–O distance elongates to 1.91 Å
while C (CO)–O (O2) bond contracts to 1.27 Å in TS5–6. CO2 is
released upon further elongation of the O–O bond and IM6 forms,
IM6 is 93.16 kcal/mol more stable than the reactants.
In path II, CO and O2 approach two Pt atoms to form adsorption
complex IM4* with adsorption energy 64.45 kcal/mol, which is
more stable than IM4 by 11.53 kcal/mol. IM4* has to pass through
a five-membered ring transition state (TS4 —5 ) with an energy bar-
rier of 30.19 kcal/mol. Similar to path I, the formation of the per-
Fig. 8. Potential energy curve for CO oxidation promoted by Pt3Ru along path I oxo-like intermediate IM5* is exothermic by 39.73 kcal/mol. The
(black line), path II (red line) and path III (blue line). The energies include zero-point O–O bond lengthens further to transition state TS5 —6 with an
energy corrections. (For interpretation of the references to colour in this figure energy barrier of 9.25 kcal/mol. A CO2 molecule forms upon the
legend, the reader is referred to the web version of this article.)
formation of IM6*.
In path III, CO and O2 adsorb on Pt and Ni atoms respectively.
Adsorption of CO takes place on Pt atom while O2 in bridge position
on Ni atom to form IM4** intermediate, which is 78.79 kcal/mol
Pt3 M ðM ¼ Ni; Mo; Ru; Pd; RhÞ þ CO þ O2
more stable than the reactants. IM4** converts into IM5** via tran-
Figs. 3–12 show the predicted potential energy profiles and the sition state TS4 —5 with an energy barrier of 42.44 kcal/mol. With
structures of all stationary points for CO oxidation promoted by the distance of O–O getting longer, IM5** further evolves into IM6**,
Pt3Ni, Pt3Mo, Pt3Ru, Pt3Pd and Pt3Rh bimetallic clusters. a complex of Pt3NiO with CO2 with 88.84 kcal/mol more stable
Three models of CO and O2 co-adsorption on Pt3Ni are shown in than the reactants.
Figs. 3 and 4. In path I, CO and O2 initially are co-adsorbed on the The potential energy curve and the structure along the reaction
single Pt atom to form IM4 and the adsorption energy path of other clusters are very similar to those of Pt3Ni. The calcu-
(52.92 kcal/mol) is larger than that on Pt4. The CO bond length lated barriers of the first transition state along path I are
is 1.14 Å, nearly the same as that in the adsorption of CO, but the 20.58 kcal/mol, 20.29 kcal/mol and 21.02 kcal/mol for Pt3Rh, Pt3Ru
O–O bond slightly shrinks from 1.30 Å (single adsorption) to and Pt3Pd clusters, while those corresponding barriers into the
1.25 Å (co-adsorption). Conversion from IM4 to IM5 takes place path II are 41.05 kcal/mol, 30.83 kcal/mol and 34.93 kcal/mol
through a transition state TS4–5 with a barrier of 21.08 kcal/mol, respectively. Path I is energetically more favorable than path II.
slightly lower than that on Pt4. In TS4–5, the O–O bond elongates However, in path III (CO on Pt and O2 on the doped metal), the bar-
to 1.35 Å, and the distance of C (CO)–O (O2) is shortened to rier of first transition state is higher than those in path I and path II
1.65 Å. After overcoming a relatively high energy barrier transition for all clusters, i.e. the bifunctional mechanism is not favorable.
242 X. Lian et al. / Computational Materials Science 96 (2015) 237–245

Fig. 9. Optimized geometries of intermediates and transition states involved in the CO oxidation in paths I, II and III on Pt3Pd. Pt, Pd, C and O atoms are represented by blue,
teal, gray and red spheres. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

The reactions barriers of three paths for all clusters are summa-
rized in Fig. 13. For path I, the reactions barriers on all clusters are
similar. The barriers in path I are much lower than the other two
pathways except path III on Pt3Mo. For path II, although the barri-
ers of the transition states are higher than that in path I for all clus-
ters, the barriers in bimetallic clusters are much lower than that on
Pt4 except for Pt3Rh. The comparison of reaction barriers in path III
with those in paths I and II finds that bifunctional mechanism work
only on Pt3Mo.

3.3. Reaction rate constants

The pre-exponential factors and rate constants of a reaction


Fig. 10. Potential energy curve for CO oxidation promoted by Pt3Pd along path I can be determined according to transition-state theory (TST)
(black line), path II (red line) and path III (blue line). The energies include zero-point
[57,58]:
energy corrections. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)  
kB T Q TS DH i
kTST ¼ exp  ð2Þ
h Q IM RT

It is worth noting that in Pt3Mo, the first energy barriers in path kB T Q TS


I and path II are similar to the other clusters. However, in path III, A0;i ¼ ð3Þ
h Q IM
the energy barrier is only 15.41 kcal/mol and it is lower than that
in path I (19.54 kcal/mol) and path II (30.93 kcal/mol), the lowest where DHi is the enthalpy difference between the initial stateand
one among all barriers. The bifunctional mechanism occurs feasi- the transition state, kB is the Boltzmann constant, T is temperature;
bly on Pt3Mo. Mo atom prefers to adsorb O2 molecule and activates Q TS and Q IM are the partition functions of transition state and inter-
O2 at low energy when CO adsorbs on Pt site, and this facilitates mediate, R is the gas constant and h is the Planck constant.
the oxidation of CO. The predicted constants are presented in Table 2. In all cases,
On all those clusters, the reaction intermediates and transition the rate constants for the formation of the initial OCOO* interme-
states along the reaction path are all more stable than the reac- diate outbalance the rate constants for dissociation of the O–O
tants, thus that the reaction proceeds readily without thermal acti- bond from the OCOO*. A difference of more than 20 orders of
vation [56]. The reaction takes two steps [55]: the first step magnitude suggests that O2 scission to form OCOO* is the slow-
involves the O–O elongation and CO binding of CO to one O atom est step for CO oxidation. Thus, the dissociation of O–O is the
of O2 to form a peroxide-like (OOCO) intermediate, and the second rate-determining step for CO oxidation. The rate constants of
step involves the cleavage of O–O bond of peroxide to produce rate-determing step in path I are larger than those of paths II
CO2. and III by over 10 orders of magnitude indicating the path I is
X. Lian et al. / Computational Materials Science 96 (2015) 237–245 243

Fig. 11. Optimized geometries of intermediates and transition states involved in the CO oxidation in paths I, II and III on Pt3Rh. Pt, Rh, C and O atoms are represented by blue,
dark-cyan, gray and red spheres. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 12. Potential energy curve for CO oxidation promoted by Pt3Rh along path I
Fig. 13. The energy barrier of rate-determining step for CO oxidation on Pt3M
(black line), path II (red line) and path III (blue line). The energies include zero-point
(M = Pt, Ni, Mo, Ru, Pd, Rh).
energy corrections. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

to improve the performance of catalyst. As shown in Fig. 14, the


favored over paths II and III. The synergetic effect of binary bimetallic clusters have similar HOMO (highest occupied molecu-
alloys plays a significant role in CO oxidation reaction except lar orbital) and LUMO (lowest unoccupied molecular orbital) ener-
path III in Pt3Ni and path II in Pt3Rh according to the predicted gies and HOMO–LUMO energy gaps, thus they have similar
rate constants. The doping metal M (Ni, Mo, Ru, Pd, Rh) exhibits interactions with O2 to form adsorption complexes. The energy dif-
a positive ligand effect on the rate-determing step for CO oxida- ference [42] between E (LUMO, O2) and E (HOMO, Pt3Mo) is the
tion on Pt-based alloys, i.e. with improved catalytic activity. smallest and that between E (LUMO, Pt3Mo) and E (HOMO, O2) is
Pt3Mo has the highest rate constants of rate-determing step for larger than Pt4 and Pt3Pd. Therefore, the small energy difference
CO oxidation among all clusters [59] judged from the rate con- makes Pt3Mo easier to activate O2. The natural atomic charge for
stants listed in Table 2. O2 adsorption on bimetallic clusters is listed in Table 3. O2 gets
the largest amount of electrons when adsorbed on Pt3Mo, thus
3.4. Molecular orbital and nature bond orbital analysis with the highest activation on the O–O bond. The frontier molecu-
lar orbital and nature bond orbital (NBO) charge analyses reveal
The O–O bond breaking is the rate-determining step for CO oxi- the relative catalytic activities of those bimetallic clusters and
dation according to reaction barrier and reaction rate constant Pt3Mo is the best catalyst for CO oxidation among the clustered
analyses. Further analysis on the bonding of O2 with catalyst helps investigated.
244 X. Lian et al. / Computational Materials Science 96 (2015) 237–245

Table 2
Pre-exponential factor and rate constants for CO oxidation on Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh). (At 298 K, PCO = PO2 = 0.1 bar).

O2(ads) + CO(ads) ? OOCO(ads) OOCO(ads) ? CO2(ads) + O(ads)


A (1/s) K (1/s) A (1/s) K (1/s)
Pt4 Path I 1.18  1011 3.07  105 2.41  1014 3.94  105
Path II 7.43  1010 2.87  1018 6.68  1013 2.09  1011
Pt3Ni Path I 1.05  1011 3.36  105 2.80  1014 3.61  105
Path II 8.18  1010 5.90  1012 7.74  1013 1.27  107
Path III 1.52  1012 2.33  1020 9.50  1015 1.87  103
Pt3Mo Path I 1.82  1011 8.49  104 6.61  1013 2.18  106
Path II 5.10  1011 1.05  1011 2.22  1014 3.10  109
Path III 5.29  1010 0.264 9.05  1013 1.18  1011
Pt3Ru Path I 1.03  1011 1.35  104 1.25  1014 3.36  104
Path II 1.25  1012 3.06  1011 5.38  1013 7.51  108
Path III 3.50  1011 8.72  1015 3.08  1013 1.69  109
Pt3Pd Path I 1.21  1011 4.60  105 6.97  1014 1.01  106
Path II 3.58  1011 8.62  1015 3.20  1014 3.28  1010
Path III 1.75  1011 1.07  1017 4.22  1013 6.63  108
Pt3Rh Path I 1.03  1011 8.27  105 1.99  1014 1.88  106
Path II 1.04  1012 8.12  1019 1.83  1013 9.10  1010
Path III 1.03  1012 1.92  1014 1.47  1014 1.46  109

the bifunctional mechanism, to catalyze CO oxidation with high


efficiency. The high catalytic efficiency of Pt3Mo (possibly Mo
doped Pt cluster) on CO oxidation deserves experimental trial.

Acknowledgments

WQT thank supports from the State Key Lab of Urban Water
Resource and Environment (HIT) (2014TS01) and the Open Project
of State Key laboratory of Supramolecular Structure and Materials
(JLU) (SKLSSM201404).

References

Fig. 14. Energies of HOMOs and LUMOs of O2 and Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh). [1] B. Kalita, R.C. Deka, J. Am. Chem. Soc. 131 (2009) 13252–13254.
[2] K. Liu, A.Q. Wang, T. Zhang, ACS Catal. 2 (2012) 1165–1178.
[3] S. Kageyama, Y. Sugano, Y. Hamaguchi, J. Kugai, Y.J. Ohkubo, S. Seino, T.
Nakagawa, S. Ichikawa, T.A. Yamamoto, Mater. Res. Bull. 48 (2013) 1347–1351.
Table 3 [4] A. Ates, P. Pfeifer, O. Görke, Chem. Ing. Tech. 85 (2013) 664–672.
Natural atomic charges of O2 adsorbed on Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh). [5] M. Garcia-Dieguez, E. Iglesia, J. Catal. 301 (2013) 198–209.
[6] M.M.V.M. Souza, N.F.P. Ribeiro, M. Schmal, Int. J. Hydrogen Energy 32 (2007)
Pt4 Pt3Mo Pt3Ru Pt3Pd Pt3Ni Pt3Rh 425–429.
[7] S. Desai, M. Neurock, Electrochim. Acta 48 (2003) 3759–3773.
O2 0.343 0.711 0.439 0.428 0.465 0.326
[8] Y. Shimodaira, T. Tanaka, T. Miura, A. Kudo, H. Kobayashi, J. Phys. Chem. C 111
(2007) 272–279.
[9] Y.Z. Li, Y. Yu, J.G. Wang, J. Song, Q. Li, M.D. Dong, C.J. Liu, Appl. Catal., B 125
(2012) 189–196.
4. Conclusions [10] L.Q. Liu, F. Zhou, L.G. Wang, X.J. Qi, F. Shi, Y.Q. Deng, J. Catal. 274 (2010) 1–10.
[11] N. Todoroki, H. Osano, T. Maeyama, H. Yoshida, T. Wadayama, Appl. Sur. Sci.
256 (2009) 943–947.
Theoretical exploration on the reactivity of Pt3M (M = Pt, Ni, Mo, [12] S. Jaatinen, P. Salo, M. Alatalo, K. Kokko, Surf. Sci. 566 (2004) 1063–1066.
Ru, Pd, Rh) clusters toward CO oxidation, aiming at understanding [13] M. Ojeda, B.Z. Zhan, E. Iglesia, J. Catal. 285 (2012) 92–102.
[14] W. An, Y. Pei, X.C. Zeng, Nano Lett. 8 (2008) 195–202.
the improved catalytic activity of Pt-based bimetallic catalysts, has
[15] H.Y. Kim, H.M. Lee, G. Henkelman, J. Am. Chem. Soc. 134 (2012) 1560–1570.
been carried out in the present work. The adsorption of CO on Pt [16] D.Y. Wang, H.L. Chou, Y.C. Lin, F.J. Lai, C.H. Chen, J.F. Lee, B.J. Hwang, C.C. Chen,
and M sites of Pt3M (M = Pt, Ni, Mo, Ru, Pd, Rh) clusters through J. Am. Chem. Soc. 134 (2012) 10011–10020.
[17] W.L. Guo, W.Q. Tian, X. Lian, F.L. Liu, M. Zhou, P. Xiao, Y.H. Zhang, Comput.
three pathways was well-established. Although CO and O2 have
Theor. Chem. 1032 (2014) 73–83.
weaker adsorption on a single Pt site (path I) than path II and path [18] F.F. Zhao, C. Liu, P. Wang, S.P. Huang, H.P. Tian, J. Alloys Compd. 577 (2013)
III, the lower reaction barriers in path I make this path possible. 669–676.
The lower energies of all intermediates and transition states in [19] G. Samjeské, H.S. Wang, T. Löffler, H. Baltruschat, Electrochim. Acta 47 (2002)
3681–3692.
all three reaction paths warrant the processing of the reaction [20] L.C. Ordonez, P. Roquero, P.J. Sebastian, J. Ramirez, Int. J. Hydrogen Energy 32
without thermal activation. The O2 scission to form OCOO* is the (2007) 3147–3153.
rate-determining-step with high energy barrier (thermodynamic [21] J.M. Jaksic, L.J. Vracar, S.J. Neophytides, S. Zafeiratos, G. Papakonstantinou, N.V.
Krstajic, M.M. Jaksic, Surf. Sci. 598 (2005) 156–173.
control) and real low rate constants (kinetic control). Bimetallic [22] B.C. Vicentea, R.C. Nelsona, J. Singh, S.L. Scotta, J.A. Bokhoven, Catal. Today 160
clusters enhance the catalytic activity through lowering LUMO (2011) 137–143.
energy and raising the HOMO energy to match those of O2 and [23] C. Dupont, Y. Jugnet, D. Loffreda, J. Am. Chem. Soc. 128 (2006) 9129–9136.
[24] J. Kugai, T. Moriya, S. Seino, T. Nakagawa, Y.J. Ohkubo, H. Nitani, T.A.
inducing more charge transfer to O2. Different from the other Yamamoto, Int. J. Hydrogen Energy 38 (2013) 4456–4465.
bimetallic clusters, Pt3Mo takes path III (CO on Pt and O2 on Mo), [25] A. Sebetci, Comput. Mater. Sci. 58 (2012) 77–86.
X. Lian et al. / Computational Materials Science 96 (2015) 237–245 245

[26] J. Suntivich, Z.C. Xu, C.E. Carlton, J. Kim, B.H. Han, S.W. Lee, N. Bonnet, N. J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O.
Marzari, L.F. Allard, H.A. Gasteiger, K. Hamad-Schifferli, Y. Shao-Horn, J. Am. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K.
Chem. Soc. 135 (2013) 7985–7991. Morokuma, G.A. Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S.
[27] A. Velázquez-Palenzuela, P.L. Cabot, F. Centellas, J.A. Garrido, C. Arias, R.M. Dapprich, A.D. Daniels, M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck, K.
Rodríguez, E. Brillasi, Int. J. Hydrogen Energy 35 (2010) 11591–11600. Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G. Baboul, S. Clifford, J.
[28] F. Maillard, G.Q. Lu, A. Wieckowski, U. Stimming, J. Phys. Chem. B 109 (2005) Cioslowski, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R.L.
16230–16243. Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A. Nanayakkara, M.
[29] H. Igarashi, H. Uchida, M. Watanabe, Chem. Lett. 29 (2000) 1262–1263. Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, C. Gonzalez, J.A.
[30] A. Parinyaswan, S. Pongstabodee, A. Luengnaruemitchai, Int. J. Hydrogen Pople, Gaussian03, revision E.01, Gaussian Inc., Wallingford, CT, 2004.
Energy 31 (2006) 1942–1949. [44] A.D. Becke, Phys. Rev. A 38 (1988) 3098–3100.
[31] M. Krausa, W. Vielstich, J. Electroanal. Chem. 379 (1994) 307–314. [45] C.T. Lee, W.T. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789.
[32] T. Frelink, W. Visscher, J.A.R. Veen, Surf. Sci. 335 (1995) 353–360. [46] A.D. Becke, J. Chem. Phys. 98 (1993) 5648–5652.
[33] Y. Tong, H.S. Kim, P.K. Babu, P. Waszczuk, A. Wieckowski, E. Oldfield, J. Am. [47] P.J. Hay, W.R. Wadt, J. Chem. Phys. 82 (1985) 299.
Chem. Soc. 124 (2002) 468–473. [48] K. Raghavachari, J.S. Binkley, R. Seeger, J.A. Pople, J. Chem. Phys. 72 (1980) 650.
[34] J.C. Davies, J. Bonde, Á. Logadóttir, J.K. Nørskov, I. Chorkendorff, Fuel. Cells. 4 [49] C.M. Chang, C. Cheng, C.M. Wei, J. Chem. Phys. 128 (2008) 124710.
(2005) 429–435. [50] C. Li, S.X. Yang, S.S. Li, J.B. Xia, J.B. Li, J. Phys. Chem. C 117 (2013) 483–488.
[35] M. Watanabe, S.J. Motoo, J. Electroanal. Chem. 60 (1975) 259–266. [51] L. Guo, R.J. Zhang, L.L. Guo, S.S. Niu, Comput. Theor. Chem. 1036 (2014) 7–15.
[36] C. Roth, A.J. Papworth, I. Hussain, R.J. Nichols, D.J. Schiffrin, J. Electroanal. [52] F.R. Negreiros, L. Sementa, G. Barcaro, S. Vajda, E. Aprá, A. Fortunelli, ACS Catal.
Chem. 581 (2005) 79–85. 2 (2012) 1860–1864.
[37] J.L. Gómez de la Fuente, F.J. Pérez-Alonso, M.V. Martínez-Huerta, M.A. Peña, [53] T. Davran-Candan, A.E. Aksoylu, R. Yildirim, J. Mol. Catal. A: Chem. 306 (2009)
J.L.G. Fierro, S. Rojas, Catal. Today 143 (2009) 69–75. 118–122.
[38] F. Wang, D.J. Zhang, Y. Ding, J. Phys. Chem. C 114 (2010) 14076–14082. [54] Y. Gao, N. Shao, S. Bulusu, X.C. Zeng, J. Phys. Chem. C 112 (2008) 8234–8238.
[39] E.F. Fialko, A.V. Kikhtenko, V.B. Goncharov, K.I. Zamaraev, J. Phys. Chem. B 101 [55] C.Y. Liu, Y.Z. Tan, S.S. Lin, H. Li, X.J. Wu, L. Li, Y. Pei, X.C. Zeng, J. Am. Chem. Soc.
(1997) 5772–5773. 135 (2013) 2583–2595.
[40] W.T. Wallace, R.L. Whetten, J. Am. Chem. Soc. 124 (2002) 7499–7505. [56] H.T. Chen, J.G. Chang, S.P. Ju, H.L. Chen, J. Comput. Chem. 31 (2010) 258–265.
[41] S.J. Li, X. Zhou, W.Q. Tian, J. Phys. Chem. A 116 (2012) 11745–11752. [57] A.A. Gokhale, S. Kandoi, J.P. Greeley, M. Mavrikakis, J.A. Dumesic, Chem. Eng.
[42] Z.F. Xu, Y.X. Wang, J. Phys. Chem. C 115 (2011) 20565–20571. Sci. 59 (2004) 4679–4691.
[43] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, [58] P.R.P. Barreto, A.F.A. Vilela, R. Gargano, J. Mol. Struct. – Theochem. 639 (2003)
J.J.A. Montgomery, T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S. Iyengar, J. 167–176.
Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson, [59] S.M.M. Ehteshami, Q.Y. Jia, A. Halder, S.H. Chan, S. Mukerjee, Electrochim. Acta
H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. 107 (2013) 155–163.
Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J.E. Knox, H.P. Hratchian,

You might also like