You are on page 1of 6

Article

pubs.acs.org/JPCC

CO Oxidation over Strained Pt(100) Surface: A DFT Study


Fuzhu Liu,† Chao Wu,*,‡ Guang Yang,§ and Shengchun Yang*,†,⊥

School of Science, Key Laboratory of Shaanxi for Advanced Materials and Mesoscopic Physics, State Key Laboratory for Mechanical
Behavior of Materials, Xi’an Jiaotong University, Xi’an 710049, People’s Republic of China

Frontier Institute of Science and Technology, Xi’an Jiaotong University, Xi’an 710054, People’s Republic of China
§
Electronic Materials Research Laboratory, Key Laboratory of the Ministry of Education & International Center for Dielectric
Research, Xi’an Jiaotong University, Xi’an 710049, People’s Republic of China

Collaborative Innovation Center of Suzhou Nano Science and Technology, Suzhou Academy of Xi’an Jiaotong University, 215000,
Suzhou, People’s Republic of China
*
S Supporting Information

ABSTRACT: The oxidation of CO on strained Pt(100) surface was studied


using periodic density functional theory (DFT). Unlike the uniform response of
global properties (e.g., d-band center) to strain, the localized nature of
adsorption leads to complex site-dependent and adsorbate-dependent responses,
invalidating the generally believed statement of “tension strengthens binding”.
Moreover, the complex responses of reaction energetics to strain require direct
study of the reaction under strain rather than extrapolating the known behaviors
of individual adsorbates under strain or reaction energetics on unstrained
surfaces. We show that the tensile strain lowers the reaction barrier of CO
oxidation over the Pt(100) surface. This work provides a theoretical basis of
utilizing strain to improve the Pt catalysts with a higher tolerance toward CO
poisoning.

A number of important chemical reactions catalytically


occur on transition metal surfaces. Various methods,
including alloying/dealloying,1 surface masking with heter-
and upshift of its d states, in accordance with the enhanced CO
dissociation reactivity.4 Yu et al. further showed that the
continuous shift of the d-band center away from the Fermi level
oatoms,2 morphology control, or defect engineering,3 have triggers the weakening of O adsorption on the compressed
been developed to improve their catalytic performance by Pt(111) and Pt skins of Pt3Co(111) alloy compared with the
modifying the surface structure. In almost all the methods, unstrained surface.19 Gradually, “tension strengthens binding”
surface strain of the catalysts is introduced and/or varied summarized from theoretical predictions become a widely
through mechanisms like local deformation of a single metal accepted trend. 20 However, this rule is not always
phase,4 epitaxial metal overlayers,4−6 defects,7 and lattice observed.21−23 For example, Liu’s group found that the
mismatch between core and shell.8,9 The geometric and response of chemisorption of CO on the Au(001) and
electronic structures of surface atoms are perturbed, and K(001) surfaces to external strain is site-dependent:24 for
consequently, surface catalytic reactions are affected.10−12 In either metal surface, two oppositely going linear relationships
the past decade, an emerging field of strain engineered catalysis, between the CO adsorption energy and strain exist for the on-
aiming at regulating the catalytic performance of transition top and hollow sites, respectively. Similarly, Groß et al.
metals through strain, has attracted a lot of attention.3,7,13−18 discovered that for the same type of site on low-index surfaces
Strasser and co-workers observed high catalytic activity of of different metals like Pd5 and Cu6 the H adsorption energy
dealloyed Pt−Cu core−shell nanoparticles in oxygen-reduction and tensile strain exhibit opposite linear relationships.
reaction (ORR), which is attributed to surface strain induced by In principle, strain adds an additional dimension of tunability
lattice mismatch of the core and shell metals.1 Yang group on top of the traditional catalytic variables including the types
reported that icosahedral platinum alloy (Pt−Au/Ni/Pd) of metals, surfaces, and surface sites, whose different responses
nanoparticles presented higher ORR activity than octahedral to strain may be utilized to regulate the catalytic reactions. In
ones, as the tensile strain in the former facilitated the reaction this work, as an example, we try to solve or at least mitigate the
while the compressive strain in the latter hindered it.3 well-known problem of “CO poisoning” over Pt utilizing strain.
Several theoretical studies have successfully provided a CO poisoning takes place in many catalytic systems, such as
concise and insightful description of the strain effects on
catalysis by correlating reactivity and strain via the d-band Received: May 11, 2015
center shift of the catalyst.3 Nørskov et al. demonstrated that Revised: June 6, 2015
for Ru(0001) tensile strain leads to surface lattice expansion Published: June 8, 2015

© 2015 American Chemical Society 15500 DOI: 10.1021/acs.jpcc.5b04511


J. Phys. Chem. C 2015, 119, 15500−15505
The Journal of Physical Chemistry C Article

polymer electrolyte fuel cells and direct methanol fuel cells,12 To interpret the strain effects on binding strength, the d-band
which is caused by strong adsorption of CO on Pt surfaces, center model was employed.35−37 In addition, the adsorption
blocking the active sites and reducing the catalytic activity of Pt. energy is defined as ΔE = E(ads/metal) − E(metal) − E(ads),
Thus, an important and urgent issue is how to readily remove where E(ads/metal), E(metal), and E(ads) are the energies of
CO from Pt surfaces or how to improve the tolerance of CO the adsorbate−slab system, the slab, and the adsorbate,
poisoning at Pt surfaces? Based on the idea of surface strain respectively.
engineering, several techniques such as local deformation7 and Adsorbates and surfaces mainly interact through mixing the
epitaxial metal overlayers25−27 have been developed to regulate valence orbital of the adsorbates and the d states of the surfaces.
the binding strength of CO. However, a number of questions The latter is concisely represented by the d-band center model
about the atomistic behaviors of the relevant adsorbates still with the d-band center energy denoted as εd.38 The d-band
remain unclear. (1) How does strain affect multiple species center model has been widely used to rationalize the catalytic
involved in CO oxidation (e.g., O, CO) at multiple sites? (2) Is activity trends of transition metals and alloys,4,19,39,40 which is
CO−Pt binding strength the only key parameter to control? also employed here to explain how strain modifies the
(3) For the coadsorption of CO and O, required during CO electronic structure of the Pt(100) surface and subsequently
oxidation, how does the adsorption of one species affect the its interactions with adsorbates.
other? (4) Whether the tolerance of CO poisoning at Pt The d-band center (εd) of the Pt(100) surface rises linearly
surfaces can be improved by strain? with the increasing tensile strain (Figure 1). As the atoms are
Here we use the system of CO oxidation over the Pt(100)
surface as an example model to study the strain effects. First,
using the periodic density functional theory (DFT), we
systematically investigate the adsorption behaviors of key
species (e.g., O and CO) at multiple sites (e.g., top, bridge,
and 4-fold hollow sites) under compressive and tensile strains
(−3% to 3%). Next, we check if the response of coadsorption
to strain can be expressed as a linear combination of the
responses of each involved species. Finally, we probe the
variation of the energetics of CO oxidation under strain, thus
verifying if the higher tolerance of CO poisoning at Pt surfaces
can be achieved through surface strain engineering.
The spin unrestricted density functional theory (DFT)
calculations were carried out using the Dmol3 program
package.28,29 The generalized gradient approximation (GGA)
with the Perdew and Wang-91 (PW91) formulation of the
exchange-correlation functional was employed.30,31 The self-
consistent PW91 density was determined by iterative
diagonalization of the Kohn−Sham Hamiltonian.32 The valence Figure 1. The d-band center of the Pt(100) surface as a function of
electron wave functions were expanded into a set of atomic surface strain. The solid line only serves as a guide for eye. Top insets:
orbitals composed of the double numerical plus d-functions top views of the electron density plots of the −3% compressed (left),
(DND) basis set. Brillouin-zone integrations were performed equilibrium (middle), and 3% stretched (right) Pt(100) surfaces. The
on a grid of (8 × 8 × 1) Monkhorst−Pack k-point mesh. The color bar from red to blue indicates the reduction of the electron
width of the Fermi smearing of the Kohn−Sham states was set density. Bottom inset: top view of the Pt(100) surface. The top (t),
to kBT = 0.005 hartree. The Pt(100) surface was modeled by bridge (b), and 4-fold (4f) sites are marked by the dark red circles.
Dark and light gray circles represent the sublayer and toplayer Pt
using a (2 × 2) five-layer slab supercell with the bottom three
atoms, respectively.
layers frozen. Each slab was separated by a vacuum of 15 Å to
minimize the interactions between images. The convergence
criteria for geometry optimization were set to 10−5 hartree, moved away from their stable positions by the tensile strain, the
0.002 hartree/Å, and 0.005 Å for energy, force, and overlaps of the d states among atoms are reduced and the d-
displacement, respectively. Compressed and tensile strain bandwidth becomes narrower. Subsequently, the d-band shifts
ranging from −3% to 3% was imposed in parallel to the upwardly,15,20 which is evidenced in the electron density plots10
surface plane by changing the lattice constant of the unstrained (Figure 1, insets). The εd shift range is ∼0.2 eV from −3% to
Pt(100) surface. This approach has been shown to yield 3% strain, suggesting a fairly wide tunability of the surface
accurate estimates of adsorption energies on strained transition electronic structure; thus, it is interesting to see how the
metal surfaces.11 The calculated equilibrium lattice constant of adsorbates (e.g., O and CO) respond to this wide εd change.
3.924 Å for Pt was in agreement with the experimental value of Strain may cause various responses among different
3.920 Å.33 O2 and CO molecules were relaxed by placing them adsorbates. The stronger binding of O atom, CO, and NO
individually in a rectangular cuboid box with a vacuum space of molecules over stretched Ru(0001) surface was observed
at least 15 Å in all three directions to minimize image experimentally,7,41 while the reduction in binding strength of
interactions. The bond lengths were 1.224 Å (O2) and 1.140 Å O atom on compressed Pt(111) was rationalized as a result of
(CO), very close to the experimental values of 1.210 Å19 and the shift of d-band center away from the Fermi level.19 The
1.120 Å,4 respectively. To investigate the minimum-energy adsorption sites on the Pt(100) surface are defined in the
pathway for O adatom diffusion and CO oxidation on the bottom inset of Figure 1. For O adatom, our calculations show
strained Pt(100) surface, linear synchronous transit/quadratic that the 4f site is the most stable site, the b site takes second
synchronous transit (LST/QST) approaches were invoked.34 place, and the t site is less stable(Table S1), confirming the
15501 DOI: 10.1021/acs.jpcc.5b04511
J. Phys. Chem. C 2015, 119, 15500−15505
The Journal of Physical Chemistry C Article

Figure 2. O adsorption energy at the 4f and b sites on the Pt(100) surface versus strain. (a) O adsorption energy at the 4f and b sites. Insets: top
views of O adsorption configurations. The t site O is unstable and migrates to the 4f site (indicated by the black arrow in the bottom left inset). (b)
CO adsorption at the t, 4f, and b sites. Insets: top views of CO adsorption configurations. The solid lines only serve as a guide for eye. The color
coding of the circles is as in Figure 1.

previous studies.42,43 As illustrated in Figure 2a, with the adsorption energy increases along with the rising tensile strain.
increase of the tensile strain, the O adsorption energy difference Particularly, for the latter (the t site), the tensile strain
between the 4f and b sites decreases from 0.29 eV at −3% strain introduces evident nonlinear response of the adsorption energy.
to 0.14 eV at 3% strain. For the b site, the O adsorption energy However, an opposite trend appears for the b site, quite
and the tensile strain (or the lattice constant expansion) different from the previous report of Au(001) and K(001)
relationship is monotonic: more exothermic adsorption systems,24 which may be caused by the different d-states of the
corresponds to larger tensile strain, which needs two linear metal surfaces at the Fermi level.36 The calculation results also
segments with similar slopes for compressed and stretched reveal that the Pt(100) surface, like the stepped Cu(211) and
regions to describe accurately. For the 4f site, the flat line in the Ni(211) surfaces, exhibits complex CO binding response to
compressive strain range reflects the insensitiveness of the 4f external strain, emphasizing the oversimplicity of the rule of
site to compression. Usually, O adatoms prefer a large “tension strengthens binding”.20
coordination number when adsorbed on a metal surface.42 The CO oxidation over the lateral strained Pt(100) surface
For the 4f site, tensile strain counteracts the lateral contraction follows the Langmuir−Hinshelwood (L−H) mechanism,44
caused by the O adsorption and thus weakens the adsorption. which requires the coadsorption of CO and O. We focused
For the b site, tensile strain promotes the O adatom’s bridging on the coadsorption energy (ECO+O(ε)) as a function of tensile
coordination to the Pt atoms, where the distance between the strain (ε) on the Pt(100) surface (eq 1).
O and Pt atoms decreases from 2.027 to 2.024 Å, strengthening
the adsorption from −3.639 to −3.678 eV (Table S1). The very ECO + O(ε) = E0 + k × ε (1)
different responses of O adatom at the 4f and b sites to strain
E0 represents the coadsorption energy of CO and O at zero
might serve as a handle for strain engineered catalysis.
strain and k is the slope. ECO+O can be obtained from either
Next, we studied the CO adsorption on the Pt (100) surface
direct DFT calculations of coadsorbed configurations or by
at the 4f, t, and b sites (Figure 2b and Table 1), where the C−
summing up the individual adsorption energies of CO molecule
and O atom over the strained surface.
Table 1. CO Adsorption Energy (Ead) and the C−O Bond The directly calculated coadsorption energy (Figure 3a, the
Length (d(C−O)) on the Pt(100) Surface black line and squares) becomes more exothermic linearly with
4f site b site t site the rising tensile strain. The summed-up coadsorption energy
strain d(C−O) d(C−O) d(C−O)
can be computed using the most stable adsorption site of each
(%) Ead (eV) (Å) Ead (eV) (Å) Ead (eV) (Å) individual adsorbate, i.e., the b site for CO and the 4f site for O
−3 −1.509 1.189 −1.589 1.169 −1.200 1.152 (Figure 3a, the dashed blue line and triangles). For the
−2 −1.512 1.190 −1.608 1.170 −1.207 1.153 summed-up coadsorption energy and strain, the opposite trend
−1 −1.517 1.189 −1.624 1.171 −1.210 1.153 is observed. The values of E0’s are markedly different by over
0 −1.516 1.189 −1.638 1.171 −1.217 1.153 ∼0.4 eV, with the directly calculated E0 being the higher one.
1 −1.502 1.190 −1.644 1.171 −1.241 1.153 Apparently, the adsorption energy of coadsorbed species is not
2 −1.481 1.190 −1.646 1.171 −1.212 1.153 a simple linear combination of that of each adsorbate at its most
3 −1.466 1.188 −1.650 1.172 −1.199 1.154 stable site. In fact, on Pt(100), the most stable coadsorption
configuration of CO and O before oxidation occupies two
neighboring b sites. At the presence of CO, the O adatom is
Pt coordination is favored and the b site is the most stable pushed away from its most stable 4f site by repulsive lateral
binding position, in agreement with the previous study.12 interactions. It implies that O adatoms could be easily relocated
Overall, the CO adsorption is much weaker than that of the O with the help of CO molecules.45−47 Therefore, selecting
adsorption by ∼2.3 eV. Calculations reveal that the CO adsorption energy at the b site for O adatom instead of that at
adsorption energy at different sites changes very differently with the 4f site is more reasonable for the summing-up scheme. This
strain. All sites show slight rises in adsorption energies (less time, the summed-up coadsorption energy (Figure 3a, the red
exothermic) to compressive strain, suggesting a weakened line and dots) presents the same trend as the directly calculated
adsorption. For the 4f and t sites on a stretched surface, the ones.
15502 DOI: 10.1021/acs.jpcc.5b04511
J. Phys. Chem. C 2015, 119, 15500−15505
The Journal of Physical Chemistry C Article

H diffusion on the TiN(100) surface, there are two typical


migration pathways for O adatoms to move over the surface:
the 4f−4f and 4f-b-4f paths (Figure 4a).51 For unstrained

Figure 3. Coadsorption of CO and O on Pt(100) surface with tensile


strain. (a) Coadsorption energies (ECO+O(ε)) versus strain. ECO+O(ε)
are calculated using coadsorbed configuration (Directly calculated,
black), using both b-site adsorption energies for CO and O (Summed-
up (b+b), red), and using 4f site and b site adsorption energies for CO Figure 4. Diffusion pathways of an O adatom on the Pt(100) surface.
and O (Summed-up (4f+b), blue). (b) Extracted CO and O lateral (a) On the unstrained surface, diffusion pathways from a 4f site to its
interactions versus strain. Directly subtracted (E1CO+O − E2CO+O, red) neighboring 4f site (black) directly or via the b site (red). Insets: top
and indirectly subtracted (black). To get the latter, two steps are views of the diffusion paths. Dashed-open, pink-filled, and solid-open
involved. (1) The individual adsorption energy is calculated by circles represent the initial, transition (TS), and final states,
freezing the relaxed coadsorption structures and removing one respectively. The color coding of other circles is as in Figure 1. (b)
adsorbate (keeping the other) at a time. (2) The lateral interaction Minimum-energy paths (4f-b-4f) on the −3%, 0%, and 3% strained
is obtained by subtracting these energies from the total coadsorption Pt(100) surface.
energy. Note: the coadsorption configuration of CO and O on the
Pt(100) surface is presented in Figure S1.
surface, the diffusion barrier via the 4f−4f path is 0.275 eV
higher than that of the 4f-b-4f path by 0.048 eV. A local
Furthermore, the lateral interactions of coadsorbed CO and minimum appears in the latter when the O adatom moves onto
O can be extracted (Figure 3b). A rough estimation can be the b site, thus lowering the barrier than directly crossing the b
done (the dashed red line) by directly subtracting the summed- site. The minimum-energy paths of O diffusion under strains
up coadsorption energy using the b-site adsorption energies for are plotted in Figure 4b (for details see Table S2). A distinct
both CO and O (Figure 3a, the red line and dots) from the decrease in transition barrier by ∼0.165 eV is observed when
directly calculated ones. The lateral interactions are ∼0.2 eV the tensile strain increases from −3% to 3%, consistent with the
repulsive, and they drop slightly as the tensile strain rises. A tensile strain induced weakening of O adsorption at the 4f site
more accurate way (“the indirect subtraction”) to obtain the mentioned above. However, for the coadsorption case, the
lateral interactions is by using the frozen structures under strain most stable configuration has CO and O both adsorbed at two
with either CO or O removed, whose energies are used to neighboring b sites, and the relocation of the latter from the 4f
extract the lateral interactions (Figure 3b, the dashed black line to b site is necessary (Figure 4); thus, an easier diffusion of O
and dots). Both curves in Figure 3b have the same trend and atom could facilitate the reaction of CO oxidation under tensile
quantitatively close (differ by ∼0.03 eV). These results suggest strain.
that the trend of coadsorption energy can be quantitatively The lack of unanimous response of local properties (e.g.,
reproduced by the linear combination of the adsorption energy adsorbates at various sites) to the perturbation introduced to
of individual components at the near coadsorption config- global properties (e.g., strain or the d-band center shift) is
uration rather than the most stable configuration by itself. expected to alter the energetics of a reaction path. Figure 5
At least at low O coverage, O2 automatically dissociates on sketches the potential energy profiles of CO oxidized by O
low-index Pt surfaces such as (111), (100), and (321).48−50 adatoms via the L−H mechanism44 on the strained Pt(100)
Although O adsorption is much stronger than CO adsorption, surfaces. The activation energy barrier from the initial state to
their coadsorption has exemplified the change of favorable the transition state for the stretched surface (3%) is below that
binding site of O. Thus, the mobility of O adatoms over of the unstrained one by ∼0.29 eV, while the barrier of the
strained surface may be crucial in many reactions including CO compressed surface (−3%) is higher by ∼0.57 eV. For the
oxidation.42 The diffusion of an O adatom from one 4f site to a transition state of the stretched surface (3%), the O atom
nearby 4f site is investigated under different strains. Similar to moves slightly to the 4f site compared with the originally
15503 DOI: 10.1021/acs.jpcc.5b04511
J. Phys. Chem. C 2015, 119, 15500−15505
The Journal of Physical Chemistry C Article


*
ASSOCIATED CONTENT
S Supporting Information
Adsorption energy of O atom and O2 and the bond length of
O2 (Table S1); the energy difference and diffusion barrier of O
atom under strained surface (Table S2); the configuration of O
and CO coadsorption on Pt surface (Figure S1). The
Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.jpcc.5b04511.

■ AUTHOR INFORMATION
Corresponding Authors
Figure 5. Energy profiles of CO oxidation on the Pt(100) surface. IS, *E-mail: ysch1209@mail.xjtu.edu.cn (S.Y.).
TS, and FS refer to the initial, transition, and final states, respectively. *E-mail: chaowu@mail.xjtu.edu.cn (C.W.).
The activation energies are labeled. Insets: the geometry of the key
states during the CO oxidation on the −3%, 0%, and 3% strained Notes
The authors declare no competing financial interest.


Pt(100) surface. The color coding of the circles is as in Figure 1.

ACKNOWLEDGMENTS
This work is supported by National Natural Science
favored b site in the initial state, whereas the CO molecule tilts Foundation of China (No. 51271135, 21203143), the program
about 19° relative to the vertical configuration. The distance for New Century Excellent Talents in university (No. NCET-
between the C atom and the O adatom is 2.712 Å, which is 12-0455), the Fundamental Research Funds for the Central
smaller than the initial state by 0.328 Å (Figure 5, inset). The Universities, the Natural Science Foundation and the project of
big variation in reaction energetics is the collective results of Innovative Team of Shanxi Province (No. 2015JM5166 and
2013KCT-05).


different responses of the adsorbed O and CO to strain,
invalidating the simple statement that all adsorptions are
REFERENCES
strengthened by tensile strain. In addition, our results are in line
(1) Strasser, P.; et al. Lattice-Strain Control of the Activity in
with the experimental and theoretical observations that the
Dealloyed Core-Shell Fuel Cell Catalysts. Nat. Chem. 2010, 2, 454−
activity of nanoparticles with tensile strained surface is higher 460.
than that of compressed surface.3,7 (2) Xie, S.; et al. Atomic Layer-by-Layer Deposition of Pt on Pd
In summary, we have studied CO oxidation over the Pt(100) Nanocubes for Catalysts with Enhanced Activity and Durability toward
surface under surface strain (−3% to 3%) using the periodic Oxygen Reduction. Nano Lett. 2014, 14, 3570−6.
DFT calculations. Through investigating this model system, we (3) Wu, J.; Qi, L.; You, H.; Gross, A.; Li, J.; Yang, H. Icosahedral
hope to utilize strain to help solve the problem of CO Platinum Alloy Nanocrystals with Enhanced Electrocatalytic Activities.
J. Am. Chem. Soc. 2012, 134, 11880−11883.
poisoning over Pt. Our main observations are summarized as (4) Mavrikakis, M.; Hammer, B.; Norskov, J. K. Effect of Strain on
(i) in contrast to the single linear relationship between the the Reactivity of Metal Surfaces. Phys. Rev. Lett. 1998, 81, 2819−2822.
upward shift of the d-band center and the increasing strain (5) Roudgar, A.; Gross, A., Local Reactivity of Metal Overlayers:
(from compressive to tensile), the adsorption energies of key Density Functional Theory Calculations of Pd on Au. Phys. Rev. B
reactants (e.g., O adatom and CO molecule) exhibit bi- or 2003, 67.
multisegment correlations to strain. These complex site- (6) Sakong, S.; Gross, A. Dissociative Adsorption of Hydrogen on
Strained Cu Surfaces. Surf. Sci. 2003, 525, 107−118.
dependent responses of different adsorbates to strain are (7) Wintterlin, J.; Zambelli, T.; Trost, J.; Greeley, J.; Mavrikakis, M.
rooted in the localized nature of adsorption events. For Atomic-Scale Evidence for an Enhanced Catalytic Reactivity of
example, the adsorption strengths of O adatom and CO Stretched Surfaces. Angew. Chem., Int. Ed. 2003, 42, 2850−2853.
molecule on their most stable binding sites present opposite (8) Strasser, P.; et al. Lattice-Strain Control of the Activity in
changes over the same strained surfaces. (ii) For coadsorption Dealloyed Core-Shell Fuel Cell Catalysts. Nat. Chem. 2010, 2, 454−60.
(9) Wang, Z.; Chen, Z.; Zhang, H.; Zhang, Z.; Wu, H.; Jin, M.; Wu,
of CO and O, the lateral interactions relocate the O away from
C.; Yang, D.; Yin, Y. Lattice-Mismatch-Induced Twinning for Seeded
its most stable binding site; thus, linearly combining the Growth of Anisotropic Nanostructures. ACS Nano 2015, 9, 3307−13.
information on each involved adsorbate at its most stable (10) Ruban, A.; Hammer, B.; Stoltze, P.; Skriver, H. L.; Norskov, J. K.
binding site does not qualitatively reproduce the coadsorption. Surface Electronic Structure and Reactivity of Transition and Noble
(iii) The O diffusion barrier decreases on the tensile strained Metals. J. Mol. Catal. A: Chem. 1997, 115, 421−429.
surface, which may facilitate the CO oxidation. (iv) The (11) Grabow, L.; Xu, Y.; Mavrikakis, M. Lattice Strain Effects on Co
Oxidation on Pt(111). Phys. Chem. Chem. Phys. 2006, 8, 3369−74.
activation energy of CO oxidation on the strained Pt (100)
(12) Tsuda, M.; Kasai, H. Ab Initio Study of Alloying and Straining
surface is evidently reduced compared to unstrained surface, Effects on Co Interaction with Pt. Phys. Rev. B 2006, 73, 155405.
which is desired for curing the CO poisoning. (v) The complex (13) Kibler, L. A.; El-Aziz, A. M.; Hoyer, R.; Kolb, D. M. Tuning
responses of reaction energetics to strain require direct study of Reaction Rates by Lateral Strain in a Palladium Monolayer. Angew.
the reaction under strain rather than extrapolation of known Chem., Int. Ed. 2005, 44, 2080−2084.
behaviors of individual adsorbates under strain or reactions on (14) Jiang, Q.; Liang, L. H.; Zhao, D. S. Lattice Contraction and
Surface Stress of Fcc Nanocrystals. J. Phys. Chem. B 2001, 105, 6275−
unstrained surfaces. We have shown that the tensile strain can
6277.
be utilized to enhance the reactivity of CO oxidation over the (15) Wu, J.; Li, P.; Pan, Y.-T.; Warren, S.; Yin, X.; Yang, H. Surface
Pt(100) surface; thus, strain is a useful handle to remedy or Lattice-Engineered Bimetallic Nanoparticles and Their Catalytic
even solve the CO poisoning problem. Properties. Chem. Soc. Rev. 2012, 41, 8066−8074.

15504 DOI: 10.1021/acs.jpcc.5b04511


J. Phys. Chem. C 2015, 119, 15500−15505
The Journal of Physical Chemistry C Article

(16) Yang, J.; Yang, J.; Ying, J. Y. Morphology and Lateral Strain (38) Hammer, B.; Norskov, J. K. Electronic Factors Determining the
Control of Pt Nanoparticles via Core-Shell Construction Using Alloy Reactivity of Metal Surfaces [1995, 343, 211]. Surf. Sci. 1996, 359, 306.
AgPd Core toward Oxygen Reduction Reaction. ACS Nano 2012, 6, (39) Wang, C.-M.; Fan, K.-N.; Liu, Z.-P. Origin of Oxide Sensitivity
9373−9382. in Gold-Based Catalysts: A First Principle Study of Co Oxidation over
(17) Greeley, J.; Krekelberg, W. P.; Mavrikakis, M. Strain-Induced Au Supported on Monoclinic and Tetragonal ZrO2. J. Am. Chem. Soc.
Formation of Subsurface Species in Transition Metals. Angew. Chem., 2007, 129, 2642−2647.
Int. Ed. 2004, 43, 4296−300. (40) Schnur, S.; Gross, A., Strain and Coordination Effects in the
(18) Dong, N.; Zhang, C.; Liu, H.; Li, J.; Wu, X.; Han, P. Stress Adsorption Properties of Early Transition Metals: A Density-
Effects on Stability and Diffusion Behavior of Sulfur Impurity in Functional Theory Study. Phys. Rev. B 2010, 81.
Nickel: A First-Principles Study. Comput. Mater. Sci. 2014, 90, 137− (41) Gsell, M.; Jakob, P.; Menzel, D. Effect of Substrate Strain on
142. Adsorption. Science 1998, 280, 717−720.
(19) Xu, Y.; Ruban, A. V.; Mavrikakis, M. Adsorption and (42) Pedersen, M. O.; Osterlund, L.; Mortensen, J. J.; Mavrikakis, M.;
Dissociation of O-2 on Pt-Co and Pt-Fe Alloys. J. Am. Chem. Soc. Hansen, L. B.; Stensgaard, I.; Laegsgaard, E.; Norskov, J. K.;
2004, 126, 4717−4725. Besenbacher, F. Diffusion of N Adatoms on the Fe(100) Surface.
(20) Francis, M. F.; Curtin, W. A. Mechanical Work Makes Phys. Rev. Lett. 2000, 84, 4898−4901.
Important Contributions to Surface Chemistry at Steps. Nat. Commun. (43) Bogicevic, A.; Stromquist, J.; Lundqvist, B. I. First-Principles
2015, 6, 6261−6261. Diffusion-Barrier Calculation for Atomic Oxygen on Pt(111). Phys.
(21) Andersson, M. P.; Bligaard, T.; Kustov, A.; Larsen, K. E.; Rev. B 1998, 57, R4289−R4292.
Greeley, J.; Johannessen, T.; Christensen, C. H.; Norskov, J. K. (44) Yuan, D. W.; Liu, Z. R.; Chen, J. H., Catalytic Activity of Pd
Toward Computational Screening in Heterogeneous Catalysis: Pareto- Ensembles over Au(111) Surface for Co Oxidation: A First-Principles
Optimal Methanation Catalysts. J. Catal. 2006, 239, 501−506. Study. J. Chem. Phys. 2011, 134.
(22) Norskov, J. K.; Bligaard, T.; Rossmeisl, J.; Christensen, C. H. (45) Rzeszotarski, P.; Kaszkur, Z. Surface Reconstruction of Pt
Towards the Computational Design of Solid Catalysts. Nat. Chem. Nanocrystals Interacting with Gas Atmosphere. Bridging the Pressure
2009, 1, 37−46. Gap with in Situ Diffraction. Phys. Chem. Chem. Phys. 2009, 11, 5416−
(23) Zhang, L.; Iyyamperumal, R.; Yancey, D. F.; Crooks, R. M.; 5421.
Henkelman, G. Design of Pt-Shell Nanoparticles with Alloy Cores for (46) Tao, F.; Dag, S.; Wang, L.-W.; Liu, Z.; Butcher, D. R.; Salmeron,
the Oxygen Reduction Reaction. ACS Nano 2013, 7, 9168−9172. M.; Somorjai, G. A. Restructuring of Hex-Pt(100) under Co Gas
(24) Pala, R. G.; Liu, F. Determining the Adsorptive and Catalytic Environments: Formation of 2-D Nanoclusters. Nano Lett. 2009, 9,
Properties of Strained Metal Surfaces Using Adsorption-Induced 2167−2171.
Stress. J. Chem. Phys. 2004, 120, 7720−4. (47) Carenco, S. Carbon Monoxide-Induced Dynamic Metal-Surface
(25) Xu, C.; Goodman, D. W. Adsorption and Reaction of Formic Nanostructuring. Chem.Eur. J. 2014, 20, 10616−10625.
Acid on a Pseudomorphic Palladium Monolayer on Mo(110). J. Phys. (48) Liu, D.-J.; Evans, J. W., Dissociative Adsorption of O-2 on
Chem. 1996, 100, 245−252. Unreconstructed Metal (100) Surfaces: Pathways, Energetics, and
(26) Schlapka, A.; Lischka, M.; Gross, A.; Kasberger, U.; Jakob, P. Sticking Kinetics. Phys. Rev. B 2014, 89.
Surface Strain Versus Substrate Interaction in Heteroepitaxial Metal (49) McEwen, J. S.; Bray, J. M.; Wu, C.; Schneider, W. F. How Low
Layers: Pt on Ru(0001). Phys. Rev. Lett. 2003, 91, 016101. Can You Go? Minimum Energy Pathways for O-2 Dissociation on
(27) Zhou, W.-P.; Yang, X.; Vukmirovic, M. B.; Koel, B. E.; Jiao, J.; Pt(111). Phys. Chem. Chem. Phys. 2012, 14, 16677−16685.
Peng, G.; Mavrikakis, M.; Adzic, R. R. Improving Electrocatalysts for (50) Wu, C.; Schmidt, D. J.; Wolverton, C.; Schneider, W. F.
O-2 Reduction by Fine-Tuning the Pt-Support Interaction: Pt Accurate Coverage-Dependence Incorporated into First-Principles
Monolayer on the Surfaces of a Pd3Fe(111) Single-Crystal Alloy. J. Kinetic Models: Catalytic No Oxidation on Pt (111). J. Catal. 2012,
Am. Chem. Soc. 2009, 131, 12755−12762. 286, 88−94.
(28) Delley, B. An All-Electron Numerical-Method for Solving the (51) Marlo, M.; Milman, V. Density-Functional Study of Bulk and
Local Density Functional for Polyatomic-Molecules. J. Chem. Phys. Surface Properties of Titanium Nitride Using Different Exchange-
1990, 92, 508−517. Correlation Functionals. Phys. Rev. B 2000, 62, 2899−2907.
(29) Delley, B. From Molecules to Solids with the Dmol(3)
Approach. J. Chem. Phys. 2000, 113, 7756−7764.
(30) Perdew, J. P.; Chevary, J. A.; Vosko, S. H.; Jackson, K. A.;
Pederson, M. R.; Singh, D. J.; Fiolhais, C. Atoms, Molecules, Solids,
and Surfaces - Applications of the Generalized Gradient Approx-
imation for Exchange and Correlation. Phys. Rev. B 1992, 46, 6671−
6687.
(31) White, J. A.; Bird, D. M. Implementation of Gradient-Corrected
Exchange-Correlation Potentials in Car-Parrinello Total-Energy
Calculations. Phys. Rev. B 1994, 50, 4954−4957.
(32) Kresse, G.; Furthmuller, J. Efficient Iterative Schemes for Ab
Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys.
Rev. B 1996, 54, 11169−11186.
(33) Zemann, J. Wyckoff Rwg - Crystal Structures. Acta Crystallogr.
1966, 21, 455.
(34) Halgren, T. A.; Lipscomb, W. N. Synchronous-Transit Method
for Determining Reaction Pathways and Locating Molecular
Transition-States. Chem. Phys. Lett. 1977, 49, 225−232.
(35) Hammer, B.; Norskov, J. K. Why Gold Is the Noblest of All the
Metals. Nature 1995, 376, 238−240.
(36) Hammer, B.; Norskov, J. K. Electronic Factors Determining the
Reactivity of Metal Surfaces. Surf. Sci. 1995, 343, 211−220.
(37) Vojvodic, A.; Norskov, J. K.; Abild-Pedersen, F. Electronic
Structure Effects in Transition Metal Surface Chemistry. Top. Catal.
2014, 57, 25−32.

15505 DOI: 10.1021/acs.jpcc.5b04511


J. Phys. Chem. C 2015, 119, 15500−15505

You might also like