You are on page 1of 6

Materials Today Nano 1 (2018) 41e46

Contents lists available at ScienceDirect

Materials Today Nano


j o u r n a l h o m e p a g e : h t t p s : / / w w w . e v i s e . c o m / p r o fi l e / # / M T N A N O / l o g i n

Full Length Article

PdePt nanoalloy transformation pathways at the atomic scale*


Min Tang a, c, Beien Zhu b, c, Jun Meng b, Xun Zhang a, Wentao Yuan a, Ze Zhang a,
Yi Gao b, **, Yong Wang a, *
a
State Key Laboratory of Silicon Materials and Center of Electron Microscopy, School of Materials Science and Engineering, Zhejiang University, Hangzhou
310027 China
b
Division of Interfacial Water and Key Laboratory of Interfacial, Physics and Technology, Shanghai Institute of Applied Physics, Chinese Academy of Science,
Shanghai 201800 China

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

Article history: Nanoalloys have attracted considerable attention for their wide applications in materials, optics, catal-
Available online 14 April 2018 ysis, and biomedicine, which largely rely on their composition-, size-, and shape-dependent properties.
Because these properties change dynamically with working conditions, the knowledge of the complex
Keywords: transformation pathway of nanoalloys is highly demanded. Herein, we combined the in-situ aberration-
Transformation pathways corrected scanning transmission electron microscopy and multiscale modeling to fully resolve the whole
Palladiumeplatinum core-shell
transformation trajectory of a bimetallic nanoalloy (PdePt) at the atomic level. The transformation from
nanoparticles
core-shell to solid-solution structure is a multistep and temperature-dependent pathway, which includes
In-situ STEM
Alloying
monometallic-surface refacetting, bimetallic-surface refacetting, and alloy mixing, owing to the different
atomic activation barriers of surface diffusion and bulk migration. The critical role of shell thickness in
determining the transformation pathway was also revealed and explained. In particular, a unique partial
core-shell structure with an anisotropic surface pattern was observed in the annealing process of ul-
trathin core-shell nanoparticles. This study offers a fundamental insight into the structure evolution of
nanoalloys, which is beneficial for the development of functionalized nanoparticles with kinetic stability.
© 2018 Elsevier Ltd. All rights reserved.

Nanoalloys have attracted extensive attention for their potential experiments have shown that both structures and compositions of
applications in nanocatalysis, nano-engineering, and nanomedicine nanoalloys may considerably transform during thermal treatment
[1,2]. The unique properties of nanoalloys are determined not only and under real reaction conditions [19e26]. For instance, Vara et al.
by their morphologies and sizes, as for other nanomaterials, but recently reported that both the shape and the mixing pattern of
also by their unique chemical compositions and atomic distribu- PdePt core shell NPs change at a high temperature, consequently
tions [1e9]. For Pt-based bimetallic alloys, the solid solution changing their catalytic activities [27]. As a result, understanding
structures and core-shell structures exhibit distinct properties. The how the structure of a controlled core-shell nanoalloy changes
PdePt solid-solution alloy nanoparticles (NPs) have a higher under a given condition is crucial for the effective tuning of prop-
hydrogen-storage capacity than the core-shell-type NPs [10], and erties, which requires detailed information about the complex
their core-shell NPs have been widely employed as model catalysts transformation pathways.
for oxygen reduction reaction in fuel cells [11e15]. In particular, the In the past decade, much effort has been contributed to study
core-shell PdePt NPs with controlled shell thickness and the complex pathways of formation [28], transformation [29e32],
morphology have attracted great interests because of their distinct and dissolution [33,34] of individual nanoscale building blocks. Chi
catalytic properties [16e18]. Interestingly, recent in-situ et al. identified the distinct stages of surface element rearrange-
ments of Pt3Co NPs under thermal treatment [35]. Very recently, Ye
* et al. presented a breakthrough work for mapping the short-lived
Given his role as Editor-in-Chief, Ze Zhang had no involvement in the peer-
review of this article and has no access to information regarding its peer-review. intermediate state of single gold NPs [36] to provide the possibil-
Full responsibility for the editorial process for this article was delegated to ity of designing a kinetically stabilized surface. However, despite
Zhifeng Ren of Materials Today Physics. the above achievements, only limited information is available on
* Corresponding author.
the transformation pathways of nanoalloys because it remains a
** Corresponding author.
E-mail addresses: gaoyi@sinap.ac.cn (Y. Gao), yongwang@zju.edu.cn (Y. Wang).
challenge to monitor the structure evolution of a single NP during
c
These authors give the same contributions to this work. the long-time annealing process at the atomic scale.

https://doi.org/10.1016/j.mtnano.2018.04.003
2588-8420/© 2018 Elsevier Ltd. All rights reserved.
42 M. Tang et al. / Materials Today Nano 1 (2018) 41e46

In this study, we employed in situ scanning transmission elec- NP, which reveals the continuous lattice fringes from the Pd core to
tron microscopy (STEM) annealing experiments to fully monitor the Pt shell, indicating an epitaxial relationship between them. As
the whole pathway of the transformation of PdePt NPs from shown in Fig. 1(b), the interplanar spacing of 1.945 and 1.415 Å on
segregated core-shell structures to randomly mixed structures. The the edge and corner of the cubic NP can be indexed to FCC Pt {100}
dynamic evolution of morphology, elemental diffusion, and alloy and {110} planes, respectively. No misfit dislocation was observed
mixing for individual NPs during annealing was tracked as a because the lattice mismatch between Pt and Pd is 0.8% only. The
function of temperature and time, by high-angle annular dark-field distribution of chemical components in the bimetallic PdePt cubic
(HAADF) STEM imaging. The temperature-dependent trans- NPs at room temperature was assessed using STEM/EDX mapping,
formation pathways were identified and rationalized by the density as shown in Fig. 1(c and d), respectively. As there is a relatively large
functional theory calculations and multiscale modeling. This study difference in the atomic number of 78Pt and 46Pd, the contrast
facilitates the tailoring of the physical and chemical properties of present in the atomic-scale HAADF-STEM images can be used to
nanoalloys for catalytic applications. analyze the elemental distribution of Pt and Pd [39,40]. Further
The core-shell PdePt cubic NPs were prepared by a chemical quantification of the distribution is very challenging for large-sized
method reported in a previous work [37] (see Supplemental nanoalloy particles. It is out of the scope of this work and needs
Material for details). The highly monodispersed PdePt NPs (to further studies in the future [41,42].
prevent the nanoparticle coalescence) were heated on a silicon Fig. 2(a) shows the STEM image of a PdePt cube kept at 300  C
nitride film using a double-tilt DENS TEM holder [38]. Because the for 140 min. No obvious structural change was observed. With the
core-shell structures are not the equilibrium states of NPs, the increase in temperature to 400  C, a refacetting of the core-shell NP
structures will transform during the annealing process. The in situ without alloy mixing was observed. As shown in Fig. 2(b, d), the
annealing experiments were performed in an aberration-corrected morphology of the monitored NP became more and more round
STEM (Titan ChemiSTEM, FEI). During the long-time heating pro- with time. The ultra-high resolution images (see the second and
cess, special attention has been paid to obtain atomic-resolution third rows of Fig. 2) clearly show that the surface refacetting
HAADF-STEM images as NPs rotate frequently at a high tempera- resulted from the changes in the Pt shell thickness on the {110} and
ture. Besides, EDX mapping has also been performed at the same {100} surfaces. In the beginning, the thicknesses of the atomic
temperature to acquire in-situ information of the elemental dis- layers of Pt shell were the same on the {110} and {100} facets. The Pt
tribution of PdePt alloys. shell thickness on the {100} facet increased from 4 atomic layers at
The steps in the annealing process are shown in Fig. 1(a), where 300  C to 6 layers, 9 layers, 11 layers after 30, 35, and 55 min at
the temperature was increased to 600  C over 400 min. Fig. 1(b) 400  C, respectively. Furthermore, the Pt shell thickness on the
shows the atomic HAADF-STEM image of a corner of a typical PdePt {110} facet decreased from 4 atomic layers to 0 atomic layers. The

Fig. 1. (a) Annealing sequence for in situ experiments, plotted as the thermal annealing temperature with respect to time. The specific annealing conditions for the HAADF-STEM
images shown in Fig. 2 are labeled with red stars in this plot. (b) Atomic-resolution HAADF-STEM image of a corner of a typical PdePt NP at room temperature. (ced) HAADF-STEM
image and corresponding EDX elemental map of a PdePt NP at RT.
M. Tang et al. / Materials Today Nano 1 (2018) 41e46 43

Fig. 2. (aef) Atomic-resolution HAADF-STEM images of a single PdePt NP acquired at different temperatures and annealing times. (gel; mer) Corresponding details of {110} and
{100} facets, respectively. It should be noted that the electron beam may induce some contamination (the bright area in the center of the particle) when we focus the beam on the
particle to adjust the zone axis to obtain high-resolution HAADF images (for more details, please refer Figure S5).

STEM images and the EDX elemental mapping in Fig. 3(aec) show image and the Pd bulk. When the sample temperature was
that during refacetting, there was no obvious migration of the Pt increased to 500  C and held for 75 min, the migration of Pt atoms
atoms into the bulk, which indicates that the surface refacetting into the core of the NP was observed as evidenced by the EDX
results from the surface diffusion of Pt atoms. Note that the Pt mapping (Fig. 3(def)), where more Pt atoms are observed in the
atoms seen in the middle of the particle in Fig. 3(c) resulted from middle of the particle, compared with those in Fig. 3(c). Interest-
the overlapping between the Pt(100) surface parallel to the STEM ingly, on the {100} facets, Pt shell with 11 atomic layers still re-
mains, while the area of the Pd {110} facets increases. This further
refacetting makes the morphology of the nanoalloy more round
(shown by yellow arrows in Fig. 2(e, q)). When the temperature was
increased to 600  C, the transformation of the monitored NP from a
core-shell structure to a random mixed structure was completed.
The EDX mapping in Fig. 3(gei) shows that Pt and Pd atoms were
homogeneously mixed. To exclude the effect of beam irradiation,
the morphology of other NPs (only exposed to electron beam when
taking images) at a high temperature was also acquired (refer to
Figures S1 and S3, Supplemental Material), showing a similar
phenomenon as observed in Fig. 2, which implies that beam irra-
diation is not a dominant factor in the refacetting and alloying
process.
To understand the experimental results, the refacetting process
observed at 400  C is interpreted first. We divide the refacetting
process into two stages. In the first stage, the Pt shell becomes
thinner on the {110} facets, while it becomes thicker on the {100}
facets, indicating the directional movement of Pt atoms from the
{110} facets to {100} facets. The Pt shell remains complete until a
monolayer Pt is present on the {110} facets (Fig. 2(aec)). Thus, this
stage is defined as the monometallic-surface refacetting stage. It can
be understood that by chamfering the cube, the surface energy re-
duces. The initial cubic structure is obviously much different from
its equilibrium structure, which is predicted to be a truncated oc-
tahedron composed mainly of {111} and {100} facets according to
the well-known Wulff rule [43e45]. Because the STEM image can
offer only 2-D structure information but not the {111} facet fraction,
Fig. 3. HAADF-STEM images and corresponding EDX elemental maps of a PdePt NP, it would be difficult to directly compare the refacetted structures
which was annealed at different temperatures: 400  C (aec), 500  C (def), and 600  C with the Wulff constructed structure. However, we can still find the
(gei).
44 M. Tang et al. / Materials Today Nano 1 (2018) 41e46

reason for refacetting by measuring the ratio of d100 to d110 (R100/110), PtePt (PdePd) first-neighbor bonding energy (εptpt : 0.42 eV,
where d100 and d110 are the distances from the center of the particle εpdpd : 0.34 eV) was fitted using the surface energies calculated at
to the {100} and {110} facets, respectively. When the NP reaches its DFT level (Table S1). Moreover, the PdePt bonding energy
equilibrium structure, the R100/110 should be equal to the ratio of (εptpd : 0.38 eV) was fitted using the surface segregation energy of
g100 to g110 (g the surface tension) as stated by George Wulff. The Pt on the Pd surface. A series of model systems were tested using
measured R100/110 from Fig. 2(aec) is 0.712, 0.793, and 0.839 both the DFT calculations and fitting parameters. The difference
respectively, while the ideal R100/110 calculated by the surface ten- between the DFT results and the fitted ones was reasonably small
sion of Pt (100) and (110) surfaces is 0.917. Thus, it can be concluded (Table. S2). All details are provided in the Supplemental Material.
that the NP tends to reach its equilibrium shape by directional The change in the energy caused by the KMC trial is considered by
surface atom diffusion. Interestingly, the second refacetting stage counting the number of broken and formed nearest-neighbor
named as the bimetallic-surface refacetting stage is found in our bonds. For example, if a Pt atom jumps from position 1 to posi-
experiments. At the end of the first stage, the refacetting process is tion 2, then
not finished because the R100/110 is considerably different from the
ideal one. However, because there is only a monolayer Pt on the    
{110} facets, further chamfering exposes the Pd core on the {110} DE ¼ εPtPt CNPt
2 2
þ εPdPt CNPd 1
 εPtPt CNPt 1
þ εPdPt CNPd (1)
facets. Because Pd has a lower surface tension than Pt [44,45], this
exposure of Pd core triggers an expansion of Pd atoms on the {110} where CNPt 1 /CN 1 and CN 2 /CN 2 are the numbers of first-neighbor
Pd Pt Pd
facets, which squeezes the Pt {100} facets, resulting in a more round Pt/Pd atoms before and after the movement. A Metropolis-
shape. Hastings algorithm was then carried out on the basis of the
As observed from our experiments, the whole transformation energy barrier and the energy balance. If the trial is accepted
process of the cubic PdePt core-shell nanoalloy involves two indi- eventually, we refer it as an effective KMC step.
vidual processes, i.e. the surface refacetting process and the alloying The simulated NP had a size of about 9 nm, containing 45,563
process. Two energy barriers play important roles in these two atoms and 2 Pt layers. A 60-million-KMC-step simulation at 400  C
processes, i.e. the diffusion energy barriers (DED) of the Pt atoms on was performed first to simulate the surface refacetting process. It
the {110} and {100} surfaces and the migration energy barriers can be seen from Fig. 4(b and c) and Supplemental Material that our
(DEM) of the Pt atom from Pt subsurface layer to the Pd core. DFT simulation well reproduces the experimental observations: (1) The
calculations were performed to calculate the energy barriers by Pt shell on {100} facets became thicker with time. (2) The direc-
using the Vienna ab initio simulation package (VASP) [46]. The GGA- tional diffusion of Pt atoms from {110} facets to {100} facets
PBE functional and the projector augmented-wave methods were exposed the Pd atoms on {110} facets. Besides, the {111} facets were
adopted [47,48]. The transition state during the diffusion or found at the truncated vertexes. Further statistical analysis showed
migration was found by the nudged elastic band method imple- that the number of atoms on {110} facets increased with KMC steps,
mented in VASP. More details can be seen in the Supplemental thus representing the expansion of {110} facets. Furthermore, the
Material. The results showed that DED is 0.86 eV, while DEM is concentration of Pd atoms on the exposed {110} facets increased
approximately 0.94 eV. The corresponding energy pathways can be from 9% at 10 M KMC steps to 76% at 60 M KMC steps. It can also be
found in Figure S4 in the Supplemental Material. At 400  C, DED can observed that the partial core-shell structure showed some sta-
be overcome, while it is difficult to overcome DEM. Therefore, we bility after 50 M KMC steps, and no large change was found in the
consider that the refacetting process occurs without alloy mixing. structure and composition. The annealing simulation was per-
When the temperature is relatively higher (500  C and 600  C), both formed by 50 M KMC steps at each temperature (400  C, 500  C,
the energy barriers can be overcome. Thus, it can be concluded that and 600  C). As expected, the alloying process began at 500  C
the alloying began at 500  C (Fig. 3(def)) and finished completely at (Figure S9). The final structure at 600  C was a random mixed round
600  C (Fig. 3(gei)). It can be predicted that if a high temperature structure (Fig. 4(d)). Using the simulation results, the whole
treatment is applied without an annealing process, the trans- multistep PdePt core-shell alloy transformation is interpreted in
formation process could thus be a one-step process. This was Fig. 4.
confirmed by an experiment where the NPs were heated directly to As discussed above, the goal of the refacetting process is to
600  C, where we observed the surface refacetting and alloying change the controlled shape to the equilibrium shape. If the shell
simultaneously (Figures S5 and S6). thickness of the NP is relatively larger than its size, it may reach the
The above analysis indicates that (1) the difference between the equilibrium structure without the bimetallic-surface refacetting
surface diffusion barrier and the bulk migration barrier makes the process. To confirm this idea, we first ran the developed KMC
transformation process a selective multistep process. (2) The sur- simulation using a NP of the same size but with 4 Pt layers. The
face energy plays an important role in the surface refacetting pro- results showed that the unique partial core-shell structure did not
cess. To confirm our understanding, an effective kinetic Monte appear for this case (Fig. 5, Figure S10 in the Supplemental
Carlo (KMC) simulation model is established to reproduce the Material). Followed by the theoretical simulation results, we per-
experimental observations. In each KMC trial, a random atom i is formed similar experiments using NPs with thicker Pt shells.
picked to move it to its nearby empty crystal lattice site. The Consistent results were obtained: in the surface refacetting process,
acceptance rate of this trial is controlled by both the energy barrier the shell was too thick to expose the subsurface Pd atoms. It is
and the energy balance of the movement. The possibility of over- known that the catalytic property of the core-shell NP is very
coming the diffusion/migration energy barrier is determined first. sensitive to its shell thickness. Besides the known effect of shell
DED is used if it is surface diffusion, while DEM is used if it is bulk thickness on the electronic structure, we show here that the shell
migration. By doing this, the surface diffusion and the bulk thickness could also change the intermediates appearing in the
migration are selected. If this overcomes the energy barrier, then transformations of a core-shell structure. Some unique structures
the energy change for moving the atom is further determined. To may appear in the case of an ultrathin core-shell structure. For
reduce the computational cost to make the simulation of large- example, the partial core-shell structure found in our study may
sized NPs feasible, we adopted a simplified model to describe the have a special catalytic property for some kinds of reactions
change in energy. The model assumes that the contributions to the because NPs have selective Pt and Pd facets on the surface
surface energy are from the missing first-neighbor bonds [49]. The simultaneously.
M. Tang et al. / Materials Today Nano 1 (2018) 41e46 45

Fig. 4. Selective mechanism of the transformation from core-shell structure to mixed alloys. The 3-D structures were obtained from the kinetic Monte Carlo simulations. Orange
balls: Pt atoms; blue balls: Pd atoms.

In summary, by monitoring the transformation pathway of a monometallic-surface refacetting, bimetallic-surface refacetting,


single PdePt NP from a cubic core-shell structure to a round alloyed and alloy mixing. Depending on the heating treatments, these
structure, we revealed a selective mechanism for alloy trans- stages may occur in a stepwise manner or simultaneously. By the
formation. Three main stages may occur during the transformation: theoretical analysis and kinetic Monte Carlo simulations, we

Fig. 5. (1e6) KMC simulated results of similar transformation of the 9-nm PdePt NP with 4 Pt shells, 50 M effective KMC steps at each temperature. Orange balls: Pt atoms, blue
balls: Pd atoms. (aef) Experimentally observed alloy transformation of the PdePt NP with thicker shells.
46 M. Tang et al. / Materials Today Nano 1 (2018) 41e46

further explained that the refacetting process occurs with the [14] S. Xie, S.I. Choi, N. Lu, L.T. Roling, J.A. Herron, L. Zhang, J. Park, J. Wang,
M.J. Kim, Z. Xie, M. Mavrikakis, Y. Xia, Nano Lett. 14 (2014) 3570e3576.
purpose of transforming a NP from a controlled structure to its
[15] J. Zhang, M.B. Vukmirovic, Y. Xu, M. Mavrikakis, R.R. Adzic, Angew. Chem. Int.
equilibrium structure. We also showed that the shell thickness Ed. 44 (2005) 2132e2135.
plays an important role in deciding the transformation pathway. [16] B. Lim, M. Jiang, P.H. Camargo, E.C. Cho, J. Tao, X. Lu, Y. Zhu, Y. Xia, Science 324
Several intermediates with anisotropic shell thickness were char- (5932) (2009) 1302e1305.
[17] P. Strasser, S. Koh, T. Anniyev, J. Greeley, K. More, C. Yu, Z. Liu, Nat. Chem. 2
acterized during the refacetting process in situ observations. In (2010) 454e460.
particular, a unique partial core-shell structure with an anisotropic [18] D. Wang, H.L. Xin, R. Hovden, H. Wang, Y. Yu, D. Muller, F. DiSalvo, H. Abruna,
surface pattern was observed with the ultrathin layer core-shell NP. Nat. Mater. 12 (2013) 81e87.
[19] F. Tao, M.E. Grass, Y. Zhang, D.R. Butcher, J.R. Renzas, Z. Liu, J.Y. Chung,
This work helps to understand and control the alloy transformation B.S. Mun, M. Salmeron, G.A. Somorjai, Science 322 (5903) (2008) 932e934.
pathway and reveals the potential usage of the intermediates in [20] C. Cui, L. Gan, M. Heggen, S. Rudi, P. Strasser, Nat. Mater. 12 (2013) 765e771.
future studies. [21] H. Zheng, J. Wang, J.Y. Huang, J. Wang, Z. Zhang, S.X. Mao, Nano Lett. 13 (12)
(2013) 6023e6027.
[22] Y. Jiang, H. Li, Z. Wu, W. Ye, H. Zhang, Y. Wang, C. Sun, Z. Zhang, Angew. Chem.
Acknowledgments Int. Ed. 55 (2016) 12427e12430.
[23] J. Villanova, R. Daudin, P. Lhuissier, D. Jauffres, S. Lou, C.L. Martin, S. Laboure,
R. Tucoulou, G.M. Criado, L. Salvo, Mater. Today 20 (2017) 354e359.
We acknowledge the support of the National Nature Science [24] Y.A. Wu, L. Li, Z. Li, A. Kinaci, M.K. Chan, Y. Sun, J.R. Guest, I. McNulty, T. Rajh,
Foundation of China (51390474, 11574340, 11604357, 91645103, Y. Liu, ACS Nano 10 (2016) 3738e3746.
11234011, 11327901, and 21773287). B.Z. thanks the Natural Science [25] S. Prabhudev, M. Bugnet, G.Z. Zhu, C. Bock, G.A. Botton, ChemCatChem 7
(2015) 3655e3664.
Foundation of Shanghai (16ZR1443200). The authors thank Prof.
[26] J. Sun, L. He, Y.C. Lo, T. Xu, H. Bi, L. Sun, Z. Zhang, S.X. Mao, J. Li, Nat. Mater. 13
Shengbai Zhang and Prof. Riccardo Ferrando for helpful discussions. (2014) 1007e1012.
The computational resources utilized in this research were pro- [27] M. Vara, L.T. Roling, X. Wang, A.O. Elnabawy, Z.D. Hood, M. Chi, M. Mavrikakis,
vided by Shanghai Supercomputer Center, National Super- Y. Xia, ACS Nano 11 (2017) 4571e4581.
[28] H. Zheng, R.K. Smith, Y.W. Jun, C. Kisielowski, U. Dahmen, A.P. Alivisatos,
computing Center in Tianjin and Shenzhen, and Special Program for Science 324 (5932) (2009) 1309e1312.
Applied Research on Super Computation of the NSFC-Guangdong [29] S.H. Tolbert, A.P. Alivisatos, Science 265 (5170) (1994) 373e376.
Joint Fund (the second phase) under Grant No. U1501501. [30] P. Buffat, J.P. Borel, Phys. Rev. A 13 (1976) 2287e2298.
[31] D.H. Son, S.M. Hughes, Y. Yin, A. Paul Alivisatos, Science 306 (5698) (2004)
1009e1012.
Appendix A. Supplementary data [32] Y. Sun, Y. Xia, Science 298 (5601) (2002) 2176e2179.
[33] M.J. Mulvihill, X.Y. Ling, J. Henzie, P. Yang, J. Am. Chem. Soc. 132 (2010)
268e274.
Supplementary data related to this article can be found at [34] M.N. O'Brien, M.R. Jones, K.A. Brown, C.A. Mirkin, J. Am. Chem. Soc. 136 (2014)
https://doi.org/10.1016/j.mtnano.2018.04.003. 7603e7606.
[35] M. Chi, C. Wang, Y. Lei, G. Wang, D. Li, K.L. More, A. Lupini, L.F. Allard,
N.M. Markovic, V.R. Stamenkovic, Nat. Commun. 6 (2015) 8925.
References [36] X. Ye, M.R. Jones, L.B. Frechette, Q. Chen, A.S. Powers, P. Ercius, G. Dunn,
G.M. Rotskoff, S.C. Nguyen, V.P. Adiga, A. Zettl, E. Rabani, P.L. Geissler,
[1] R. Ferrando, J. Jellinek, R.L. Johnston, Chem. Rev. 108 (3) (2008) 845e910. A.P. Alivisatos, Science 354 (6314) (2016) 874e877.
[2] D.S. Su, B. Zhang, R. Schlogl, Chem. Rev. 115 (2015) 2818e2882. [37] N. Lu, J. Wang, S. Xie, J. Brink, K. McIlwrath, Y. Xia, M.J. Kim, J. Phys. Chem. C
[3] B.T. Sneed, C.N. Brodsky, C.H. Kuo, L.K. Lamontagne, Y. Jiang, Y. Wang, F.F. Tao, 118 (2014) 28876e28882.
W. Huang, C.K. Tsung, J. Am. Chem. Soc. 135 (2013) 14691e14700. [38] Y. Jiang, Y. Wang, Y.Y. Zhang, Z. Zhang, W. Yuan, C. Sun, X. Wei, C.N. Brodsky,
[4] L. Li, L. Zhou, S.O. Chikn, D.H. Anjum, M.B. Kanoun, J. Scaranto, M.N. Hedhili, C.-K. Tsung, J. Li, X. Zhang, S.X. Mao, S. Zhang, Z. Zhang, Nano Res. 7 (2014)
S. Khalid, P.V. Laveille, L.D. Souza, A. Clo, J.M. Basset, ChemCatChem 7 (2015) 308e314.
819e829. [39] Y. Jiang, Y. Wang, J. Sagendorf, D. West, X. Kou, X. Wei, L. He, K.L. Wang,
[5] B.T. Sneed, A.P. Young, D. Jalalpoor, M.C. Golden, S. Mao, Y. Jiang, Y. Wang, S. Zhang, Z. Zhang, Nano Lett. 13 (2013) 2851e2856.
C.K. Tsung, ACS Nano 8 (7) (2014) 7239e7250. [40] S.I. Sanchez, M.W. Small, J. Zuo, R.G. Nuzzo, J. Am. Chem. Soc. 131 (2009)
[6] Y. Yang, L. Lai, Y. Guo, Z. Dai, R. Zhang, C. Sun, X. Zhou, J. Electroanal. Chem. 8683e8689.
783 (2016) 132e139. [41] A. Backer, G.T. Martinez, A. Rosenauer, S.V. Aert, Ultramicroscopy 134 (2013)
[7] S.K. Konda, A. Chen, Mater. Today 19 (2016) 100e108. 23e33.
[8] S.B. Vendelbo, C.F. Elkjær, H. Falsig, I. Puspitasari, P. Dona, L. Mele, B. Morana, [42] K.H.W. van den Bos, A. Backer, G. Martinez, N. Winckelmans, S. Bals, P. Nellist,
B.J. Nelissen, R. Rijn, J.F. Creemer, P.J. Kooyman, S. Helveg, Nat. Mater. 13 S.V. Aert, Phys. Rev. Lett. 116 (2016) 246101.
(2014) 884e890. [43] G.Z. Wulff, Z. Kristallogr, Cryst. Mater 34 (1901) 449.
[9] Y. Yue, D. Yuchi, P. Guan, J. Xu, L. Guo, J. Liu, Nat. Commun. 7 (2016) 12251. [44] B. Zhu, J. Meng, Y. Gao, J. Phys. Chem. C 121 (2017) 5629e5634.
[10] H. Kobayashi, M. Yamauchi, H. Kitagawa, Y. Kubota, K. Kato, M. Takata, J. Am. [45] B. Zhu, Z. Xu, C. Wang, Y. Gao, Nano Lett. 16 (2016) 2628e2632.
Chem. Soc. 132 (2010) 5576e5577. [46] G. Kresse, J. Hafner, Phys. Rev. B 47 (1) (1993) 558e561.
[11] J. Park, L. Zhang, S.I. Choi, L.T. Roling, N. Lu, J.A. Herron, S. Xie, J. Wang, [47] P.E. Blochl, O. Jepsen, O.K. Andersen, Phys. Rev. B 49 (23) (1994)
M.J. Kim, M. Mavrikakis, Y. Xia, ACS Nano 9 (3) (2015) 2635e2647. 16223e16233.
[12] K. Sasaki, H. Naohara, Y. Cai, Y.M. Choi, P. Liu, M.B. Vukmirovic, J.X. Wang, [48] G. Kresse, D. Joubert, Phys. Rev. B 59 (3) (1999) 1758e1775.
R.R. Adzic, Angew. Chem. Int. Ed. 49 (2010) 8602e8607. [49] W.F. Hosford, Materials Science: an Intermediate Text, Cambridge University
[13] J.X. Wang, H. Inada, L. Wu, Y. Zhu, Y. Choi, P. Liu, W.P. Zhou, R.R. Adzic, J. Am. Press, 2006.
Chem. Soc. 131 (2009) 17298e17302.

You might also like