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

Article

pubs.acs.org/JPCC

Controllable Synthesis of Concave Nanocubes, Right Bipyramids, and


5‑Fold Twinned Nanorods of Palladium and Their Enhanced
Electrocatalytic Performance
Zhibin Shao, Wei Zhu,* Hong Wang, Qianhui Yang, Shaolin Yang, Xiaodi Liu, and Guanzhong Wang*
Hefei National Laboratory for Physical Sciences at Microscale, and Department of Physics, University of Science and Technology of
China, Hefei, Anhui, 230026, People’s Republic of China
*
S Supporting Information

ABSTRACT: Concave palladium nanocrystals are attractive for their superior catalytic ability
arising from high densities of atomic steps and kinks. However, it is still a challenge to generate
the concave surface, which is not favored by thermodynamics owing to its higher surface energy.
In this study, concave palladium nanocubes have been synthesized kinetically in high yield via a
facile one-step wet chemical method using sodium ascorbate (NaA) as the reductant in an
aqueous solution. This process allows independent control of the average edge length and the
surface curvature of the nanocubes, respectively. The particle morphology can be tuned by
changing the reducing rate during the reaction. Right bipyramids and 5-fold twinned nanorods
with concave surfaces have also been synthesized with two reductants at the different stages or an
appropriate amount of ascorbic acid only. Remarkable enhancements in both electrocatalytic
activity and stability are observed on concave Pd nanocubes and twinned nanocrystals over
conventional Pd nanocrystals with flat surfaces and commercial Pd/C.

■ INTRODUCTION
As a superior catalyst, palladium nanocrystals have gained
namically lowered the surface energy of concave surface and
caused the formation of concave structure. They demonstrated
exceptional attention in recent years.1−7 To optimize their that the angels of the concave structure of the cubic Pd
catalytic ability, a variety of shapes have been synthesized and nanocrystals could be roughly tuned by changing the CTAC/
investigated, including cube, octahedron, tetrahedron, decahe- CTAB ratio. Despite recent progress, it remains a challenge to
dron, icosahedron, bipyramid, plate, bar, rod, and wire.8−14 fabricate Pd concave nanocrystals by a simple and rapid route
Corresponding research has demonstrated that the particle in seed-free aqueous solutions and control the degree and
catalytic activity and stability are correlated with their morphology of concavities.
shapes.15−18 Particularly, because of the presence of low- In this work, we investigated Pd concave nanocubes
coordinate atomic steps and kinks at high densities, nano- synthesized from a fast one-step wet chemical method using
particles with concave surfaces have intrinsically higher catalytic NaA as the reductant in an aqueous solution. The surface
ability than their convex counterparts.19,20 Thus, the research of curvature and morphology were modulated kinetically by the
Pd nanocrystals with concave surface has been of particular concentration of NaA in the initial stage of reaction and the
interest. However, due to their higher surface energy, concave reducing rate during the reaction, respectively. Concave right
surfaces are much harder to form than convex ones on bipyramids and 5-fold twinned nanorods of Pd were also
nanoparticles. Huang et al. reported the synthesis of Pd concave obtained by using L-ascorbic acid (AA) and NaA as reductants
tetrahedral nanocrystals with {110} and {111} using a in turn or AA only at appropriate concentration. The as-
solvothermal method in the presence of formaldehyde.21 prepared Pd concave nanocrystals exhibited superior catalytic
Also, various noble-metal concave nanocubes, including Pd,22 activity and stability in formic acid and ethanol electro-
Pt,23 Rh,24 Ag,25 and Au−Pd,26 as well as Pd and Au concave oxidation over the normal Pd nanocrystals and commercial Pd/
twinned nanoparticles,27−29 have been prepared with seed- C.


mediated methods that boost directionally controlled over-
growth. In these works, curved structures were obtained MATERIALS AND METHODS
through kinetically controlled synthesizing processes by tuning
a set of reaction parameters like the concentration of precursor Materials. Palladium nitrate dihydrate (PdNO3·2H2O,
and reductant, the injection rate, and the reaction temperature. 39.5%), (+)-sodium L-ascorbate (NaA, 99%), L-ascorbic acid
Zhang et al. reported the direct preparation of concave Pd (AA, 99.7%), and cetyltrimethylammonium bromide (CTAB,
nanocubes by the synergism of two capping reagents,
cetyltrimethylammonium chloride (CTAC) and cetyltrimethy- Received: March 12, 2013
lammonium bromide (CTAB).30 They concluded that the Revised: June 11, 2013
coadsorption of Br− and Cl− ions on the surface thermody- Published: June 11, 2013

© 2013 American Chemical Society 14289 dx.doi.org/10.1021/jp402519u | J. Phys. Chem. C 2013, 117, 14289−14294
The Journal of Physical Chemistry C Article

99%) (all from Sinopharm Chemical Reagent Co., Ltd.) were


introduced as purchased without further purification. Their
aqueous solutions were freshly prepared before use.
Synthesis of Concave Pd Nanocubes. In a typical
synthesis of concave nanocubes, CTAB (182.0 mg) and NaA
(19.8 mg) were dissolved in 15 mL of deionized water. The
solution was put in a 50 °C water bath under magnetic stirring.
When thermal equilibrium was established, 5 mL of Pd-
(NO3)2·2H2O (10.8 mg) aqueous solution was added in
rapidly. In the reaction solution, the concentration of CTAB,
NaA, and Pd(NO3)2·2H2O was 25 mM, 5 mM, and 0.8 mM,
respectively. Under magnetic stirring, the combined solution
was allowed to react for 5 min at 50 °C. Then, the solution was
centrifuged, and the product was washed three times with water
to remove excess CTAB before characterization.
Synthesis of Concave Pd 5-fold Twinned Nanorods
and Right Bipyramids. In a typical synthesis of concave 5-
fold twinned nanorods and right bipyramids, CTAB (182.0 mg)
and AA (18.0 mg) were dissolved in 15 mL of deionized water.
The solution was put in a 50 °C water bath under magnetic
stirring. When thermal equilibrium was established, 5 mL of
PdNO3·2H2O (10.8 mg) aqueous solution was added in
rapidly. In the reaction solution, the concentration of CTAB,
AA, and Pd(NO3)2·2H2O was 25 mM, 5 mM, and 0.8 mM,
respectively. After 2 min, with magnetic stirring, 1 mL aqueous
solution of NaA (19.8 mg) was added rapidly. In the reaction Figure 1. SEM (a) and TEM (b) images of Pd concave nanocubes
solution, the concentration of NaA was 5 mM. The final synthesized by reducing Pd(NO3)2·2H2O in an aqueous solution
solution reacted for 5 min at 50 °C, still under magnetic containing 5 mM NaA and 0.25 M CTAB. Models, SEM images, TEM
stirring. Then, the solution was centrifuged, and the product images, electron diffraction patterns of concave nanocubes orientated
was washed three times with water to remove excess CTAB along the [100] (c), [110] (d), and [111] (e) directions, respectively.
before characterization.
Characterization. Field-emission scanning electron micro- image (Figure 1b) present four pointed star shapes, confirming
scope (FESEM; JEOL JSM-6700F) and high-resolution the concave structure of the surfaces. The universality of
transmission electron microscope (HRTEM; JEOL model concave surfaces is further demonstrated in Figure 1, parts c, d,
2010) were used to characterize the morphology and structure and e. X-ray diffraction was conducted on product nanocubes
of the concave Pd nanocrystals. X-ray diffraction (XRD; MAC spread on a Si substrate. The pattern (Figure S1, all figures
MXPAHF) patterns and selected area electron diffraction labeled by a leading S can be found in the Supporting
(SAED) were used to determine their crystal structures. Information, SI) with fcc Pd (JCPDS file no. 87-0641), while
Electrochemical Measurements. The electrochemical the exceptionally intense (200) peak suggests that most of the
measurements were performed at room temperature with a Pd nanocubes are preferentially oriented with their {100} facets
standard three-electrode system (CH Instrument, model 620B) parallel to the substrate. Both the selected area electron
with a platinum coil counter electrode and a saturated Ag/AgCl diffraction (SAED) and high resolution TEM measurements
reference electrode. The working electrode was prepared by (Figure S2 of the SI ) on a single concave nanocube clearly
transferring 15 μL of the aqueous suspension of the prepared indicate their single-crystalline nature. While the SAED pattern
nanoparticles to a glassy carbon electrode with effective surface is indexed to the [001] zone axis of an fcc Pd single crystal, the
area of 0.196 cm2. After naturally drying in air for 2 h, the HRTEM images show the fringes with lattice spacing of 0.20
electrode was covered with 15 μL of Nafion dispersed in water nm, corresponding to the (200) lattice plane of fcc Pd.
(0.05%). With these electrodes, underpotential deposition of To understand the formation mechanism of Pd concave
Cu was performed in a solution of 0.05 M H2SO4 and 0.05 M nanocubes, the reaction parameters were systematically
CuSO4 at 5 mV/s sweep rate. Formic acid electro-oxidation investigated. With all other parameters constant with those in
cyclic voltammetry (CV) measurements were conducted in a the typical process, Figure 2 shows that NaA concentration
solution of 0.5 M H2SO4 and 0.5 M HCOOH at 50 mV/s effectively modulates the degree of concavity of the nanocubes.
sweep rate. Chronoamperometry (CA) experiments for catalyst The angle between the (100) facet and the concave surface
stability were conducted at 0.3 VAg/AgCl bias.


gradually increases with the concentration of NaA from 2.5 to
7.5 mM, while the size of nanocubes slowly decreases.
RESULTS AND DISCUSSION Homogeneous nucleation is maintained in the demonstrated
Figure 1 shows SEM and TEM images of the product NaA range. NaA is the key reductant due to its strong
synthesized using the above-mentioned typical procedure for reducibility in our process that tunes the formation of concave
Pd nanocubes. The SEM image (Figure 1a) reveals that the Pd nanocubes. As discussed by Chernov et al. and Zhang et
product can be synthesized in high purity (>95%) and relatively al.,31,32 when the side faces are capped by Br−, the reduced
large quantity. The as-prepared product has a cubic atoms prefer to nucleate and grow from the edges and corners.
morphology with the edge length of 30 ± 3 nm. Each face of When the reducing rate of metal atoms is greater than the
the cube is concave at the center. The nanocubes in the TEM surface diffusion33 rate on the particle, the newly deposited
14290 dx.doi.org/10.1021/jp402519u | J. Phys. Chem. C 2013, 117, 14289−14294
The Journal of Physical Chemistry C Article

precursor concentration is deviated from optimized value, the


shape of nanocubes tends to be irregular (Figure S5 of the SI).
Since the reduction rate of Pd precursor is a key factor for
the degree of cavity, it is interesting to investigate the effect of
changing the reduction rate in the middle of the reaction. First,
accelerating the reduction rate at an appropriate time after a
low initial rate was attempted. In the procedure of synthesizing
the Pd nanocubes with flat faces, 1 mL of 0.1 M NaA was added
in at 50 s after the reaction start time. SEM and TEM images
(Figure 3) of the product show that the edges and corners of

Figure 3. Schematic illustration of shape evolution, when accelerating


the reduction rate at 50 s after a low initial rate was performed. (a)
SEM images, (b) TEM images, and (c) HRTEM images of Pd
nanocubes that were prepared with 2.5 mM NaA at 0 s and the
additional 5 mM NaA at 50 s.

the nanocubes are more convex than those prepared by the


standard procedure. This indicates that the sudden acceleration
of the reduction rate results in preferred Pd deposition onto the
edges and corners. A transformation from the conventional
nanocubes to concave ones is achieved by accelerating the
reducing rate. Second, accelerating or decelerating the
reduction rate at an appropriate time after a high reducing
rate at the beginning was attempted. In the procedure of
synthesizing the concave nanocubes, 100 μL of 14 M HNO3 or
1 mL of 2 M NaOH was added into the reaction solution at t =
50 s. As the pH was decreased with the addition of HNO3, the
Figure 2. SEM images, TEM images, and HRTEM images of Pd
nanocubes synthesized with different NaA concentration: (a−c) 2.5,
reducing rate of Pd precursor decreased, and more Pd atoms
(d−f) 3.3, (g−i) 4.2, (j−l) 5.0, and (m−o) 7.5 mM. The insets in SEM adsorbed at edges and corners could migrate to side faces
images are 300% enlarged. (p) Summary curve of the NaA owing to surface diffusion. Figure 4, parts c and d, shows that
concentration-dependent the angle between (100) and concave the convex portion at edges and corners of cubes is more
surface and average edge length of nanocubes. obtuse than that of the typical product. In contrast, NaOH
increased pH and accelerated the reduction process. More
adsorbed Pd atoms grew at edges and corners, leading to
sharper cube edges than those under the typical and low pH
atoms will not migrate timely to the side surfaces, causing the conditions (Figure 4, parts a and b). Moreover, different from
formation of concave nanocubes. However, when the NaA adding HNO3 at 50 s, if HNO3 is introduced at 30 s, then the
concentration exceeds 7.5 mM, the degree of concavity stops concave structure on the product is either eliminated or
increasing. Instead, irregular shaped nanocubes with square replaced by some concave spots (Figure 4, parts e and f). This
convex on the surface start to appear (Figure S3 of the SI). is because reducing the growth rate at the earlier reaction stage
Alternatively, the average size of the concave nanocubes is more makes enough Pd atoms migrate to the {100} side faces and fill
sensitive to the concentration of CTAB (also a key factor in the cavity to form nearly flat faces. Like the case of adding
controlling the shape and size of Pd nanocrystals27). When HNO3, by adding a small amount of Pd(NO3)2 after the
CTAB concentrations are 50 mM and 12.5 mM, nanocubes standard procedure, the product consists mainly of larger cubes
obtained are ca. 50 and 20 nm in size, respectively (Figure S4 of (Figure S6 of the SI). It is believed that the weak reducibility
the SI). However, the concentration of Pd(NO3)2·2H2O from the residual of the reductant can no longer keep the great
precursor is not a preferred tuning parameter. When the reducing rate, and the cavity can be filled by the Pd atoms
14291 dx.doi.org/10.1021/jp402519u | J. Phys. Chem. C 2013, 117, 14289−14294
The Journal of Physical Chemistry C Article

preferential exposition of {111} planes with low surface energy,


which causes the formation of twinned nanocrystals. A dose of
NaA solution added in at 2 min after the reaction start time
resulted concave product. Shown in the SEM and TEM images
(Figure 5, parts e−h), it is identified that the product contains
24% of 5-fold twinned nanorods with starlike cross section,
42% of concave right bipyramids, and 34% of concave
nanocubes. The nanocubes are analogous with those prepared
with NaA as the reductant. Besides, many variation of triangular
and diamond shaped concave nanoparticles are observed in the
SEM and TEM images. Computer modeling (Figure S8 of the
SI) demonstrates that they are the projections of concave right
bipyramids. The high resolution TEM image (Figure S9 of the
SI) of a right bipyramid viewed along [011] zone axes
demonstrates the existence of {111} twin plane. The 5-fold
twinned nanorods are exposed as either starlike cross section
(the upper inset in Figure 5g) or star fruit like projections at
different roll angles (Figure S10 of the SI). The electron
diffraction pattern (Figure S11d of the SI) of a 5-fold nanorod
corresponds to the superposition of square ⟨100⟩ and
Figure 4. Schematic illustration of shape evolution, when accelerating rectangular ⟨112⟩ zone patterns.34 The HRTEM image (Figure
or decelerating the reduction rate at an appropriate time after a high S11c of the SI) taken over a ridge at the five-pointed star end
reducing rate at the beginning was performed. SEM images and TEM reveals the symmetric domains with a lattice spacing of 0.27
images of Pd nanocubes prepared using the typical procedure, with nm. Confirmed with the double diffraction patterns in the
additional (a, b) 1 mL of 2 M NaOH (t = 50 s), (c, d) 100 μL of 14 M SAED image (Figure S11d of the SI), it is classified that the
HNO3 (t = 50 s), and (e, f) 100 μL of 14 M HNO3 (t = 30 s) added.
nanorods have a cyclic penta-tetrahedral twin structure. Just like
The insets in SEM images are 300% enlarged.
the prior discussed growth mechanism for the concave single
crystal nanocubes, the rapid reducing rate provided by
migrating from the corners. The size fluctuation (see Figure S7 supplemental NaA leads to preferred Pd atoms deposition on
of the SI) of particles is coincident with the results caused by the edge zone of twinned seeds, which results concave twinned
atoms surface diffusion. These outcomes indicate that the cavity nanocrystals. Our investigation indicates that the formation of
morphology can be readily tuned by adjusting the reducing rate, concave structure depends on the time of adding NaA. If NaA
and even over tuning of reductant or precursor concentration was introduced at 20 s (Figure S12a of the SI), then most of the
can lead concave nanocubes back to general nanocubes. products are nanocubes along with only a few pentagonal
Using AA as the reductant, as described in the second typical
bipyramids. When the growth rate becomes too fast at the
procedure, 5-fold twinned nanorods and right bipyramids were
initiate stage, the compensation effect cannot be achieved
obtained as well as nanocubes (see Figure 5, parts a−d). As
anymore and twins are consumed, which leads to seeds
Xiong et al. reported, the reduction rate is critical to the
evolving into single crystal nanomaterials like cubes. If NaA is
formation of twinned nanostructures.10 The mild reducing
introduced at a later time point, then the growth ends and few
power of AA ensures a slow reduction, the high surface energy
Pd atoms are left to reform the edges and corners. Another
which arises from twins forming can be compensated by the
driving factor is the initial concentration of AA. The results in
Figure S12b of the SI show that the formation of Pd concave
twinned nanocrystals is maintained up to initial AA
concentration of 15 mM. However, when AA concentration
is increased to over 30 mM, 90% of the products are nanocubes
(Figure S12c of the SI).
Since the concave palladium nanocrystals have a high density
of low-coordinate atomic steps and kinks, as well as a high
surface-area-to-volume ratio, they are predicted to have
enhanced catalytic performance. The influence on the catalytic
activities was evaluated by electrocatalytic oxidation of formic
acid. The electrochemically active surface area (ECSA) of Pd
catalyst was calculated from the charges associated with the
stripping of a Cu monolayer underpotentially deposited on
their surface.35−37 As shown in Figure S13a of the SI, it is clear
that the concave nanocrystals (1.3 cm2 for concave nanocubes
and 1.1 cm2 for concave twinned-nanocrystals) have a higher
surface area than those with flat facets (0.8 cm2 for nanocubes
Figure 5. SEM images (a), TEM images (b), high-magnification
morphology images (c), and models (d) of 5-fold twinned nanorods, and 1.0 cm2 for twinned-nanocrystals). Figure 6a shows cyclic
right bipyramids, and nanocubes of Pd, synthesized with 5 mM of AA voltammograms (CVs) normalized against the ECSAs of glassy
as the reducing agent. And corresponding images (e−h) of concaved carbon electrodes modified with different Pd catalysts measured
Pd nanocrystals, synthesized initially with AA, and accelerated by NaA in a 0.5 M H2SO4 with 0.5 M HCOOH solution. The concave
at 120 s. nanocubes exhibit a maximum current density of 8.5 mA cm−2,
14292 dx.doi.org/10.1021/jp402519u | J. Phys. Chem. C 2013, 117, 14289−14294
The Journal of Physical Chemistry C Article

features. By controlling the concentration of CTAB and NaA,


the size and the curvature of concave nanocubes can be
individually controlled. By changing the reducing rate in the
reaction, the product morphology can be designed. We have
also presented a synthesis method of concave right bipyramids
and 5-fold twinned nanorods. The electrochemical activity and
stability of different shaped nanocrystals have been evaluated in
electro-oxidation of formic acid and ethanol against commercial
Pd/C. Compared with the general nanocrystals, concave
nanocrystals perform the higher electrocatalytic ability that
increases with the degree of concavity under our controlled
range.


*
ASSOCIATED CONTENT
S Supporting Information
Additional figures (Figure S1−S16) about morphology, crystal
structure, electrocatalytic performance of concave nanocrystals.
This material is available free of charge via the Internet at
http://pubs.acs.org.

■ AUTHOR INFORMATION
Corresponding Author
*Tel: +86-551-63603323; fax: +86-551-63606266; e-mail:
zhuw@ustc.edu.cn (W.Z.), gzwang@ustc.edu.cn (G.W.).
Notes
The authors declare no competing financial interest.

Figure 6. (a) Cyclic voltammograms (CV) of four different Pd


nanocrystals and commercial Pd/C measured in 0.5 M H2SO4 + 0.5 M
■ ACKNOWLEDGMENTS
This work was supported by the Natural Science Foundation of
HCOOH solution at a scan rate of 50 mVs −1. (b) Corresponding China (Grant Nos. 50772110, 50721091), the National Basic
chronoamperometry curves at 0.3 VAg/AgCl. The current values were Research Program of China (2009CB939901, 2011CB921400),
normalized to ECSA. and the Fundamental Research Funds for the Central
Universities (Grant No. WK2030000004).
2.2 times higher than the conventional nanocubes (3.8 mA
cm−2), and 9.4 times higher than the commercial Pd/C (0.9
mA cm−2). The concave twinned nanocrystals (11.0 mA cm−2)
■ REFERENCES
(1) Meng, H.; Sun, S. H.; Masse, J.-P.; Dodelet, J.-P. Electrosynthesis
show the highest maximum current density, 1.8 times higher of Pd Single-Crystal Nanothorns and Their Application in the
Oxidation of Formic Acid. Chem. Mater. 2008, 20, 6998−7002.
than the nanocrystals with flat surfaces (6.0 mA cm−2).
(2) Bai, Z. Y.; Yang, L.; Li, L.; Lv, J.; Wang, K.; Zhang, J. A Facile
Moreover, the mass activity of concave nanocrystals (see Figure Preparation of Hollow Palladium Nanosphere Catalysts for Direct
S14a of the SI) is also higher than the conventional Formic Acid Fuel Cell. J. Phys. Chem. C 2009, 113, 10568−10573.
nanocrystals and commercial Pd/C. However, the concave (3) Bianchini, C.; Shen, P. K. Palladium-Based Electrocatalysts for
nanocubes appear to have the same mass activity as that from Alcohol Oxidation in Half Cells and in Direct Alcohol Fuel Cells.
the concave twinned-nanocrystals, responsible for the higher Chem. Rev. 2009, 109, 4183−4206.
surface area. Chronoamperometry (CA) experiments at 0.3 (4) Ge, J. J.; Xing, W.; Xue, X. Z.; Liu, C. P.; Lu, T. H.; Liao, J. H.
VAg/AgCl in Figure 6b demonstrate that the electrochemical Controllable Synthesis of Pd Nanocatalysts for Direct Formic Acid
stability of concave nanocrystals is also superior to that of the Fuel Cell (DFAFC) Application: From Pd Hollow Nanospheres to Pd
conventional nanocrystals. Furthermore, the formic acid Nanoparticles. J. Phys. Chem. C 2007, 111, 17305−17310.
(5) Li, H. Q.; Sun, G. Q.; Jiang, Q.; Zhu, M. Y.; Sun, S. G.; Xin, Q.
electro-oxidation activities and the electrochemical stabilities
Synthesis of Highly Dispersed Pd/C Electro-Catalyst with High
of five batches of concave nanocubes with comparable size but Activity for Formic Acid Oxidation. Electrochem. Commun. 2007, 9,
different degrees of concavity are investigated in Figure S15 of 1410−1415.
the SI. It shows that the more concavity results higher (6) Huang, X. Q.; Tang, S. H.; Mu, X. L.; Dai, Y.; Chen, G. X.; Zhou,
enhancement of catalytic ability. Similar behavior was also Z. Y.; Ruan, F. X.; Yang, Z. L.; Zheng, N. F. Freestanding Palladium
observed with ethanol electro-oxidation in alkaline solution Nanosheets with Plasmonic and Catalytic Properties. Nat. Nanotechnol.
(Figures S16 and S13b of the SI). The concave Pd nanocrystals 2011, 6, 28−32.
demonstrate the best catalytic activity and stability. (7) Shen, X. S.; Wang, G. Z.; Hong, X.; Zhu, W. Simple-Cubic

■ CONCLUSIONS
In summary, we have presented a high-yielding, nonseeded-
Microcubes Assembled by Palladium Nanocubes. Crystengcomm 2009,
11, 753−755.
(8) Lim, B. K.; Xiong, Y. J.; Xia, Y. N. A Water-Based Synthesis of
Octahedral, Decahedral, and Icosahedral Pd Nanocrystals. Angew.
growth, one-step synthesis method of concave palladium Chem., Int. Ed. 2007, 46, 9279−9282.
nanocubes. By introducing a relatively strong reducing agent (9) Xiong, Y. J.; Cai, H. G.; Wiley, B. J.; Wang, J. G.; Kim, M. J.; Xia,
of NaA, the preferential overgrowth on both {110} and {111} Y. N. Synthesis and Mechanistic Study of Palladium Nanobars and
facets can be achieved to yield Pd nanocubes with concave Nanorods. J. Am. Chem. Soc. 2007, 129, 3665−3675.

14293 dx.doi.org/10.1021/jp402519u | J. Phys. Chem. C 2013, 117, 14289−14294


The Journal of Physical Chemistry C Article

(10) Xiong, Y. J.; Cai, H. G.; Yin, Y. D.; Xia, Y. N. Synthesis and (29) DeSantis, C. J.; Skrabalak, S. E. Core Values: Elucidating the
Characterization of Fivefold Twinned Nanorods and Right Bipyramids Role of Seed Structure in the Synthesis of Symmetrically Branched
of Palladium. Chem. Phys. Lett. 2007, 440, 273−278. Nanocrystals. J. Am. Chem. Soc. 2012, 135, 10−13.
(11) Xiong, Y. J.; McLellan, J. M.; Chen, J. Y.; Yin, Y. D.; Li, Z. Y.; (30) Zhang, J. W.; Zhang, L.; Xie, S. F.; Kuang, Q.; Han, X. G.; Xie,
Xia, Y. N. Kinetically Controlled Synthesis of Triangular and Z. X.; Zheng, L. S. Synthesis of Concave Palladium Nanocubes with
Hexagonal Nanoplates of Palladium and Their SPR/SERS Properties. High-Index Surfaces and High Electrocatalytic Activities. Chemistry −
J. Am. Chem. Soc. 2005, 127, 17118−17127. A European Journal 2011, 17, 9915−9919.
(12) Niu, W. X.; Li, Z. Y.; Shi, L. H.; Liu, X. Q.; Li, H. J.; Han, S.; (31) Chernov, A. Theory of Stability of Face Forms of Crystals. Sov.
Chen, J. A.; Xu, G. B. Seed-Mediated Growth of Nearly Monodisperse Phys. Cryst. 1972, 16, 734−753.
Palladium Nanocubes with Controllable Sizes. Cryst. Growth Des. (32) Zhang, H.; Jin, M. S.; Xia, Y. N. Noble-Metal Nanocrystals with
2008, 8, 4440−4444. Concave Surfaces: Synthesis and Applications. Angew. Chem., Int. Ed.
(13) Niu, W. X.; Zhang, L.; Xu, G. B. Shape-Controlled Synthesis of 2012, 51, 7656−7673.
(33) Hunson, J. B. Surface Science: An Introduction. Wiley, New York,
Single-Crystalline Palladium Nanocrystals. Acs Nano 2010, 4, 1987−
1998, pp 289−304.
1996.
(34) Berhault, G.; Bausach, M.; Bisson, L.; Becerra, L.; Thomazeau,
(14) Huang, X. Q.; Zheng, N. F. One-Pot, High-Yield Synthesis of 5-
C.; Uzio, D. Seed-Mediated Synthesis of Pd Nanocrystals: Factors
Fold Twinned Pd Nanowires and Nanorods. J. Am. Chem. Soc. 2009, Influencing a Kinetic- or Thermodynamic-Controlled Growth Regime.
131, 4602−4603. J. Phys. Chem. C 2007, 111, 5915−5925.
(15) Zhou, W. P.; Lewera, A.; Larsen, R.; Masel, R. I.; Bagus, P. S.; (35) Cuesta, A.; Kibler, L. A.; Kolb, D. M. A Method to Prepare
Wieckowski, A. Size Effects in Electronic and Catalytic Properties of Single Crystal Electrodes of Reactive Metals: Application to Pd(hkl). J.
Unsupported Palladium Nanoparticles in Electrooxidation of Formic Electroanal. Chem. 1999, 466, 165−168.
Acid. J. Phys. Chem. B 2006, 110, 13393−13398. (36) Chierchie, T.; Mayer, C. Voltammetric Study of the Under-
(16) Shao, M. H.; Yu, T.; Odell, J. H.; Jin, M. S.; Xia, Y. N. Structural potential Deposition of Copper on Polycrystalline and Single Crystal
Dependence of Oxygen Reduction Reaction on Palladium Nanocryst- Palladium Surfaces. Electrochim. Acta 1988, 33, 341−345.
als. Chem. Commun. 2011, 47, 6566−6568. (37) Okada, J.; Inukai, J.; Itaya, K. Underpotential and Bulk
(17) Tian, N.; Zhou, Z. Y.; Yu, N. F.; Wang, L. Y.; Sun, S. G. Direct Deposition of Copper on Pd(111) in Sulfuric Acid Solution Studied by
Electrodeposition of Tetrahexahedral Pd Nanocrystals with High- in Situ Scanning Tunneling Microscopy. Phys. Chem. Chem. Phys.
Index Facets and High Catalytic Activity for Ethanol Electrooxidation. 2001, 3, 3297−3302.
J. Am. Chem. Soc. 2010, 132, 7580−7581.
(18) Niu, Z. Q.; Peng, Q.; Gong, M.; Rong, H. P.; Li, Y. D.
Oleylamine-Mediated Shape Evolution of Palladium Nanocrystals.
Angew. Chem., Int. Ed. 2011, 50, 6315−6319.
(19) Somorjai, G. A.; Blakely, D. W. Mechanism of Catalysis of
Hydrocarbon Reactions by Platinum Surfaces. Nature 1975, 258, 580−
583.
(20) Tian, N.; Zhou, Z. Y.; Sun, S. G.; Ding, Y.; Wang, Z. L. Synthesis
of Tetrahexahedral Platinum Nanocrystals with High-Index Facets and
High Electro-Oxidation Activity. Science 2007, 316, 732−735.
(21) Huang, X. Q.; Tang, S. H.; Zhang, H. H.; Zhou, Z. Y.; Zheng, N.
F. Controlled Formation of Concave Tetrahedral/Trigonal Bipyr-
amidal Palladium Nanocrystals. J. Am. Chem. Soc. 2009, 131, 13916−
13917.
(22) Jin, M. S.; Zhang, H.; Xie, Z. X.; Xia, Y. N. Palladium Concave
Nanocubes with High-Index Facets and Their Enhanced Catalytic
Properties. Angew. Chem., Int. Ed. 2011, 50, 7850−7854.
(23) Yu, T.; Kim, D. Y.; Zhang, H.; Xia, Y. N. Platinum Concave
Nanocubes with High-Index Facets and Their Enhanced Activity for
Oxygen Reduction Reaction. Angew. Chem., Int. Ed. 2011, 50, 2773−
2777.
(24) Zhang, H.; Li, W.; Jin, M.; Zeng, J.; Yu, T.; Yang, D.; Xia, Y.
Controlling the Morphology of Rhodium Nanocrystals by Manipulat-
ing the Growth Kinetics with a Syringe Pump. Nano Lett. 2010, 11,
898−903.
(25) Xia, X. H.; Zeng, J.; McDearmon, B.; Zheng, Y. Q.; Li, Q. G.;
Xia, Y. N. Silver Nanocrystals with Concave Surfaces and Their
Optical and Surface-Enhanced Raman Scattering Properties. Angew.
Chem., Int. Ed. 2011, 50, 12542−12546.
(26) DeSantis, C. J.; Peverly, A. A.; Peters, D. G.; Skrabalak, S. E.
Octopods versus Concave Nanocrystals: Control of Morphology by
Manipulating the Kinetics of Seeded Growth via Co-Reduction. Nano
Lett. 2011, 11, 2164−2168.
(27) Shen, X. S.; Wang, G. Z.; Hong, X.; Zhu, W. Shape-Controlled
Synthesis of Palladium Nanoparticles and Their SPR/SERS Properties.
Chin. J. Chem. Phys. 2009, 22, 440−446.
(28) Wu, H. L.; Chen, C. H.; Huang, M. H. Seed-Mediated Synthesis
of Branched Gold Nanocrystals Derived from the Side Growth of
Pentagonal Bipyramids and the Formation of Gold Nanostars. Chem.
Mater. 2009, 21, 110−114.

14294 dx.doi.org/10.1021/jp402519u | J. Phys. Chem. C 2013, 117, 14289−14294

You might also like