One-Pot Synthesis of Penta-Twinned Palladium Nanowires

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Research Article

www.acsami.org

One-Pot Synthesis of Penta-twinned Palladium Nanowires and Their


Enhanced Electrocatalytic Properties
Hongwen Huang,†,‡,⊥ Aleksey Ruditskiy,§,⊥ Sang-Il Choi,† Lei Zhang,† Jingyue Liu,∥ Zhizhen Ye,‡
and Younan Xia*,†,§

The Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology and Emory University, Atlanta,
Georgia 30332, United States

State Key Laboratory of Silicon Materials and Department of Materials Science and Engineering, Zhejiang University, Hangzhou,
Zhejiang 310027, People’s Republic of China
§
School of Chemistry and Biochemistry, Georgia Institute of Technology, Atlanta, Georgia 30332, United States

Department of Physics, Arizona State University, Tempe, Arizona 85287, United States
*
S Supporting Information

ABSTRACT: This article reports the design and successful implementa-


tion of a one-pot, polyol method for the synthesis of penta-twinned Pd
nanowires with diameters below 8 nm and aspect ratios up to 100. The
key to the success of this protocol is the controlled reduction of Na2PdCl4
by diethylene glycol and ascorbic acid through the introduction of NaI and
HCl. The I− and H+ ions can slow the reduction kinetics by forming
PdI42− and inhibiting the dissociation of ascorbic acid, respectively. When
the initial reduction rate is tuned into the proper regime, Pd decahedral
seeds with a penta-twinned structure appear during nucleation. In the
presence of I− ions as a selective capping agent toward the Pd(100)
surface, the decahedral seeds can be directed to grow axially into penta-twinned nanorods and then nanowires. The Pd nanowires
are found to evolve into multiply twinned particles if the reaction time is extended beyond 1.5 h, owing to the involvement of
oxidative etching. When supported on carbon, the Pd nanowires show greatly enhanced specific electrocatalytic activities, more
than five times the value for commercial Pd/C toward formic acid oxidation and three times the value for Pt/C toward oxygen
reduction under an alkaline condition. In addition, the carbon-supported Pd nanowires exhibit greatly enhanced electrocatalytic
durability toward both reactions. Furthermore, we demonstrate that the Pd nanowires can serve as sacrificial templates for the
conformal deposition of Pt atoms to generate Pd@Pt core−sheath nanowires and then Pd−Pt nanotubes with a well-defined
surface structure.
KEYWORDS: penta-twinned nanowires, bimetallic nanotubes, oxidative etching, reduction kinetics, electrocatalytic properties

■ INTRODUCTION
Palladium nanocrystals have received considerable attention as
toward both FAO and ORR by modifying their surface
electronic structures.15,16 In addition, the use of nanowires
a primary catalyst for the electro-oxidation of formic acid can also result in substantial improvement in electrocatalytic
(FAO) and therefore the fabrication of direct formic acid fuel durability with respect to their 0D counterparts, owing to the
cells (DFAFCs).1−3 They are also emerging as an alternative to suppression of Ostwald ripping, reduction in sintering and
the Pt-based electrocatalyst toward the oxygen reduction detachment from the support, and improvement in resistance
reaction (ORR).3 Tremendous efforts have been devoted to toward metal dissolution.17−19 These merits make penta-
the synthesis of Pd nanocrystals with diversified shapes as a twinned nanowires a promising candidate for an active and
means to tailor and optimize their catalytic and electrocatalytic durable electrocatalyst.
properties through surface structure engineering.4,5 To this end, Like other metals, there is no intrinsic driving force for Pd
Pd nanocrystals with a wide variety of shapes, including atoms to grow into nanowires (or nanorods) with a highly
tetrahedra,6 cubes,7,8 octahedra,9 decahedra,10 icosahedra,10,11 anisotropic morphology due to the isotropy in interaction
right bipyramids, 12 and penta-twinned nanorods/nano- among the atoms.4,20 As a result, it has been difficult to
wires,13,14 have all been reported in recent years. Among synthesize Pd penta-twinned nanowires as uniform samples in
them, penta-twinned nanowires are particularly attractive large quantities. Nevertheless, there exist several reports in
because of their unique structural features, including the lattice
strain at a twin boundary and the highly anisotropic one- Received: August 11, 2017
dimensional (1D) morphology. It has been reported that the Accepted: August 21, 2017
lattice strain can greatly enhance the electrocatalytic activities Published: August 21, 2017

© 2017 American Chemical Society 31203 DOI: 10.1021/acsami.7b12018


ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

literature. Using a hydrothermal process at 200 °C, Zheng and Aldrich. Ethylene glycol (EG; lot no. K43B24) was purchased from J.
co-workers demonstrated the synthesis of Pd penta-twinned T. Baker. All the chemicals were used as received. Deionized (DI)
nanowires with an average diameter of 9 nm and lengths up to water with a resistivity of 18.2 MΩ·cm was used throughout the
3 μm.13 Although it was a significant advance, the authors experiments.
Synthesis of Pd Nanowires in DEG. In a standard synthesis, 105
offered essentially no discussion of the underlying mechanism. mg of PVP, 100 mg of AA, 100 mg of NaI, and 5 μL of aqueous HCl
Subsequently, Huang and co-workers also prepared Pd penta- (12 M) were dissolved in 8 mL of DEG hosted in a 24 mL vial and
twinned nanowires using a hydrothermal route.14 The authors heated at 160 °C in an oil bath for 15 min under magnetic stirring.
attributed the success of their synthesis to good control over Subsequently, 3 mL of DEG containing 30 mg of Na2PdCl4 was added
oxidative etching, caused by O 2 /halides, through the in one shot through a pipet. After the reaction had proceeded for 1 h,
introduction of small organic molecules to scavenge the halide it was terminated by immersing the vial in an ice−water bath. Finally,
ions, thus limiting their availability in the solution. Similar to the product was collected by centrifugation (12,000 rpm, 10 min),
what had been established for the synthesis of Ag nanowires,21 washed with ethanol three times, and re-dispersed in water.
the suppression of oxidative etching could protect the twinned Synthesis of Pd@Pt Core−Sheath Nanowires. In a typical
procedure, 2.0 mL of the Pd nanowires (0.20 mg/mL), 100 mg of AA,
seeds from being oxidized and dissolved during the initial stage 66.6 mg of PVP, 54 mg of KBr, and 13 mL of EG were mixed in a 50
of the synthesis. However, no mechanistic insight was mL three-neck flask and heated at 110 °C for 15 min under magnetic
presented with regard to the formation of decahedral stirring. The reaction temperature was then quickly ramped to 200 °C
seeds.4,20 Indeed, the formation of decahedral seeds represents within 30 min. Meanwhile, 1.2 mg of Na2PtCl6·6H2O was dissolved in
a critical step in the synthesis of penta-twinned nanowires. As 12 mL of EG and this solution was then injected into the flask using a
established in recent studies, the reaction kinetics play a pivotal syringe pump at a rate of 4 mL/h. After completion of the injection,
role in controlling the formation of Pd seeds with a decahedral the reaction mixture was kept at 200 °C for another 10 min. The
structure.22 It was further demonstrated that the Pd decahedral reaction was quenched by immersing the flask in an ice bath. The
seeds could be employed to produce penta-twinned nanorods products were collected by centrifugation (12,000 rpm, 10 min),
washed with ethanol three times, and re-dispersed in DI water.
made of Pd, Ag, and Cu through seed-mediated growth.23−25 In Synthesis of Pt-Based Nanotubes. In a typical process, 0.5 mL
addition to the involvement of a two-step process, the aspect of the aqueous suspension of Pd@Pt core−sheath nanowires (0.48
ratios of these nanorods were typically limited to values below mg/mL), 25 mg of FeCl3, 0.15 mL of HCl (12 M), 50 mg of PVP, 300
10. mg of KBr, and 7 mL of DI water was mixed in a 20 mL vial. The vial
In this work, we designed and successfully implemented a was then immersed in an oil bath held at 95 °C under magnetic
one-pot protocol for the facile synthesis of Pd penta-twinned stirring. After 2 h, the products were collected by centrifugation
nanowires with an average diameter as thin as 7.8 nm and (12,000 rpm, 10 min), washed with ethanol, and re-dispersed in water.
aspect ratios up to 100. The protocol was based upon the Kinetic Studies of the Pd Nanowire Synthesis. During a
reduction of a Pd(II) precursor by a combination of diethylene standard Pd nanowire synthesis, 100 μL aliquots were drawn with a
micropipette in 5 min intervals over 1 h, with the first aliquot drawn
glycol and ascorbic acid. By controlling the amounts of NaI and immediately after the injection of the Na2PdCl4 precursor. The
HCl, introduced to form PdI42− species and inhibit the aliquots were thoroughly mixed with 1.4 mL of ice-cold aqueous NaI
dissociation of ascorbic acid, respectively, the initial reduction solution (200 mg/mL). The resultant mixture was centrifuged at
rate could be tuned into the right regime for the generation of 16,500 rpm for 15 min to remove the solid product. A 100 μL aliquot
Pd decahedral seeds. The relatively slow reduction rate of the resultant supernatant was injected into 0.9 mL of aqueous NaI
associated with PdI42− was also instrumental to the induction (200 mg/mL) inside a cuvette, and the Pd(II) concentration was
and continuation of one-dimensional growth along the penta- measured using UV−vis spectroscopy.
twinned axis of a decahedral seed. Furthermore, the I− ions Electrochemical Measurements for FAO and ORR under an
Alkaline Condition. Prior to the electrochemical measurements, the
could serve as a selective capping agent toward the Pd(100) carbon-supported Pd nanowires (denoted as Pd nanowires/C) with a
surface, thereby promoting the formation of penta-twinned Pd content of 20 wt % (determined by ICP-MS) were prepared by
nanorods and then nanowires. When the reaction time was loading the as-prepared nanowires on a carbon support (Vulcan XC-
prolonged, however, the nanowires were observed to evolve 72). The catalyst ink was then prepared by ultrasonically mixing of the
into multiply twinned particles without defined geometry. Our catalyst with 960 μL of isopropyl alcohol and 40 μL of 5 wt % Nafion
analysis implies that oxidative etching was responsible for the solution for 1 h. The working electrode was obtained by drop-casting
destabilization of the penta-twinned nanowires when they were 10 μL of an aqueous suspension of the catalyst ink onto a precleaned,
suspended in a solution phase containing halide ions and glassy carbon electrode (geometrical area of 0.196 cm2, Pine
exposed to oxygen from the air at an elevated temperature. We Instruments). The loading amount of metal Pd or Pt for each catalyst
was kept at 3 μg. The electrochemical measurements were carried out
also demonstrated enhanced electrocatalytic activities and
with a three-electrode system on an IM6 electrochemical workstation
durability for the penta-twinned nanowires toward FAO and (Zahner, Germany). A Pt wire and Ag/AgCl were used as the counter
ORR under an alkaline condition. Besides, using the as- and reference electrodes, respectively. All potentials were converted to
prepared Pd nanowires as sacrificial templates, we demon- values with reference to reversible hydrogen electrode (RHE).
strated the fabrication of Pd@Pt core−sheath nanowires and To remove the surface ligand, such as I− ions, the Pd nanowires/C
Pd−Pt nanotubes. loaded on electrodes were cycled between 0.08 and 1.5 V versus RHE


(denoted VRHE) for 400 cycles in a N2-saturated aqueous solution
containing HClO4 (0.1 M) at a scan rate of 50 mV s−1. The activities
EXPERIMENTAL DETAILS of Pd nanowires/C and commercial Pd/C (10 wt % Pd loading,
Chemicals and Materials. Diethylene glycol (DEG; ≥99.0%, lot Sigma) for FAO were obtained by cycling the potential between 0.08
no. BCBK6125 V), sodium tetrachloropalladate(II) (Na2PdCl4; and 1.1 V for three cycles in a N2-saturated aqueous solution
99.998%), hexachloroplatinate(IV) hexahydrate (Na2PtCl6·6H2O; containing HClO4 (0.5 M) and HCOOH (0.5 M) at a scan rate of 50
98.0%), poly(vinylpyrrolidone) (PVP; MW ≈ 55,000), aqueous HCl mV s−1. For ORR under an alkaline condition, the cyclic voltammo-
solution (37 wt % or 12 M), iron(III) chloride (FeCl3; 97%), L- grams (CVs) for Pd nanowires/C and commercial Pt/C catalyst (20
ascorbic acid (AA; ≥99.0%), sodium iodide (NaI; ≥99.5%), and wt % Pt loading, Johnson Matthey) were measured in a N2-saturated
potassium bromide (KBr; 99.0%) were all obtained from Sigma- KOH (0.1 M) solution between 0.2 and 1.2 VRHE at a sweep rate of 50

31204 DOI: 10.1021/acsami.7b12018


ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

mV s−1. The ORR measurements for Pd nanowires/C and commercial


Pt/C were performed using the rotating disk electrode (RDE) method
in O2-saturated 0.1 M KOH at a rotation rate of 1,600 rpm and a
sweep rate of 10 mV s−1. The accelerated durability tests (ADTs) were
performed at room temperature by applying cyclic sweeps between 0.4
and 1.0 VRHE at a sweep rate of 100 mV s−1 for 10,000 cycles. The
electrochemically active surface areas (ECSAs) of the carbon-
supported Pd nanowires and commercial Pd/C catalyst were
calculated from the CVs for Cu underpotential deposition (CuUPD),
which were conducted in aqueous solutions containing H2SO4 (0.05
M) and CuSO4 (0.05 M) at a scan rate of 5 mV s−1. Consistent with
previous studies, the ECSA of commercial Pt/C in alkaline media was
estimated by integrating the hydrogen desorption regime in CV of
commercial Pt/C.26
Instrumentation. The X-ray diffraction (XRD) patterns were
obtained on a diffractometer (Paralytical XRD-600) operated at 40 kV
and 40 mA with filtered Cu Kα radiation at λ = 0.154 nm. The
transmission electron microscopy (TEM) images were obtained using
a Hitachi HT7700 microscope operated at 120 kV. The high-
resolution TEM (HRTEM) and high-angle annular dark-field scanning
TEM (HAADF-STEM) images were acquired using a JEOL JEM
2200FS STEM/TEM microscope equipped with a CEOS (Heidelberg,
Germany) probe corrector to provide a nominal image resolution of
0.07 nm. Elemental analysis of products was conducted using an
inductively coupled plasma mass spectrometer (ICP-MS, PerkinElmer,
NexION 300Q). UV−vis absorption spectra were recorded with a
Lambda 750 spectrometer (PerkinElmer).

■ RESULTS AND DISCUSSION


Characterizations of the Pd Nanowires. We first
analyzed the as-obtained nanowires using XRD. As shown in
Supporting Information Figure S1, all the peaks could be
indexed to the diffraction from fcc Pd (JCPDS No. 46-1043).
No other crystalline phases such as PdO were detected. As
shown in the TEM image (Figure 1A), the Pd nanowires had
an average length of 720 nm, together with a purity over 85%.
Figure 1. Morphological and structural characterization of the Pd
Figure 1B shows another TEM image of the Pd nanowires nanowires obtained using the standard procedure. (A, B) TEM images
recorded at a higher magnification, highlighting good at low and high magnifications, respectively, and (C) HAADF-STEM
uniformity along each nanowire, together with an average image. (D, E) HRTEM images of the region marked by a dashed box
diameter of 7.8 nm. The HAADF-STEM image in Figure 1C in panel C and the middle section of a nanowire, respectively. The red
further confirmed the good uniformity in diameter along the arrow and dotted line indicate the twin boundary running along the
long axis of an individual nanowire. A closer look at the long axis of the nanowire. The inset in panel D illustrates the
nanowire ends (Figure 1D) revealed the presence of a penta- orientation of the electron beam relative to the nanowire. (F) Fourier
twinned structure. A straight twin boundary could be seen transform pattern derived from the lattice fringes in panel E.
running along the length of the nanowire, as marked by the red
arrow. The lattice fringes only appeared on one side of the initial formation of Pd decahedral seeds, we studied the reaction
imaged nanowire, indicating that the direction of the incident kinetics underlying the nanowire synthesis. Previous work has
electron beam was parallel to one of the five side faces, as shown that the type of seed generated during the nucleation
schematically illustrated in the inset of Figure 1D. Such an phase can be controlled by tuning the initial reduction rate of a
orientation corresponds to the superposition of ⟨110⟩ and salt precursor into the proper regime.22,29 The synthesis of Pd
⟨111⟩ zones.27 The HRTEM image taken from the middle nanowires is a significantly more complicated system compared
section of a nanowire (Figure 1E) clearly shows lattice fringes to those described in the previous reports on this subject.
with spacing of 0.23 and 0.20 nm, which could be indexed to Specifically, in the nanowire synthesis, both DEG and AA
the {111} and {200} planes of fcc Pd, respectively. The Fourier served as reducing agents, with the reducing power of AA
transform (FT) pattern displayed in Figure 1F could be dependent on the concentration of HCl. However, since both
assigned to two sets of diffraction patterns corresponding to the of these components are in excess when compared to the
⟨110⟩ and ⟨111⟩ zones of Pd, further confirming the existence Pd(II) precursor, we can approximate the reaction system as
of twinned planes in the nanowire. Taken together, we can pseudo-first order. To quantify the kinetic parameters for the
conclude that the Pd nanowires possessed a penta-twinned nanowire synthesis, we used UV−vis analysis to measure the
structure bound by 10 {111} facets at the two ends and five concentration of Pd(II) precursor remaining in the solution
{100} side facets, which is a common feature of one- over the course of reaction. The experiment was conducted for
dimensional metal nanostructures composed of fcc metals both the standard synthesis and a modified procedure in which
such as Cu, Ag, and Au.27,28 the I− was substituted by a molar equivalent of Br− in order to
Effect of Reduction Kinetics on the Formation of Pd quantify the impact of a specific halide on the reaction kinetics.
Nanowires. In order to elucidate the driving forces behind the Figure S2A shows a TEM image of the Pd nanostructures
31205 DOI: 10.1021/acsami.7b12018
ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

obtained by replacing I− with Br− in the standard synthesis. The faster reduction in the bromide-mediated system. Notably, the
products consisted of a mixture of decahedra (marked by the rate constant of the iodide-mediated reaction was nearly
blue arrows), truncated right bipyramids (marked by the red identical to that of the sulfate-mediated synthesis of Pd
arrows), and nanocubes. To better distinguish the shapes of a decahedra (5.95 × 10−5 s−1), which was previously reported by
decahedron and a truncated right bipyramid, their correspond- our group.22 This result strongly suggests that Pd decahedral
ing 3D models are presented in Figure S2B,C, respectively. seeds are created in the initial stage of a nanowire synthesis, and
When the initial PdCl42− precursor is exposed to an excess the seeds subsequently evolve into pentagonal nanowires
amount of Br− or I−, rapid ligand exchange will occur, through axial overgrowth. When the k values are multiplied
generating a PdI42− or a PdBr42− complex. This change is driven by the initial concentration of PdCl42− (9.27 × 10−3 M), we
by the greater stability of the resultant complexes, with the obtain the initial reduction rates (ro): 4.47 × 10−7 and 1.62 ×
stability trending in the order of PdCl42− < PdBr42− < PdI42−.30 10−5 M s−1 for the iodide- and bromide-mediated syntheses,
In the case of standard synthesis, the characteristic absorption respectively.
peak of PdI42− at 408 nm was used to track the Pd(II) To further study the impact of the reduction kinetics, we
precursor concentration, while for the modified procedure the varied the concentrations of Na2PdCl4 and NaI. As shown in
332 nm absorption peak corresponding to PdBr42− was utilized. Figure 3A, increasing the amount of Na2PdCl4 to 45 mg
The reaction was sampled in 5 min intervals for the duration of
1 h. The concentration of the Pd(II) precursor remaining in
solution is directly proportional to the absorbance of the
characteristic peak. Using this information, the rate constant for
the reduction can be obtained by plotting the integrated form
of the pseudo-first-order reaction rate law:
ln[A]t = −kt + ln[A]0 (1)
where [A]0 and [A]t represent the concentrations of the Pd(II)
precursor at the beginning of a synthesis and at a specific time
point, respectively; k is the rate constant; and t is time. The
derived plots for both the iodide- and bromide-mediated
nanowire syntheses are shown in Figure 2. The iodide-mediated

Figure 3. TEM images of Pd nanostructures obtained using the


standard procedure with the addition of different amounts of
Na2PdCl4, (A) 45 and (B) 25 mg, respectively; and different amounts
of NaI, (C) 150 and (D) 50 mg, respectively.

Figure 2. Plot showing the natural log of Pd(II) concentration over


time for the iodide- and bromide-mediated nanowire syntheses. For produced a mixture of Pd nanocubes (with an average size of
the bromide-mediated synthesis, the characteristic peak of PdBr42− 14 nm) and right bipyramids (with an average size of 30 nm).
disappeared after 15 min due to the involvement of rapid reduction. When the amount of Na2PdCl4 was decreased to 25 mg,
The rate constant (k) and initial reduction rate (ro) are labeled on the uniform Pd nanowires were obtained (Figure 3B). Further
fitting curves. decreasing the amount of Na2PdCl4 to 15 mg resulted in no
solid product in the reaction system, due to the very slow
synthesis showed a slow, linear decline in ln[Pd2+] over the reduction rate. Variation in the amount of NaI added could also
course of 1 h, as expected from a pseudo-first-order reaction. be used to adjust the reaction kinetics. Since PdI42− is more
On the other hand, the decline in ln[Pd2+] for the bromide- stable than PdCl42−, a high concentration of NaI would
mediated reaction was comparatively rapid, with the character- appreciably slow the reduction rate.31,32 As illustrated in Figure
istic absorption peak disappearing after 15 min of reaction. This 3C,D, short Pd nanowires, as well as right bipyramids and small
result suggests a significantly more rapid reduction rate in the amounts of nanocubes, were produced if 150 mg of NaI was
case of the bromide-mediated system, which is consistent with used. In contrast, the product contained 15% of right
the relative stabilities of the PdI42− and PdBr42− complexes. The bipyramids and 85% of nanocubes when the amount of NaI
k values, derived from the slope of the fitted lines, were 4.82 × was reduced to 50 mg. All these results consistently indicate
10−5 and 1.75 × 10−3 s−1 for the iodide- and bromide-mediated that the formation of Pd nanowires requires a relatively slow
reactions, respectively. The more than 2 orders of magnitude reduction rate for the precursor. It should be noted, however,
difference between the rate constants confirms the significantly that changing the concentration of I− not only impacts the
31206 DOI: 10.1021/acsami.7b12018
ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

reduction kinetics but also influences the capping of Pd(100)


facets and the rate of oxidative etching.
On the basis of the experimental investigations described
above, we can posit a mechanism for the formation of the Pd
nanowires, summarized in Figure 4. Immediately after the

Figure 4. Schematic illustration of the nucleation and growth pathway


for the formation of Pd nanowires with a penta-twinned structure.
Homogeneous nucleation of Pd atoms initially produces decahedral
seeds. The capping of the Pd(100) facets by I− induces selective
deposition of Pd atoms on the Pd(111) facets, resulting in the
formation of pentagonal nanowires.

injection of the Na2PdCl4 precursor into the reaction solution,


the PdCl42− complex is converted to PdI42− because of ligand Figure 5. TEM images of the samples obtained at various reaction
exchange with the excess I−. The PdI42− complex then times for a standard nanowire synthesis: (A) 0.5, (B) 1.5, (C) 2, and
undergoes reduction by the combination of DEG and AA, (D) 4 h, respectively.
producing Pd(0) atoms. The reducing power of AA is
moderated through the introduction of H+ ions from HCl,
which serve to inhibit the dissociation of AA into ascorbate
monoanion, which is a much stronger reducing agent than AA. the sample. The XRD pattern (Figure S4) recorded from these
When the concentration of Pd(0) atoms reaches super- multiply twinned particles only showed peaks corresponding to
saturation, homogeneous nucleation occurs, resulting in the fcc Pd, indicating that the particles and nanowires had the same
formation of decahedral seeds. The remaining I− ions act as a composition. These observations clearly indicate that the penta-
capping agent to passivate the Pd(100) facets of the decahedra, twinned Pd nanowires were not stable in the reaction solution
as previously observed for the seeded growth of Pd nanorods.23 under ambient conditions at an elevated temperature. Zheng
Consequently, the subsequent nucleation of Pd(0) is and co-workers also reported that their Pd nanowires would
heterogeneous, occurring on the Pd(111) surfaces of the first shrink to Pd nanorods and then transform into multiply
decahedra. Additionally, the slow reduction of PdI42− limits the twinned particles as the reaction time was extended to 24 h.13
supply of Pd(0), further promoting the asymmetric overgrowth However, no detailed mechanistic discussion was provided in
of the decahedra. The result is the axial growth of the decahedra that report. The intrinsic instability of the penta-twinned Pd
along the ⟨110⟩ direction, thereby generating pentagonal nanowires is likely responsible for the limited number of
nanorods and, after further growth, nanowires. This mechanism reports on their synthesis in the literature.
is consistent with those proposed for the syntheses of Cu and To elucidate the mechanism responsible for the instability of
Ag nanowires with a pentagonal cross-section.21,27,28 Pd nanowires, we designed and conducted a set of experiments.
Oxidative Etching and Stability of the Pd Nanowires. To simplify the study, we dispersed the purified Pd nanowires
As mentioned in the Introduction, previous attempts to prepared using the standard procedure in DEG containing
synthesize Pd nanowires showed product instability at long different chemical species and then monitored their morpho-
reaction times. In order to follow the structural evolution of Pd logical changes as a function of aging time. As shown in Figure
nanowires, we characterized the products sampled at different 6A, the shape of the nanowires was retained after aging at 160
stages of a synthesis by TEM. At t = 0.5 h (Figure 5A), the °C for 4 h in the presence of PVP (15 mg/mL). When the
sample mainly contained nanowires with an average diameter of aging time was extended to 12 h, short nanorods with varying
7.5 nm and an average length of 450 nm. As the reaction time diameters and rough surfaces were observed (Figure 6B). In
was extended to t = 1.5 h, however, a small number of irregular contrast, most of the Pd nanowires were quickly broken into
particles appeared, as shown in Figure 5B. These particles small particles after aging in a DEG solution containing NaI (50
exhibited a tadpole-like geometry (with one end being mg/mL), PVP (15 mg/mL), and HCl (30 mM) for just 10 min
significantly larger than the other), together with a multiply (Figure 6C). If the aging time was increased to 60 min, Pd
twinned structure. Upon closer examination by TEM (Figure particles much larger in size would be formed, as shown in
S3), it was found that most of these particles were positioned at Figure 6D. The TEM image in the inset reveals that the as-
the ends of short Pd nanowires. As the reaction continued, the obtained particles possessed a multiply twinned structure.
number and size of such irregular particles also increased. At t = These observations indicate that the presence of NaI and HCl
2 h (Figure 5C), the tadpole-like particles grew to 200−400 nm was responsible for the instability of the penta-twinned Pd
in length while the average length of the nanowires could reach nanowires suspended in a solution phase. Similar to Cl−, I−
2 μm. At t = 4 h (Figure 5D), essentially all of the Pd nanowires contributes to the oxidative etching of Pd nanostructures by
disappeared, with only multiply twinned particles remaining in promoting the formation of a PdI42− stable complex and acting
31207 DOI: 10.1021/acsami.7b12018
ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

nanowires, causing fragmentation of the Pd nanowires into


short nanorods and small particles, as the synthesis was
continued.
On the basis of experimental observations described above,
we can propose a plausible mechanism for the time-dependent
shape evolution of Pd nanowires. Pd nanorods are formed from
initial decahedral nuclei, followed by their growth along the
⟨110⟩ direction into longer nanowires. At longer time periods,
two reactions occur in solution. The first is the reduction of
Pd2+ ions into Pd(0) atoms and their subsequent deposition
onto the {111} facets at the ends of the Pd nanowires, which
results in the longer nanowires observed in Figure 5C. The
second is the fracturing of the Pd nanowires into smaller parts,
such as nanorods, through oxidative etching. We expect that the
newly formed fracture is enclosed by high-index facets, rather
than the {111} facets found at the ends of standard nanowires.
The preferential deposition of Pd at these high-energy sites
serves to lower the total surface energy, resulting in the
formation of a roughly spherical segment at the site of fracture.
Upon further surface diffusion and Pd atom deposition the
tadpole-like particle is subsequently created. Eventually, all of
the nanowires are transformed into the irregular, twinned
particles owing to the combination of oxidative etching,
reduction of the released Pd2+ ions, and their deposition on
the surface of the irregular particle, as shown in Figure 5D.
It is well-known that the strength of oxidative etching is
strongly correlated with the reaction temperature.35 Therefore,
as the reaction temperature is decreased, the rate of oxidative
etching should be observably slowed. On the basis of this
argument, we carried out another series of experiments
conducted at 150 °C to verify the proposed mechanism for
the shape evolution of the Pd nanowires. As expected, Pd
nanowires, with average length of 350 nm and average diameter
of 8 nm, were obtained at t = 1 h, as shown in Figure S5A. After
Figure 6. TEM images of the products obtained by aging the Pd
extending the reaction time to t = 6 h, Pd nanowires and
nanowires at 160 °C. The nanowires were dispersed in different
solutions and aged under various conditions: (A, B) in DEG tadpole-like particles were produced (Figure S5B). Because of
containing PVP (15 mg/mL) for 4 and 12 h, respectively; (C, D) in the relatively slow rate of the oxidative etching at lower
DEG containing NaI (50 mg/mL), PVP (15 mg/mL), and HCl (30 temperature, when compared to standard procedure, the Pd
mM) for 10 and 60 min, respectively; (E) in DEG containing NaI (50 nanowires could be retained even after t = 6 h.
mg/mL), PVP (15 mg/mL), and HCl (30 mM) for 60 min, with Evaluation of Electrocatalytic Properties. The electro-
argon being bubbled into the solution. (F) TEM image of the product catalytic properties of the penta-twinned Pd nanowires toward
shown in panel E at a higher magnification. The dashed boxes in panel FAO and ORR in alkaline media were then evaluated. Prior to
F highlight the roughened side surface and fragmentation of the Pd the electrochemical measurements, we prepared Pd nanowires/
nanowires caused by the aging. C catalysts by loading the as-prepared samples on Vulcan XC-
72 carbon. For FAO, we benchmarked the electrocatalytic
as a charge carrier, according to the following reaction properties of the Pd nanowires/C against a commercial Pd/C
equation:10,33,34 under identical conditions. In the first step, the CVs of the Pd
nanowires/C and commercial Pd/C for the deposition and
Pd + O2 + 4I− + 4H+ → PdI4 2 − + 2H 2O (2)
stripping of a Cu monolayer underpotentially deposited (UPD)
From eq 2, it is clear that O2 could act as an oxidant and play a were collected, as shown in Figure S6. The specific ECSAs of
critical role in destabilizing the Pd nanowires. To validate the the Pd nanowires/C (26.2 m2 gPd−1) and commercial Pd/C
proposed mechanism, we bubbled argon through the DEG (44.9 m2 gPd−1) were then obtained by normalizing the ECSAs,
solution containing PVP (15 mg/mL), NaI (50 mg/mL), and which were calculated by integrating the corresponding
HCl (30 mM). Figure 6E shows a TEM image of the products stripping charges for CuUPD (420 and 460 μC cm−2 for the
obtained after aging at 160 °C for 60 min, revealing that most Pd nanowires and Pd/C, respectively), against the Pd mass.36
of the Pd nanowires were retained. Although argon was The electrocatalytic activities of the catalysts were further
bubbled into the DEG system, there was still a minor oxidative evaluated by recording CV curves in N2-saturated solutions
etching effect, as shown in Figure 6F. Some of the nanowires containing 0.5 M HClO4 and 0.5 M HCOOH at a scan rate of
were found to have roughened side faces (boxed by the large 50 mV s−1. Figure 7A shows the CVs for FAO after normalizing
dashed rectangle) as well as fractures (boxed by the small to the corresponding ECSAs of the Pd nanowires/C and Pd/C.
dashed rectangle). Considering these observations, we believe The Pd nanowires/C showed a specific current density of 13.89
that the oxidative etching likely began at the high-energy mA cm−2, more than five times the value for commercial Pd/C
twinned boundaries running along the sides of the Pd (Table S1). Such a great enhancement can be attributed to the
31208 DOI: 10.1021/acsami.7b12018
ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

Figure 7. Electrocatalytic performance of Pd nanowires/C toward FAO and ORR. (A) CVs of Pd nanowires/C and commercial Pd/C for FAO
recorded in N2-saturated aqueous solutions containing 0.5 M HClO4 and 0.5 M HCOOH at a scan rate of 50 mV s−1. The current density was
normalized to the corresponding ECSA. (B) Positive-going ORR polarization curves of the Pd nanowires/C and commercial Pt/C recorded in O2-
saturated 0.1 M KOH aqueous solutions with a sweep rate of 10 mV s−1. (C) Mass and specific activities of the carbon-supported Pd nanowires and
commercial Pt/C at 0.9 VRHE. (D) Comparison of mass activities for Pd nanowires/C and commercial Pt/C at 0.9 VRHE before (left column) and
after (right column) 10,000 cycles of ADTs. The color scheme in panel B also applies to panels C and D.

advantageous structural features of Pd nanowires, including the (Figure 7C) were further derived by normalizing the kinetic
surface strain and exposed {100} facets on the side faces.15 We current densities, which can be calculated by following the
also conducted a durability test for the Pd nanowires/C and Koutecky−Levich equation, against the corresponding metal
Pd/C in solutions containing 0.5 M HCOOH and 0.5 M mass and ECSA, respectively. Specifically, the mass and specific
HClO4 for 1,000 s with the potential fixed at 0.4 VRHE. The activities of Pd nanowires/C were 0.22 A mgPd−1 and 0.83 mA
chronoamperometric curves show that the initial and final cm−2, which were 2.0 and 3.0 times as high as those of
current densities of the Pd nanowires/C were both much commercial Pt/C catalyst (mass activity of 0.11 A mgPt−1 and
higher than those of Pd/C, indicating a great improvement in specific activity of 0.28 mA cm−2), respectively. The much
durability (Figure S7). higher activity of Pd nanowires/C is believed to originate from
For ORR under an alkaline condition, the electrocatalytic their unique structural characteristics, including, for example,
properties of Pd nanowires/C were compared with commercial surface strain, ultrathin diameter, and 1D morphology, which
Pt/C under the same measurement conditions. The CVs of Pd can optimize the surface electronic structure and thus boost the
nanowires/C and commercial Pt/C shown in Figure S8 were activity toward ORR.15,16,38
recorded at room temperature in N2-saturated 0.1 M KOH The long-term durability is another important criterion when
solutions at a sweep rate of 50 mV s−1. Because hydrogen can screening a potential catalyst for practical applications. We thus
penetrate into the Pd lattice, no obvious peak was observed in further evaluated the durability of the catalysts through an ADT
the hydrogen adsorption/desorption region for the CVs.37 at room temperature. As shown in Figure S9 and Figure 7D, the
From the CV of commercial Pt/C catalyst, the specific ECSA Pd nanowires/C only showed a drop of 13.8% in mass activity
(39.7 m2 gPt−1) was determined by normalizing the ECSA, after 10,000 cycles of ADTs, against a loss of 31.8% for the
which can be derived from the charges associated with the commercial Pt/C, indicating an excellent catalytic durability
desorption of hydrogen, against Pt mass. The electrocatalytic toward ORR in alkaline media. The TEM images in Figure S10
activities of Pd nanowires/C and commercial Pt/C were further compare the morphologies of Pd nanowires/C and commercial
measured using the RDE method. Pt/C before and after 10,000 cycles of ADTs. Compared to the
Figure 7B shows the positive-going ORR polarization curves serious sintering for commercial Pt/C, the Pd nanowires/C
of Pd nanowires/C and commercial Pt/C obtained at room well-retained their highly anisotropic morphology after the
temperature in O2-saturated 0.1 M KOH solutions at a rotation ADTs, demonstrating their superior structural stability. We
speed of 1,600 rpm. The positive potential shift for Pd argued that such a proponent structural stability was
nanowires/C implies the enhancement in catalytic activity responsible for their remarkable long-term durability.
when compared to commercial Pt/C. The mass and specific Syntheses of Pd@Pt Core−Sheath Nanowires and
activities of Pd nanowires/C and commercial Pt/C at 0.9 VRHE Pd−Pt Bimetallic Nanotubes. Pt-based one-dimensional
31209 DOI: 10.1021/acsami.7b12018
ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

nanostructures have been demonstrated as effective electro- which is the approximate thickness of the Pt shell deposited on
catalysts toward the oxygen reduction reaction (ORR).39−41 the Pd nanowires.


However, it remains a challenge to produce Pt-based one-
dimensional nanostructures. To this end, we employed the Pd CONCLUSION
nanowires, obtained using the standard procedure, as seeds to
generate the Pd−Pt bimetallic one-dimensional nanostructures. In summary, we have successfully synthesized ultrathin, penta-
The seeded overgrowth was conducted by injecting Na2PtCl6 twinned Pd nanowires, with an average diameter of 7.8 nm and
with a syringe pump into an EG-based mixture of PVP, AA, an average length of up to 720 nm using a one-pot approach.
KBr, and Pd nanowires at 200 °C (see Experimental Section for Adjusting the reduction rate of the Pd(II) precursor into the
details). As shown in Figure 8A, bimetallic nanowires with an appropriate regime, in combination with Pd(100) capping by I−
ions, was found to be crucial to the success of this synthesis.
The Pd nanowires were found to evolve into irregular, multiply
twinned particles if the reaction time was prolonged beyond 1.5
h. Further investigation revealed that this instability originated
from oxidative etching induced by the O2 from the air in
combination with I− ions. The as-prepared Pd nanowires were
also demonstrated to possess promising electrocatalytic proper-
ties toward the FAO and ORR in alkaline media. Finally, Pd@
Pt core−sheath nanowires were synthesized by using the Pd
nanowires as templates. Subsequent selective removal of Pd
from these bimetallic structures produced well-defined Pd−Pt
bimetallic nanotubes. We believe that the extensive under-
standing of the formation of Pd penta-twinned nanowires, as
well as the facile approach to the production of Pd and Pd−Pt
one-dimensional nanostructures, presented here can greatly
advance the utility of noble-metal nanocrystals.


*
ASSOCIATED CONTENT
S Supporting Information

The Supporting Information is available free of charge on the


ACS Publications website at DOI: 10.1021/acsami.7b12018.
Catalytic activities comparison; weight percentages; XRD
Figure 8. Morphological and compositional characterization of the patterns; TEM images; low- and high-magnification
Pd@Pt core−sheath nanowires, as well as the Pd−Pt nanotubes TEM images; CVs; chronoamperometric curves; and
prepared by selectively etching away the Pd cores. (A) TEM image ORR polarization curves (PDF)


and (B) EDX mapping of the Pd@Pt core−sheath nanowires. (C)
Low-magnification and (D) high-magnification TEM images of the Pt-
based nanotubes. The inset in panel D shows a HRTEM image of the AUTHOR INFORMATION
boxed region.
Corresponding Author
*E-mail: younan.xia@bme.gatech.edu.
average diameter of 10.3 nm were obtained. The thicker ORCID
diameter of the bimetallic nanowires corresponds to the Aleksey Ruditskiy: 0000-0002-1146-827X
deposition of Pt shells onto the surface of the Pd nanowires. Younan Xia: 0000-0003-2431-7048
EDX mapping shown in Figure 8B confirms the formation of
Author Contributions
Pd@Pt core−sheath nanowires. Since Pt is more resistant to ⊥
chemical oxidation when compared to Pd, we could selectively H.H. and A.R. made equal contributions to the work.
remove the Pd from the Pd@Pt core−sheath nanowires by Notes
using the Fe3+/Br− etchant pair (see Experimental Section for The authors declare no competing financial interest.


details). As shown in Figure 8C,D, hollow nanotubes were
obtained after selectively etching the Pd cores from the Pd@Pt ACKNOWLEDGMENTS
core−sheath nanowires. The composition of the nanotubes was
subsequently determined by ICP-MS. As shown in Table S2, This work was supported in part by a research grant from the
the significantly lower weight percentage of Pd in the final NSF (DMR 1506018) and startup funds from the Georgia
nanotubes, when compared to the initial nanowires, demon- Institute of Technology. As a visiting Ph.D. student from
strates the selective removal of Pd from the Pd@Pt core− Zhejiang University, H.H. was also partially supported by a
sheath nanowires. Notably, a significant amount of Pd was Fellowship from the China Scholarship Council. A.R. was
retained in the nanotubes after etching, suggesting the supported by a Graduate Research Fellowship from the NSF.
formation of a Pd−Pt alloyed shell. The inset in Figure 8D J.L. gratefully acknowledges the support by Arizona State
shows a HRTEM image captured at a section of a nanotube University and the use of facilities in the John M. Cowley
wall (marked by the dashed box in Figure 8D). The wall was Center for High Resolution Electron Microscopy at Arizona
composed of six atomic layers with a thickness of about 1.2 nm, State University.
31210 DOI: 10.1021/acsami.7b12018
ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

■ REFERENCES
(1) Rice, C.; Ha, S.; Masel, R. I.; Waszczuk, P.; Wieckowski, A.;
(20) Elechiguerra, J. L.; Reyes-Gasga, J.; Yacaman, M. J. The Role of
Twinning in Shape Evolution of Anisotropic Noble Metal Nanostruc-
tures. J. Mater. Chem. 2006, 16, 3906−3919.
Barnard, T. Direct Formic Acid Fuel Cells. J. Power Sources 2002, 111, (21) Sun, Y.; Gates, B.; Mayers, B.; Xia, Y. Polyol Synthesis of
83−89. Uniform Silver Nanowires: A Plausible Growth Mechanism and the
(2) Zhang, H.; Jin, M.; Xiong, Y.; Lim, B.; Xia, Y. Shape-Controlled Supporting Evidence. Nano Lett. 2002, 2, 165−168.
Synthesis of Pd Nanocrystals and Their Catalytic Applications. Acc. (22) Ruditskiy, A.; Zhao, M.; Gilroy, K. D.; Vara, M.; Xia, Y. Toward
Chem. Res. 2013, 46, 1783−1794. a Quantitative Understanding of the Sulfate-Mediated Synthesis of Pd
(3) Shao, M.; Chang, Q.; Dodelet, J. P.; Chenitz, R. Recent Advances Decahedral Nanocrystals with High Conversion and Morphology
in Electrocatalysts for Oxygen Reduction Reaction. Chem. Rev. 2016, Yields. Chem. Mater. 2016, 28, 8800−8806.
116, 3594−3657. (23) Huang, H.; Zhang, L.; Lv, T.; Ruditskiy, A.; Liu, J.; Ye, Z.; Xia,
(4) Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S. E. Shape-Controlled Y. Five-Fold Twinned Pd Nanorods and Their Use as Templates for
Synthesis of Metal Nanocrystals: Simple Chemistry Meets Complex the Synthesis of Bimetallic or Hollow Nanostructures. ChemNanoMat
Physics? Angew. Chem., Int. Ed. 2009, 48, 60−103. 2015, 1, 246−252.
(5) Habas, S. E.; Lee, H.; Radmilovic, V.; Somorjai, G. A.; Yang, P. (24) Luo, M.; Huang, H.; Choi, S.-I.; Zhang, C.; Silva, R. R.; Peng,
Shaping Binary Metal Nanocrystals Through Epitaxial Seeded Growth. H.-C.; Li, Z.-Y.; Liu, J.; He, Z.; Xia, Y. Facile Synthesis of Ag Nanorods
Nat. Mater. 2007, 6, 692−697. with No Plasmon Resonance Peak in the Visible Region by Using Pd
(6) Wang, Y.; Xie, S.; Liu, J.; Park, J.; Huang, C. Z.; Xia, Y. Shape- Decahedra of 16 nm in Size as Seeds. ACS Nano 2015, 9, 10523−
Controlled Synthesis of Palladium Nanocrystals: A Mechanistic 10532.
Understanding of the Evolution from Octahedrons to Tetrahedrons. (25) Luo, M.; Ruditskiy, A.; Peng, H.-C.; Tao, J.; Figueroa-Cosme,
Nano Lett. 2013, 13, 2276−2281. L.; He, Z.; Xia, Y. Penta-Twinned Copper Nanorods: Facile Synthesis
(7) Chang, G.; Oyama, M.; Hirao, K. Facile Synthesis of via Seed-mediated Growth and Their Tunable Plasmonic Properties.
Monodisperse Palladium Nanocubes and the Characteristics of Self- Adv. Funct. Mater. 2016, 26, 1209−1216.
Assembly. Acta Mater. 2007, 55, 3453−3456. (26) Holewinski, A.; Idrobo, J. C.; Linic, S. High-Performance Ag-Co
(8) Jin, M.; Liu, H.; Zhang, H.; Xie, Z.; Liu, J.; Xia, Y. Synthesis of Pd Alloy Catalysts for Electrochemical Oxygen Reduction. Nat. Chem.
Nanocrystals Enclosed by {100} Facets and With Sizes < 10 nm for 2014, 6, 828−834.
Application in CO Oxidation. Nano Res. 2011, 4, 83−91. (27) Johnson, C. J.; Dujardin, E.; Davis, S. A.; Murphy, C. J.; Mann,
(9) Niu, W.; Zhang, L.; Xu, G. Shape-Controlled Synthesis of Single- S. Growth and Form of Gold Nanorods Prepared by Seed-Mediated,
Crystalline Palladium Nanocrystals. ACS Nano 2010, 4, 1987−1996. Surfactant-Directed Synthesis. J. Mater. Chem. 2002, 12, 1765−1770.
(10) Xia, X.; Choi, S.; Herron, J. A.; Lu, N.; Scaranto, J.; Peng, H.; (28) Jin, M.; He, G.; Zhang, H.; Zeng, J.; Xie, Z.; Xia, Y. Shape-
Wang, J.; Mavrikakis, M.; Kim, M. J.; Xia, Y. Facile Synthesis of Controlled Synthesis of Copper Nanocrystals in an Aqueous Solution
Palladium Right Bipyramids and Their Use as Seeds for Overgrowth with Glucose as a Reducing Agent and Hexadecylamine as a Capping
and as Catalysts for Formic Acid Oxidation. J. Am. Chem. Soc. 2013, Agent. Angew. Chem., Int. Ed. 2011, 50, 10560−10564.
135, 15706−15709. (29) Wang, Y.; Peng, H.-C.; Liu, J.; Huang, C. Z.; Xia, Y. Use of
(11) Huang, H.; Wang, Y.; Ruditskiy, A.; Peng, H.-C.; Zhao, X.; Reduction Rate as a Quantitative Knob for Controlling the Twin
Zhang, L.; Liu, J.; Ye, Z.; Xia, Y. Polyol Syntheses of Palladium Structure and Shape of Palladium Nanocrystals. Nano Lett. 2015, 15,
Decahedra and Icosahedra as Pure Samples by Maneuvering the 1445−1450.
Reaction Kinetics with Additives. ACS Nano 2014, 8, 7041−7050. (30) Griffith, W. P.; Robinson, S. D.; Swars, K. In Gmelin Handbook
(12) Li, C.; Sato, R.; Kanehara, M.; Zeng, H.; Bando, Y.; Teranishi, T. of Inorganic Chemistry, 8th ed., Supplement Vol. B 2, Pd Compounds;
Controllable Polyol Synthesis of Uniform Palladium Icosahedra: Effect Griffith, W. P., Swars, K., Eds.; Springer-Verlag: Berlin, Heidelberg,
of Twinned Structure on Deformation of Crystalline Lattices. Angew. 1989; pp 208−210.
Chem., Int. Ed. 2009, 48, 6883−6887. (31) Achard, T.; Benet-Buchholz, J.; Escudero-Adan, E. C.; Riera, A.;
(13) Huang, X.; Zheng, N. One-Pot, High-Yield Synthesis of 5-Fold Verdaguer, X. N-Benzyl-N-phosphino-tert-butylsulfinamide and Its
Twinned Pd Nanowires and Nanorods. J. Am. Chem. Soc. 2009, 131, Coordination Modes with Ir(I), Cu(I), Pd(II), and Pt(II): P,S or P,O?
4602−4603. Organometallics 2011, 30, 3119−3130.
(14) Lu, N.; Chen, W.; Fang, G.; Chen, B.; Yang, K.; Yang, Y.; Wang, (32) Wang, C.; Wang, L.; Long, R.; Ma, L.; Wang, L.; Li, Z.; Xiong,
Z.; Huang, S.; Li, Y. 5-fold Twinned Nanowires and Single Twinned Y. Anisotropic Growth of Palladium Twinned Nanostructures
Right Bipyramids of Pd: Utilizing Small Organic Molecules To Tune Controlled by Kinetics and Their Unusual Activities in Galvanic
the Etching Degree of O2/Halides. Chem. Mater. 2014, 26, 2453− Replacement. J. Mater. Chem. 2012, 22, 8195−8198.
2459. (33) Liu, M.; Zheng, Y.; Zhang, L.; Guo, L.; Xia, Y. Transformation
(15) Choi, S.-I.; Herron, J. A.; Scaranto, J.; Huang, H.; Wang, Y.; Xia, of Pd Nanocubes into Octahedra with Controlled Sizes by
X.; Lv, T.; Park, J.; Peng, H.-C.; Mavrikakis, M.; Xia, Y. A Maneuvering the Rates of Etching and Regrowth. J. Am. Chem. Soc.
Comprehensive Study of Formic Acid Oxidation on Palladium 2013, 135, 11752−11755.
Nanocrystals with Different Types of Facets and Twin Defects. (34) Xiong, Y.; Chen, J.; Wiley, B.; Xia, Y.; Aloni, S.; Yin, Y.
ChemCatChem 2015, 7, 2077−2084. Understanding the Role of Oxidative Etching in the Polyol Synthesis
(16) Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C.; of Pd Nanoparticles with Uniform Shape and Size. J. Am. Chem. Soc.
Liu, Z.; Kaya, S.; Nordlund, D.; Ogasawara, H.; Toney, M. F.; Nilsson, 2005, 127, 7332−7333.
A. Lattice-Strain Control of the Activity in Dealloyed Core-Shell Fuel (35) Zheng, Y.; Zeng, J.; Ruditskiy, A.; Liu, M.; Xia, Y. Oxidative
Cell Catalysts. Nat. Chem. 2010, 2, 454−460. Etching and Its Role in Manipulating the Nucleation and Growth of
(17) Chen, Z.; Waje, M.; Li, W.; Yan, Y. Supportless Pt and PtPd Noble-Metal Nanocrystals. Chem. Mater. 2014, 26, 22−33.
Nanotubes as Electrocatalysts for Oxygen-Reduction Reactions. Angew. (36) Shao, M.; Odell, J.; Humbert, M.; Yu, T.; Xia, Y. Electrocatalysis
Chem., Int. Ed. 2007, 46, 4060−4063. on Shape-Controlled Palladium Nanocrystals: Oxygen Reduction
(18) Huang, H.; Li, K.; Chen, Z.; Luo, L.; Gu, Y.; Zhang, D.; Ma, C.; Reaction and Formic Acid Oxidation. J. Phys. Chem. C 2013, 117,
Si, R.; Yang, J.; Peng, Z.; Zeng, J. Achieving Remarkable Activity and 4172−4180.
Durability toward Oxygen Reduction Reaction Based on Ultrathin Rh- (37) Zhang, Z.; More, K. L.; Sun, K.; Wu, Z.; Li, W. Preparation and
Doped Pt Nanowires. J. Am. Chem. Soc. 2017, 139, 8152−8159. Charaterization of PdFe Nanoleaves as Electrocatalysts for Oxygen
(19) Ruditskiy, A.; Peng, H.-C.; Xia, Y. Shape-controlled Metal Reduction Reaction. Chem. Mater. 2011, 23, 1570−1577.
Nanocrystals for Heterogeneous catalysis. Annu. Rev. Chem. Biomol. (38) Huang, H.; Jia, H.; Liu, Z.; Gao, P.; Zhao, J.; Luo, Z.; Yang, J.;
Eng. 2016, 7, 327−348. Zeng, J. Understanding of Strain Effects in the Electrochemical

31211 DOI: 10.1021/acsami.7b12018


ACS Appl. Mater. Interfaces 2017, 9, 31203−31212
ACS Applied Materials & Interfaces Research Article

Reduction of CO2: Using Pd nanostructures as an Ideal Platform.


Angew. Chem., Int. Ed. 2017, 56, 3594−3598.
(39) Jiang, S.; Ma, Y.; Jian, G.; Tao, H.; Wang, X.; Fan, Y.; Lu, Y.; Hu,
Z.; Chen, Y. Facile Construction of Pt−Co/CNx Nanotube Electro-
catalysts and Their Application to the Oxygen Reduction Reaction.
Adv. Mater. 2009, 21, 4953−4956.
(40) Guo, S.; Dong, S.; Wang, E. Pt/Pd Bimetallic Nanotubes with
Petal-Like Surfaces for Enhanced Catalytic Activity and Stability
Towards Ethanol Electrooxidation. Energy Environ. Sci. 2010, 3, 1307−
1310.
(41) Koenigsmann, C.; Santulli, A. C.; Gong, K.; Vukmirovic, M. B.;
Zhou, W.; Sutter, E.; Wong, S. S.; Adzic, R. R. Enhanced
Electrocatalytic Performance of Processed, Ultrathin, Supported Pd−
Pt Core−Shell Nanowire Catalysts for the Oxygen Reduction
Reaction. J. Am. Chem. Soc. 2011, 133, 9783−9795.

31212 DOI: 10.1021/acsami.7b12018


ACS Appl. Mater. Interfaces 2017, 9, 31203−31212

You might also like