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

Letter

pubs.acs.org/NanoLett

Use of Reduction Rate as a Quantitative Knob for Controlling the


Twin Structure and Shape of Palladium Nanocrystals
Yi Wang,†,‡,⊥ Hsin-Chieh Peng,§ Jingyue Liu,∥ Cheng Zhi Huang,‡ and Younan Xia*,†,§

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

Key Laboratory of Luminescent and Real-Time Analytical Chemistry, Ministry of Education, School of Chemistry and Chemical
Engineering, Southwest University, Chongqing 400715, 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: Kinetic control is a powerful means for maneuvering


the twin structure and shape of metal nanocrystals and thus optimizing
their performance in a variety of applications. However, there is only a
vague understanding of the explicit roles played by reaction kinetics
due to the lack of quantitative information about the kinetic
parameters. With Pd as an example, here we demonstrate that kinetic
parameters, including rate constant and activation energy, can be
derived from spectroscopic measurements and then used to calculate
the initial reduction rate and further have this parameter quantitatively
correlated with the twin structure of a seed and nanocrystal. On a
quantitative basis, we were able to determine the ranges of initial
reduction rates required for the formation of nanocrystals with a
specific twin structure, including single-crystal, multiply twinned, and
stacking fault-lined. This work represents a major step forward toward
the deterministic syntheses of colloidal noble-metal nanocrystals with specific twin structures and shapes.
KEYWORDS: Nanocrystals, palladium, kinetic control, twin structure

T here is a steadily growing interest in noble-metal


nanocrystals with controlled shapes (and thus facets)
due to their unique applications in catalysis, energy harvesting/
be decahedra (or icosahedra) and thin plates, respectively. This
correlation between the shape of a nanocrystal and the twin
structure of its seed was initially formulated for Ag nanocrystals
conversion/storage, electronics, display, environmental protec- prepared using the polyol method with ethylene glycol (EG)
tion, sensing, and biomedical research.1−10 Thanks to the serving as a solvent and a precursor to the reductant for AgNO3
efforts from many groups, significant progress has been made in at an elevated temperature.14 Later, it was also observed in
understanding the formation of nanocrystals with a specific other metals including Au, Cu, Pd, Pt, and Rh, as well as their
shape by identifying the nucleation and growth pathway. In bimetallic combinations.11,15−17
general, nuclei with fluctuating structures will be formed via Despite the incredible progress, it remains a grand challenge
homogeneous nucleation when metal atoms are generated in a to rationally and reproducibly manipulate the shape of noble-
solution through the reduction or decomposition of a metal nanocrystals and thus control their properties. The
precursor. Once the nuclei have grown past a critical size, challenge can be largely attributed to the lack of a quantitative
they evolve into seeds with different twin structures, including knob that can be employed to precisely control the nucleation
those with a single-crystal, singly twinned, multiply twinned, or and growth of nanocrystals. For example, there is still no
stacking-fault-lined structure.11 The seeds will then grow into quantitative knowledge about the experimental conditions or
nanocrystals with distinctive shapes or facets as dictated by parameters that control the twin structure taken by a seed
interplay of surface capping and reaction kinetics. For example, during the nucleation process. In many cases, one has to rely on
the single-crystal seed can evolve into a cube, truncated cube, the introduction of an oxidative etchant to selectively etch away
cuboctahedron, truncated octahedron, octahedron, or rhombic seeds with twin defects and thus obtain single-crystal seeds
dodecahedron depending on the relative growth rates of {100}, only.18 While this strategy works effectively for the syntheses of
{111}, and {110} facets, which can be maneuvered via the use
of a facet-selective capping agent.12,13 For the seeds with Received: January 14, 2015
multiple twin defects and stacking faults, the final products will

© XXXX American Chemical Society A DOI: 10.1021/acs.nanolett.5b00158


Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

nanocrystals with a single-crystal structure, it cannot be applied


to those with twin defects or stacking faults. It has been
demonstrated that the symmetry of a seed could be reduced
during the growth process, depending on the exact
experimental conditions.19−21 Although symmetry reduction
can result in the formation of nanocrystals with novel shapes
deviated from the cubic symmetry of a noble metal, it will be
more complicated and challenging to control the uniformity
and purity of the products. Ideally, one needs a quantitative
knob that can be tuned to precisely manipulate the nucleation
and growth pathways in a predictable way by controlling a small
set of experimental parameters. We hypothesize that the
reduction rate of a precursor can serve as a quantitative knob
for manipulating the number of twin defects developed in a
seed and for controlling the growth pattern of a seed.
With Pd as a typical example, here we demonstrate that there
exists a quantitative correlation between the twin structure of
nanocrystals and the initial reduction rate of a synthesis. Our
study was inspired by the variable outcomes of a polyol
synthesis that typically involves the reduction of sodium
tetrachloropalladate(II) (Na2PdCl4) by ethylene glycol (EG) or
diethylene glycol (DEG) in the presence of poly-
(vinylpyrrolidone) (PVP).22,23 The twin structure undertaken
by the nanocrystals was found to be sensitive to the type of
polyol used and the reaction temperature. Figure 1a shows
transmission electron microscopy (TEM) image of the Pd
nanocrystals obtained using EG-based reduction at 140 °C. The
particles took a truncated octahedral shape with the surface
covered by a mix of {111} and {100} facets (see the inset for an
atomic model). The high-angle annular dark-field scanning-
TEM (HAADF-STEM) image in Figure 1b clearly shows a
single-crystal structure with a well-resolved, periodic arrange-
ment of atoms. In contrast, when the synthesis was conducted
in DEG and at 140 °C, multiply twinned icosahedra with an Figure 1. Palladium nanocrystals with three distinctive twin structures
average edge length of 5.8 nm were obtained in 100% purity obtained under different reaction conditions. The Pd nanocrystals
(Figure 1c). As illustrated in the inset, each icosahedron is were prepared using the standard procedure except for the variation in
enclosed by 20 {111} facets. The twin defects revealed by the the type of polyol and reaction temperature: (a,b) truncated octahedra
corresponding HAADF-STEM image (Figure 1d) confirm the in EG at 140 °C; (c,d) icosahedra in DEG at 140 °C; (e,f) nanoplates
involvement of a multiply twinned structure. When the reaction in DEG at 75 °C. (a,c,e) TEM images of the Pd nanocrystals produced
temperature of the DEG-based synthesis was decreased to 75 at each condition. The insets show the corresponding atomic models
of the nanocrystals with the twin planes or stacking faults being
°C, the final products contained 85% of nanoplates with an
delineated in red. (b,d,f) High-resolution HAADF-STEM images of
average edge length of 25 nm, together with 15% of multiply individual nanocrystals, revealing the different twin structures: (b)
twinned decahedra (Figure 1e). The HAADF-STEM image single-crystal, (d) multiply twinned, and (f) stacking fault-lined. Note
(Figure 1f) taken from the side face of a nanoplate indicates the that (f) was taken from the side face of a nanoplate.
existence of stacking faults along a direction normal to the
large, flat surfaces of a plate. Collectively, these results suggest
that Pd nanocrystals with distinctive twin structures, including by recording the UV−vis spectra of PdCl42−, the precursor to
single-crystal, multiply twinned, and stacking fault-lined, could Pd atom. At each point, 0.3 mL aliquot was sampled from the
all be obtained in high purity by simply varying the type of reaction solution using a glass pipet and immediately injected
polyol and/or the reaction temperature while keeping all other into 2.7 mL of 0.1 M aqueous HCl solution held at 0 °C. The
parameters unchanged. Because both the type of polyol and quick drop in temperature could immediately quench the
temperature will affect the reduction kinetics, we hypothesize reduction reaction and thus preserve the concentration of
that the initial reduction rate of a synthesis is the quantitative PdCl42−. The sample was then centrifuged to remove the Pd
knob responsible for the formation of seeds with different twin nanoparticles, followed by dilution of the supernatant with the
structures during the nucleation. By controlling this single same aqueous HCl solution to a level suitable for UV−vis
parameter, one would be able to deterministically obtain Pd measurement. The HCl played a critical role in preventing the
nanocrystals with different types of twin structures and thus precursor from hydrolysis when it was diluted with water to
distinctive shapes. lower concentrations.24
We have to know the rate law of a reaction in order to As shown in Figure 2, two absorption peaks appeared at 222
quantitatively analyze and compare the effects of reduction and 279 nm, similar to what was reported for PdCl42− in the
kinetics under different conditions. To this end, we derive the literature.25 The positions of these two peaks did not show any
concentrations of PdCl42− remaining in the reaction solution at shift even after the solution had been diluted with the aqueous
different time points after it had been introduced into a polyol HCl solution to a concentration as low as 1 μM, confirming the
B DOI: 10.1021/acs.nanolett.5b00158
Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Figure 2. Quantitative analysis of the reduction kinetics involved in the polyol synthesis of Pd nanocrystals. (a,b) UV−vis spectra of PdCl42−
remaining in the EG and DEG solutions after the reaction had proceeded for 0 to 30, 60, 90, 120, 180, 300, 450, and 600 s. (c) A plot showing the
percentages of PdCl42− remaining in the reaction solutions (calculated using the absorbance at 279 nm) as a function of reaction time. (d) A plot
showing the pseudo-first-order reaction kinetics involved in the syntheses of single-crystal and multiply twinned Pd nanocrystals in EG and DEG,
respectively, at 140 °C. The rate constants (k) were 1.12 × 10−2 and 5.57 × 10−4 s−1, respectively, for the EG- and DEG-based systems.

effectiveness of HCl in preventing the precursor from the precursor (PdCl42−, the reactant A), the concentration of B
hydrolysis. In the case of EG at 140 °C, the absorbance of should remain at a constant level during the entire synthesis. As
the precursor at 279 nm drastically decreased from 1.00 to 0.04 such, the kinetics involved could be approximated as a pseudo-
within just 5 min after the precursor had been introduced into first-order reaction with the rate law being simplified as
the reaction solution (Figure 2a). In comparison, the
absorbance of the precursor at 279 nm only dropped from −d[A]
rate = k[A] =
1.00 to 0.68 within 10 min when the reaction was conducted in dt (2)
DEG (Figure 2b). On the basis of the calibration curve This equation can be easily integrated to obtain
prepared using a set of Na2PdCl4 solutions with known
concentrations (Supporting Information Figure S1), we easily ln[A]t = −kt + ln[A]0 (3)
obtained the concentrations of PdCl42− remaining in the
reaction solution and the percentages of conversion from where [A]0 and [A]t represent the concentrations of the
PdCl42− to Pd(0) at different time points. As shown in Figure precursor at the beginning of a synthesis and at a specific time
2c, 96.5% of the PdCl42− was reduced to Pd atoms within 5 min point, respectively. In this case, a straight line will be obtained
in the EG-based synthesis and this value increased to 98.3% at t with a slope of −k when ln[A]t is plotted as a function of
= 10 min. In comparison, the conversion of the PdCl42− to Pd reaction time (t).
atoms in DEG was only 29.9% within the first 10 min, Figure 2d shows plots of ln[PdCl42−]t as a function of
suggesting a much weaker reducing power for DEG relative to reaction time for EG- and DEG-based syntheses, respectively,
EG. conducted at 140 °C. It can be seen that the values of
For an elementary reaction mA + nB → product, the reaction ln[PdCl42−]t indeed decrease linearly with reaction time,
rate can be described using the following equation confirming that the reduction was first-order with respect to
the precursor concentration. In addition, because the values of
rate = k[A]m [B]n (1) ln[PdCl42−]t did not show any abrupt changes deviated from
the linear plot, it is reasonable to assume that no autocatalytic
where k is the rate constant; [A] and [B] are the concentrations reaction was involved in the polyol synthesis described here.27
of reactants; and m and n represent the reaction orders. The From the slopes of the linear regression lines, the rate constants
rate constant is determined by a combination of multiple were determined to be 1.12 × 10−2 and 5.57 × 10−4 s−1 for the
experimental parameters, including the type of precursor, the reduction of PdCl42− by EG and DEG, respectively, at 140 °C.
type of reductant, solvent, and reaction temperature. In the Combined with the initial concentration of PdCl42− (1.76 ×
synthesis of noble-metal nanocrystals, the reduction of a salt 10−2 M), the corresponding initial rates were 1.97 × 10−4 for
precursor is expected to follow a second-order rate law due to EG and 9.80 × 10−6 M·s−1 for DEG (see Supporting
the involvement of collision and electron transfer between the Information Table S1 for details).
precursor and reductant.26 In the present study, because the We further extended the kinetic measurements to other
reductant (polyol, the reactant B) is in great excess relative to combinations of polyol and reaction temperature. As shown in
C DOI: 10.1021/acs.nanolett.5b00158
Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Figure 3. Determination of the rate constants under different reaction conditions and the activation energies. (a−c) Plots showing the
concentrations of remaining PdCl42− as a function of reaction time for different combinations of polyol and temperature. (d) A plot showing ln k as a
function of 1/T for three different reaction systems, where the slope of linear regression line can be used to calculate the corresponding activation
energy of the reaction using the Arrhenius equation.

Figure 3 and Table 1, the use of a polyol with a longer activation energy of a reaction. Using the reduction of PdCl42−
hydrocarbon chain or a lower reaction temperature would by DEG at 75 °C as an example, a significant drop in reducing
power and thus a much slower reduction rate is expected at
Table 1. Kinetic Parameters for the Reduction of PdCl42− by such a low temperature. As a result, it would be both
a Polyol at Different Reaction Temperatures challenging and time-consuming to experimentally determine
the kinetic parameters with high accuracy by using the UV−vis
temperature rate constant k
polyol (°C) (s−1) activation energy Ea (kJ·mol−1) method. It would be very straightforward to obtain the rate
EG 160 1.45 × 10−2
24.0
constant at this and other temperatures through the Arrhenius
150 1.28 × 10−2
equation once the activation energy of this reaction is known.
140 1.12 × 10−2
To this end, the rate constant for the reduction of PdCl42− by
130 8.75 × 10−3
DEG at 75 °C was found to be 2.26 × 10−6 s−1. Combined with
DEG 160 2.60 × 10−3 103.2
the initial concentration of PdCl42− (1.76 × 10−2 M), an initial
150 1.24 × 10−3
reduction rate of 3.98 × 10−8 M·s−1 was found to be
140 5.57 × 10−4 responsible for the formation of nanoplates with stacking faults
130 3.16 × 10−4 (Figure 1e).
TEG 160 1.40 × 10−3 110.7 As expected, the twin structure of the Pd nanocrystals would
150 5.78 × 10−4 change significantly when the reduction kinetics was adjusted
140 3.15 × 10−4 by varying the reaction temperature. Using the synthesis of
single-crystal, truncated octahedra of Pd in EG and at 140 °C as
an example, when the reaction temperature was raised to 160
decrease the rate constant. In addition, from the rate constants °C while keeping all other parameters the same, the particles in
at different temperatures, the activation energy (Ea) of a the product still had a single-crystal structure albeit they
reduction reaction involving PdCl42− and a specific polyol can showed a cuboctahedral shape (Supporting Information Figure
be derived from the Arrhenius equation S2a). The change from truncated octahedron to cuboctahedron
can be attributed to the enhancement in atom migration on the
k = Ae−Ea/ RT (6) surface as the temperature is increased.28 In contrast, multiply
where k is the rate constant of a reaction; A is the frequency twinned icosahedra started to appear when the synthesis was
factor; T is the reaction temperature (in Kelvin); and R is the conducted at 130 °C (Supporting Information Figure S2b),
universal gas constant (8.314 J·mol−1·K−1). Figure 3d shows a implying that the reduction rate associated with EG at 140 °C
plot of lnk as a function of 1/T for the reactions involving was very close to the minimum value required for the
PdCl42− and different polyols. For the reduction of PdCl42− by generation of single-crystal seeds in 100% purity. We also
EG, DEG, and triethylene glycol (TEG), their Ea values were observed a similar transition in twin structure for the seeds
found to be 24.0, 103.2, and 110.7 kJ·mol−1, respectively (Table when the synthesis was conducted in DEG at different
1). It is worth emphasizing the importance of knowing the temperatures. Multiply twinned icosahedra were obtained in
D DOI: 10.1021/acs.nanolett.5b00158
Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

100% purity when the reaction temperature was kept lower s−1) for generating multiply twinned seeds. The correlation
than or equal to 140 °C (Figure 1c and Supporting Information between the twin structure of Pd nanocrystals and the
Figure S2d). However, the products became a mix of both reduction kinetics was further validated by extending to polyol
single-crystal and multiple-twinned structures if the reaction syntheses based on triethylene glycol (TEG), as documented in
temperature was raised above 140 °C (Supporting Information Supporting Information Figure S3.
Figure S2c). When the reaction temperature was reduced below We also designed a set of experiments to demonstrate how
95 °C, the proportion of icosahedra in the product was found the twin structure of the product in a given synthesis can be
to gradually decrease, together with an increase in population controlled by manipulating reduction kinetics and be predicted
for nanoplates with stacking faults (Supporting Information based on the quantitative correlation shown in Figure 4.
Figure S2e,f). Combined together, these results qualitatively Specifically, for a synthesis with a known rate constant (k)
indicate that the twin structure developed in the seeds was determined experimentally (or derived from the Arrhenius
highly dependent on the reduction kinetics, which can be equation), the critical precursor concentration required for the
conveniently manipulated by varying a set of experimental formation of nanocrystals with a desired twin structure can be
parameters, including the types of precursor and reductant and readily estimated from the first-order rate law. For example,
their initial concentrations, as well as reaction temperature. when the reaction was conducted in EG at 140 °C with a
To quantitatively understand the effect of reduction kinetics relatively large rate constant, icosahedrons could still be
in determining the twin structure of Pd nanocrystals, we plot prepared as long as the concentration of Pd precursor was
the proportions of different types of seeds as a function of the low enough to ensure that the initial reduction rate was in the
initial reduction rates. The initial reduction rates were, in turn, range required for the formation of multiply twinned
calculated using the kinetic parameters acquired experimentally nanocrystals (Supporting Information Figure S4a,b). This
along with the initial precursor concentration (Supporting strategy was also applicable to the preparation of stacking-
Information Table S1). As shown in Figure 4, there is clearly a fault-lined nanoplates in DEG at 130 °C, where the proportions
of nanoplates gradually increased when the precursor
concentration was reduced from 1.76 × 10−2 to 3.52 × 10−4
M and 1.76 × 10−4 M, corresponding to initial reduction rates
of 1.11 × 10−7 and 5.56 × 10−8 M·s−1, respectively (Supporting
Information Figure S4c,d).
In summary, we have developed an effective method to
quantitatively analyze the reaction kinetics involved in a polyol
synthesis of noble-metal nanocrystals and further revealed a
quantitative correlation between the initial reduction rate and
the twin structure taken by seeds and nanocrystals. Such a
quantitative correlation allows us to easily control the twin
structure and thus shape taken by nanocrystals in a synthesis
without trial-and-error modification. With Pd as an example, we
have shown that simple spectroscopic measurements could give
kinetic parameters involved in the reduction of PdCl42− by a
polyol, including the rate constant, activation energy, and initial
Figure 4. A plot showing the relative distribution of the twin structure rate. By simply manipulating the initial reduction rate, Pd
of Pd nanocrystals as a function of the initial reaction rate (r0) of a
nanocrystals with distinctive twin structures, including single-
synthesis. The percentages of single-crystal, multiply twinned, and
staking fault-lined structures were acquired using TEM imaging and crystal, multiply twinned, and stacking-fault-lined, could be
are denoted in blue, black, and green colors, respectively. The twin deterministically obtained in high purity. It should be noted
planes or stacking faults are delineated in red. that the present approach might not be suitable for syntheses
involving complicated or step-by-step reactions. Nevertheless,
this work represents the first attempt to analyze the reaction
correlation between the twin structure of seeds and the initial kinetics involved in a synthesis of noble-metal nanocrystals and
reduction rate (r0). The nanocrystals gradually changed from to quantitatively correlate the initial reduction rate with the
stacking-fault-lined to multiply twinned and then single-crystal twin structure and shape taken by nanocrystals. This approach
as the initial reduction rate was increased from a level of 10−8 to not only greatly advances our understanding of the nucleation
10−6 and then 10−4 M·s−1. The rate that separates the adjacent and growth of nanocrystals but also paves the way for rational
groups of seeds with different twin structures can be readily design and deterministic synthesis of nanocrystals with desired
determined from this plot. Specifically, single-crystal seeds were and controlled twin structures and shapes.


always obtained in 100% purity when the initial rate was faster
than or equal to 1.97 × 10−4 M·s−1 whereas the initial rate had
ASSOCIATED CONTENT
to be controlled in the range of 8.70 × 10−6 M·s−1 to 4.0 × 10−7
M·s−1 in order to generate multiply twinned icosahedra in *
S Supporting Information
100% purity. When the initial rate was between the thresholds Experimental details and additional figures. This material is
for single-crystal and multiply twinned seeds (8.70 × 10−6 to available free of charge via the Internet at http://pubs.acs.org.


1.97 × 10−4 M·s−1), both single-crystal and multiply twinned
nanocrystals would appear in the final products. In the same AUTHOR INFORMATION
way, the product contained both multiply twinned and
stacking-fault-lined nanocrystals when the initial rate was Corresponding Author
further decreased below the lower threshold (4.0 × 10−7 M· *E-mail: younan.xia@bme.gatech.edu (Y.X.).
E DOI: 10.1021/acs.nanolett.5b00158
Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Present Address (23) Huang, H.; Wang, Y.; Ruditskiy, A.; Peng, H.-C.; Zhao, X.;
⊥ Zhang, L.; Liu, J.; Ye, Z.; Xia, Y. ACS Nano 2014, 8, 7041−7050.
Key Laboratory of Green Synthesis and Applications, College
of Chemistry, Chongqing Normal University, Chongqing (24) Timoshkin, A. Y.; Kudrev, A. G. Russ. J. Inorg. Chem. 2012, 57,
401331, People’s Republic of China 1362−1370.
(25) Wojnicki, M.; Pacławski, K.; Rudnik, E.; Fitzner, K. Hydro-
Author Contributions metallurgy 2011, 110, 56−61.
Y.W. and H.-C.P. contributed equally to this work. (26) Luty-Błocho, M.; Pacławski, K.; Wojnicki, M.; Fitzner, K. Inorg.
Y.X., Y.W., and H.-C.P. designed the studies; Y.W. and H.- Chim. Acta 2013, 395, 189−196.
C.P. performed the experiments; J.L. did high-resolution (27) Besson, C.; Finney, E. E.; Finke, R. G. Chem. Mater. 2005, 17,
HAADF-STEM imaging; Y.W., H.-C.P., and Y.X. analyzed the 4925−4938.
data; and Y.W., H.-C.P., and Y.X. wrote the paper together. All (28) Xia, X.; Xie, S.; Liu, M.; Peng, H.-C.; Lu, N.; Wang, J.; Kim, M.
authors discussed the results and commented on the manu- J.; Xia, Y. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 6669−6673.
script.
Notes
The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
This work was supported in part by the NSF (DMR-1215034)
and startup funds from Georgia Institute of Technology. As a
jointly supervised Ph.D. student from Southwest University,
Y.W. was also partially supported by a Fellowship from the
China Scholarship Council. J.L. gratefully acknowledges the
support by Arizona State University and the use of facilities in
the John M. Cowley Center for high-resolution electron
microscopy at Arizona State University.

■ REFERENCES
(1) Lewis, L. N. Chem. Rev. 1993, 93, 2693−2730.
(2) Burda, C.; Chen, X.; Narayanan, R.; El-Sayed, M. A. Chem. Rev.
2005, 105, 1025−1102.
(3) Li, Y.; Somorjai, G. A. Nano Lett. 2010, 10, 2289−2295.
(4) Huang, X.; Tang, S.; Mu, X.; Dai, Y.; Chen, G.; Zhou, Z.; Ruan,
F.; Yang, Z.; Zheng, N. Nat. Nanotechnol. 2011, 6, 28−32.
(5) Tian, N.; Zhou, Z.-Y.; Sun, S.-G.; Ding, Y.; Wang, Z. L. Science
2007, 316, 732−735.
(6) Talapin, D. V.; Lee, J.-S.; Kovalenko, M. V.; Shevchenko, E. V.
Chem. Rev. 2010, 110, 389−458.
(7) Anker, J. N.; Hall, W. P.; Lyandres, O.; Shah, N. C.; Zhao, J.; Van
Duyne, R. P. Nat. Mater. 2008, 7, 442−453.
(8) Stewart, M. E.; Anderson, C. R.; Thompson, L. B.; Maria, J.;
Gray, S. K.; Rogers, J. A.; Nuzzo, R. G. Chem. Rev. 2008, 108, 494−
521.
(9) Rosi, N. L.; Mirkin, C. A. Chem. Rev. 2005, 105, 1547−1562.
(10) Xia, Y.; Li, W.; Cobley, C. M.; Chen, J.; Xia, X.; Zhang, Q.;
Yang, M.; Cho, E. C.; Brown, P. K. Acc. Chem. Res. 2011, 44, 914−924.
(11) Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S. E. Angew. Chem., Int. Ed.
2009, 48, 60−103.
(12) Zeng, J.; Zheng, Y.; Rycenga, M.; Tao, J.; Li, Z.-Y.; Zhang, Q.;
Zhu, Y.; Xia, Y. J. Am. Chem. Soc. 2010, 132, 8552−8553.
(13) Chiu, C.-Y.; Li, Y.; Ruan, L.; Ye, X.; Murray, C. B.; Huang, Y.
Nat. Chem. 2011, 3, 393−399.
(14) Wiley, B.; Sun, Y.; Xia, Y. Acc. Chem. Res. 2007, 40, 1067−1076.
(15) Murphy, C. J.; Sau, T. K.; Gole, A. M.; Orendorff, C. J.; Gao, J.;
Guo, L.; Hunyadi, S. E.; Li, T. J. Phys. Chem. B 2005, 109, 13857−
13870.
(16) Xiong, Y.; Xia, Y. Adv. Mater. 2007, 19, 3385−3391.
(17) Maksimuk, S.; Teng, X.; Yang, H. J. Phys. Chem. C 2007, 111,
14312−14319.
(18) Zheng, Y.; Zeng, J.; Ruditskiy, A.; Liu, M.; Xia, Y. Chem. Mater.
2014, 26, 22−33.
(19) Zeng, J.; Zhu, C.; Tao, J.; Jin, M.; Zhang, H.; Li, Z.-Y.; Zhu, Y.;
Xia, Y. Angew. Chem., Int. Ed. 2012, 51, 2354−2358.
(20) Xia, X.; Xia, Y. Nano Lett. 2012, 12, 6038−6042.
(21) Wang, Y.; Xie, S.; Liu, J.; Park, J.; Huang, C. Z.; Xia, Y. Nano
Lett. 2013, 13, 2276−2281.
(22) Lv, T.; Wang, Y.; Choi, S.-I.; Chi, M.; Tao, J.; Pan, L.; Huang, C.
Z.; Zhu, Y.; Xia, Y. ChemSusChem 2013, 6, 1923−1930.

F DOI: 10.1021/acs.nanolett.5b00158
Nano Lett. XXXX, XXX, XXX−XXX

You might also like