Download as pdf or txt
Download as pdf or txt
You are on page 1of 59

Angewandte

A Journal of the Gesellschaft Deutscher Chemiker

International Edition Chemie www.angewandte.org

Accepted Article

Title: Surface Capping Agents and Their Roles in Shape-Controlled


Synthesis of Colloidal Metal Nanocrystals

Authors: Tung-Han Yang, Yifeng Shi, Annemieke Janssen, and


Younan Xia

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.

To be cited as: Angew. Chem. Int. Ed. 10.1002/anie.201911135


Angew. Chem. 10.1002/ange.201911135

Link to VoR: http://dx.doi.org/10.1002/anie.201911135


http://dx.doi.org/10.1002/ange.201911135
Angewandte Chemie International Edition 10.1002/anie.201911135

Surface Capping Agents and Their Roles in Shape-Controlled


Synthesis of Colloidal Metal Nanocrystals

Accepted Manuscript
Tung-Han Yang,† Yifeng Shi,† Annemieke Janssen,† and Younan Xia*

[*] Dr. T.-H. Yang, Prof. Dr. Y. Xia


The Wallace H. Coulter Department of Biomedical Engineering
Georgia Institute of Technology and Emory University, Atlanta, GA 30332 (USA)
Y. Shi, Prof. Y. Xia
School of Chemical and Biomolecular Engineering
Georgia Institute of Technology, Atlanta, GA 30332 (USA)
A. Janssen, Prof. Y. Xia
School of Chemistry and Biochemistry
Georgia Institute of Technology, Atlanta, GA 30332 (USA)
E-mail: younan.xia@bme.gatech.edu

These authors contributed equally to the preparation of this review article.

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

TOC

Surface capping agents play a pivotal role in directing the growth of colloidal metal nanocrystals
into diverse but well-controlled shapes. This article offers a comprehensive review of capping
agents, including general principles, characterizations, computational simulations, mechanistic
understandings, and removal, as well as their use in engineering the surface structures and catalytic

Accepted Manuscript
properties of metal nanocrystals.

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Abstract
Controlling the shapes and related properties of colloidal metal nanocrystals are key to the
realization of their vast applications. Surface capping agents, including neutral molecules, ionic
species, macromolecules, and biomolecules, have been extensively used to control the evolution
of seeds into nanocrystals with diverse but well-controlled shapes. Here we offer a comprehensive
review of these agents, with a focus on the mechanistic understanding of their roles in guiding the

Accepted Manuscript
shape evolution of metal nanocrystals. We begin with a brief introduction to the early history of
capping agents in electroplating and bulk crystal growth, followed by discussion of how they affect
the thermodynamics and kinetics involved in a synthesis of metal nanocrystals. We then present
representative examples involving both experimental and computational studies to highlight the
various capping agents, including their binding selectivity, molecular-level interaction with a metal
surface, and impacts on the growth of metal nanocrystals. We also showcase progress in leveraging
capping agents to generate nanocrystals with complex structures and/or enhance their catalytic
properties. Finally, we discuss various strategies for the exchange or removal of capping agents,
together with perspectives on future directions.

Keywords: Capping agent, metal nanocrystal, crystal growth, shape control, structure-property
relationship

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

From the Contents


1. Introduction
2. From Macroscopic Systems to Nanocrystals
3. Thermodynamics versus Kinetics
4. Case Studies
5. Computational Investigation

Accepted Manuscript
6. Characterization Tools
7. Nanocrystals Free of Capping Agent
8. Concluding Remarks

1. Introduction
Colloidal metal nanocrystals have received ever-increasing interest because of their unique
position in bridging bulk solids with atomic/molecular species.[1-6] In addition, nanocrystals made
of Au, Ag, Cu, Pd, Pt, Rh, Ru, and Ir have distinctive properties for applications in photonics,[7]
electronics,[8] catalysis,[9,10] energy conversion,[11,12] sensing,[13-16] and biomedicine.[17,18] In
general, the properties of a metal nanocrystal is determined by a set of physical parameters that
include composition, size, shape, and crystal structure. One can maneuver any one of these
parameters to tailor their properties and thus optimize their performance in a specific application.
Among these parameters, shape has been most extensively explored due to its strong correlations
with both photonic and catalytic properties.
A notable example can be found localized surface plasmon resonance (LSPR), which arises
from the collective oscillation of conduction electrons in response to the electromagnetic field of
incident light.[14,19] In this case, the shape determines how the conduction electrons in a nanocrystal
are polarized and how the charges are distributed on the surface, controlling the number,
wavelengths, and intensities of the LSPR peaks, as well as the ratio between absorption and
scattering. For Au spheres with a diameter of 50 nm, they have a strong LSPR absorption peak
around 525 nm, resulting in a ruby red color for their aqueous suspension. [20] When switched to
triangular plates with an edge length of 68 nm, two LSPR peaks (a strong one at 700 nm and a
shoulder around 550 nm) are observed, corresponding to the in-plane and out-of-plane dipole
resonances, respectively.[21] Their aqueous suspension gives a blue color. The ability to tune the
LSPR peak position has immediate implications for biomedical applications such as optical
4

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

contrast enhancement and photothermal therapy because soft tissues have a transparent window in
the near-infrared region (650-900 nm) to allow deep penetration.[22]
In addition to the optical properties, one can tailor (electro)catalytic activity and/or selectivity
of metal nanocrystals by maneuvering their shape to control the arrangement of atoms on the
surface.[23,24] Taking the oxygen reduction reaction (ORR) as an example, it was reported that the
specific activity (i.e., current density normalized to the electrochemically active surface area) of

Accepted Manuscript
Pt concave cubes (1.77 mA cm–2) enclosed by {720} high-index facets was 6.3 and 1.3 times
greater than those of Pt cubes and octahedra covered by {100} and {111} facets, respectively.[25]
Similar correlations between catalytic properties and surface structures have also been observed
for many other reactions such as carbon dioxide reduction,[26] formic acid oxidation,[27]
hydrogenation,[28] epoxidation,[29] and isomerization.[30] These and other examples demonstrate the
importance of shape control in tailoring the properties of metal nanocrystals and optimizing their
performance in catalytic applications.
With the development of nanochemistry, metal nanocrystals with a wide variety of shapes
have been synthesized using either one-pot or seed-mediated method.[1,5,24] In a one-pot synthesis,
the formation of nanocrystals can be separated into two stages: i) homogeneous nucleation--
assembly of atoms from precursor reduction to generate nuclei and then seeds and ii) growth—
addition of atoms onto the surfaces of seeds for their evolution into nanocrystals with distinct
shapes. As for seed-mediated growth, pre-synthesized nanocrystals with defined shapes and twin
structures are added into a growth solution for the deposition of atoms from the same or a different
metal. In either case, the shape of the nanocrystals can be changed by manipulating experimental
parameters such as temperature, solvent, metal precursor, reducing agent, colloidal stabilizer, and
most importantly, surface capping agent. In most cases, it is the capping agent that plays the
decisive role in dictating the shape evolution.
Here we aim to offer a comprehensive review of surface capping agents, with a focus on our
current understanding of their working mechanisms according to experimental and theoretical
studies. We present a number of examples to highlight the power of capping agents in nanocrystal
syntheses, with a focus on both qualitative understanding and quantitative analysis. We also
discuss recent progress in utilizing the blocking effect of capping agents to construct nanocrystals
with anisotropic and other complex structures; removing the capping agent after a synthesis or
developing methods for capping agent-free syntheses; and promoting the catalytic performance of
5

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

nanocrystals through electronic coupling with a capping agent.

2. From Macroscopic Systems to Nanocrystals


The concept of using a capping agent to modify a metal surface during its formation can be
traced back to electroplating, in which a layer of metal is deposited on the surface of a substrate
(serving as a cathode) immersed in a plating bath containing a metal precursor.[31-33] The precursor

Accepted Manuscript
is reduced by applying a current to generate metal atoms for their deposition onto the surface of
the substrate. As shown in Figure 1A, the color of a Cr-Ni-Fe alloy spoon changes dramatically
when its surface is electroplated with different metals, including Ag, Cu, Au, Pt, Pd, Rh, and Ni.[31]
The electroplated coating can have a major impact on the morphology, stability, and function of
the underlying substrate. The morphology of an electroplated object critically depends on the metal
deposition rates across the surface. If the rates are the same across the surface, a uniform layer will
be deposited, preserving the original morphology and roughness of the surface (Figure 1, B and
C). However, on a surface with a recessed region, the electric field and current density will be
concentrated at the corners and edges, resulting in higher rates of deposition and thus thicker layers
at these sites. As a result, the surface roughness will increase during electroplating (Figure 1D).
To address this issue, a “levelling agent” (e.g., thiourea or coumarin) is introduced to ensure more
or less even deposition across the surface.[32,33] By selectively passivating the corners and edges of
the recessed region, the additive can slow down the electroplating rates at these sites (Figure 1E),
enabling the formation of an evenly deposited layer.
Additives known as “habit modifiers” were widely used to control the growth behaviors and
thus shapes of bulk crystals long before nanocrystal synthesis became a subject of research. During
the growth, the habit and thus shape of a crystal could be altered through the introduction of an
ionic or molecular additive. Taking CuSO4·5H2O as an example, crystals grown from an aqueous
solution often display a mix of different facets.[34] In the presence of Fe3+, however, the surface is
dominated by {100} and {110} facets only. On the other hand, the introduction of a detergent such
as sodium lauryl sulfate promotes the formation of {11̅0} facets. Again, these additives are
believed to be able to adsorb onto different facets through selective binding, altering the growth
rates along different directions.
The capping agent is equivalent to the levelling agent employed in an electroplating process
or the habit modifier involved in the growth of a bulk crystal; they are all supposed to operate by
6

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

the same principle despite the difference in length scale. Owing to the selectivity of a capping
agent towards a specific type of facet on a growing seed, it can significantly alter the landscape of
surface free energy, growth habit, and even growth mechanism of the seed, leading to nanocrystals
with shapes deviated from the Wulff polyhedron.[35,36] In early studies, Pt nanoparticles supported
on amorphous SiO2 or Al2O3 were found to change their shape upon annealing under different
environments such as N2, H2, or H2S.[37,38] Although the exact mechanism of shape transformation

Accepted Manuscript
was not clear at that time, these studies demonstrated the feasibility to control the shape of metal
nanocrystals using gas molecules. In 1996, it was reported that a polymeric capping agent such as
sodium polyacrylate could be used to control the evolution of Pt nanocrystals into tetrahedral,
cubic, cuboctahedral, icosahedral, and irregular-prismatic shapes in an aqueous synthesis.[39] In
2002, our group demonstrated that Ag nanocubes could be obtained in the presence of poly(vinyl
pyrrolidone) (PVP) as a capping agent towards {100} facets and thus the elimination of {111}
facets in a polyol synthesis.[40] In the same year, we further demonstrated the growth of multiply-
twinned, Ag decahedral seeds into one-dimensional (1D) nanostructures such as pentagonal rods
and wires.[41-44] Again, the success of this synthesis was attributed to the selective binding of PVP
towards {100} facets, facilitating growth along the axial direction of a decahedral seed.
The phenomenon of adsorption on a solid surface can be described using a simple model
known as the Langmuir isotherm (Figure 2).[45] Despite the involvement of multiple assumptions,
the Langmuir isotherm can be easily used to fit the adsorption data. At equilibrium, the fraction
(θ) of the surface sites occupied by the adsorbate molecules can be expressed as:
𝐾𝐶𝑏𝑢𝑙𝑘
𝜃 = 𝐾𝐶 (Eq. 1)
𝑏𝑢𝑙𝑘 +1

where K stands for the equilibrium constant of adsorption and Cbulk is the concentration of the
adsorbate in the bulk solution. This equation indicates that the driving force for surface adsorption
can come from the high concentration of the adsorbate in the bulk solution and/or a strong affinity
of the adsorbate toward the surface. The surface adsorption of a capping agent should also follow
this simple model. What differentiates it from other types of adsorbates is the involvement of a
stronger binding such as covalent bonding in most cases and the difference in binding affinity
toward different types of facets. When a molecular species is capable of binding (chemically or
physically) to one particular type of facet more strongly than the other types, it should be called a
capping agent.

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

By leveraging the concept of surface capping, a variety of shapes have been achieved for
nanocrystals made of different metals and their alloys (Table 1).[40-44,46-103] These results establish
that surface capping offers an effective and versatile means for maneuvering the shape of
nanocrystals by controlling the relative surface areas of different types of facets. In this Review,
we aim to illustrate the power of surface capping in controlling the shape and thus surface structure
of metal nanocrystals. Most importantly, through case studies of experimental and theoretical

Accepted Manuscript
origins, we seek to highlight the correlations between the shape taken by a nanocrystal and the
binding selectivity of a capping agent while shedding light on the growth mechanism.

3. Thermodynamics versus Kinetics


At a fixed number of atoms, the nanocrystal can take many different forms, including those
with distinctive shapes and internal structures. Which form will prevail in the final product? The
answer to this question depends on both thermodynamics and kinetics.[104,105] The essence of these
two concepts can be explained using the energy diagram shown in Figure 3. As a form with the
least free energy, the thermodynamic product is supposed to reside at the global minimum. Any
form deviated from the global minimum is a kinetically-controlled product, a metastable form
trapped at a local minimum. All kinetic products are supposed to spontaneously transform into the
thermodynamic product, but they can be stable for different periods of time depending on the
thermal energy available for crossing the activation energy barrier. This can be illustrated using an
example involving diamond and graphite, two allotropes of carbon. Although diamond has a higher
total free energy relative to graphite, diamond can be considered “forever” once it has been created
because the kinetics is infinitely slow.
As for the shape evolution of nanocrystals, capping agent can modify both thermodynamic
and kinetic factors through selective adsorption.[106] By lowering the surface energy of a certain
type of facet, the capping agent can change the shape of nanocrystals favored by thermodynamics.
Alternatively, by posing a physical barrier to atom deposition and diffusion, the capping agent can
affect the outcome of shape evolution through kinetic means.

3.1. Thermodynamic Control


According to thermodynamics, the most favorable form of a nanocrystal can be attained by
minimizing its total free energy. For a nanocrystal, the energy of formation can be divided into
8

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

two parts: bulk and surface. Since our discussion on the most stable form of a nanocrystal involves
the same number of atoms, we can ignore the bulk part and just focus on the surface part, which is
determined by the energy required to create a new surface.[107] The surface area of a nanocrystal is
not the only determinant of total surface energy because it also depends on the crystallographic
planes involved. Figure 4A shows the three low-index planes of a face-centered cubic (fcc)
crystal.[104] For a surface, the more dangling bonds it contains, the less stable it will be and thus

Accepted Manuscript
the higher the specific surface energy. In this case, each atom on the (111) plane has three dangling
bonds. When it comes to the (100) and (110) planes, the number of dangling bonds increases to
four and six, respectively. As a result, the specific surface energy increases in the order (111) <
(100) < (110). In general, the more closely packed a given crystallographic plane is, the lower the
specific surface energy will be and vice versa.[107]
The thermodynamically favored shape with the lowest surface energy is known as a Wulff
polyhedron.[35] Based on a few parameters, including the relative surface energies of different
facets and the crystal structure, one can construct the Wulff shape. However, the presence of
capping agent will anisotropically change the surface energies of different crystallographic planes.
The capping agent selectively adsorbed onto a facet can stabilize the surface atoms and thus
decrease the surface energy. To this end, Wulff construction needs to be modified. [36] A simple
illustration of this concept is shown in Figure 4B. During the growth of an fcc nanocrystal, the
presence of a capping agent for {100} will passivate these facets to lower their surface energy.[104]
Therefore, the thermodynamically favorable shape will change from a cuboctahedron to a cube
enclosed by six {100} facets. The same argument is also valid when the synthesis is conducted in
the presence of a capping agent for {111}. In this case, octahedra encased by eight {111} facets
will be obtained to minimize the total surface energy.
A typical example can be found in the synthesis of Ag nanocubes as mediated by Cl– ions.[47]
Typically, Cl– ions are used with PVP to direct the evolution of Ag nanocrystals into a cubic shape.
A recent report analyzed the changes in surface energy for both Ag(111) and Ag(100) as the
coverage of Cl– was increased. Experimentally, it was noted that Ag nanocrystals evolved from
truncated octahedra to cuboctahedra, and finally to cubes with sharp corners and edges, as the
concentration of Cl– in the reaction solution was increased. The ab initio calculation showed that
the strong binding of Cl– to Ag(100) reversed the order of surface energies for (100) and (111).
Altogether, it was concluded that the cubic shape was a product of thermodynamic control enabled
9

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

by the capping agent.

3.2. Kinetic Control


Capping agent can also affect the shape of a nanocrystals from the kinetic perspective.
Instead of lowering the surface energy, the adsorption of capping agent on certain crystal facets
can pose a physical barrier to the surface.[5,54,108] As such, the capping layer will slow down the

Accepted Manuscript
deposition rate of atoms. Meanwhile, the presence of this passivating layer will retard the surface
diffusion of adatoms. By carefully controlling reaction conditions, a variety of shapes deviated
from the Wulff polyhedron can be obtained. Figure 5 shows the effect of surface diffusion on the
shape evolution of a 2-D nanocrystal. The square seed is originally covered a capping agent that
prevents the side faces from receiving atoms. As a result, the atoms newly formed in the solution
are preferentially deposited at the corners. The final morphology will be determined by the relative
rates responsible for deposition and diffusion. Fast diffusion of adatoms on the surface will give a
thermodynamic product further to the right, whereas slow diffusion of adatoms will give kinetic-
controlled products such as the multipod on the left side.
In essence, the competition between thermodynamics and kinetics is settled by the kinetics
of surface diffusion.[108,109] With enough energy to overcome the diffusion barrier, adatoms will
spontaneously migrate across the surface to achieve the lowest total surface energy for the system
and thus the formation of a thermodynamically-controlled shape. If the diffusion barrier is too high
to overcome, the product will stay at the local minimum with a higher total surface energy and
thus the formation of a kinetic product. Although the thermodynamic and kinetic controls seem to
be distinguishable, it is often hard to completely decouple one effect from the other in a real
synthesis. When experimental conditions are varied, the growth can switch from thermodynamic
to kinetic control and vice versa. For most syntheses, both mechanisms tend to be involved in
directing the evolution of shape.

3.3. The Dynamic Nature of Surface Capping


In order to grow into a specific size and shape, atoms have to be deposited on the surface of
a nanocrystals and allowed to diffuse around. This means the capping must be a dynamic process
in order to allow new atoms to be continuously deposited. This dynamic process can be enabled
through three different scenarios. Firstly, the capping agent constantly desorbs and adsorbs during
10

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

the growth of nanocrystal by breaking and forming the bonds between the capping agent and metal
surface. This allows new atoms to nucleate on the instantaneously exposed regions of a surface,
maintaining the growth process. Secondly, the new atoms are deposited on the region free of
capping agent due to the facet selectivity of a capping agent. The deposited atoms can then “slide
in” between the capping layer and metal surface through surface diffusion. To this end, the
interaction between the capping agent and metal needs to be kept relatively weak to facilitate the

Accepted Manuscript
sliding process. Thirdly, the capping layer only interacts with the metal surface weakly through
van der Waals (vdW) forces to maintain some free space in between.[110] In this case, however, the
capping agent may lack facet selectivity due to the missing of relatively strong chemical bonds.
There is a need to develop methods capable of resolving this dynamic process by monitoring the
variation in the surface coverage of a capping agent with the right temporal resolution. Techniques
such as second-harmonic generation (SHG) might be suitable for this purpose.[111]

4. Case Studies
Table 1 shows a list of capping agents commonly used for controlling the shapes of noble-
metal nanocrystals.[40-44,46-102] It is generally accepted that the capping agent works more
effectively in the growth stage than in nucleation. However, some capping agents may coordinate
with the precursor ions to form of new complexes during a synthesis.[112] Such ligand exchange
can have a profound impact on the reduction kinetics and thus the outcome of a nucleation process.
Here we cover capping agents capable of generating nanocrystals with controlled shapes through
facet-selective binding via both one-pot and seed-mediated routes.

4.1. Poly(vinyl pyrrolidone) (PVP)


As the most commonly used stabilizer for the synthesis of colloidal particles, PVP is also an
effective capping agent for the synthesis of metal nanocrystals enclosed by {100} facets (Figure
6). It was first reported in 2002 for the synthesis of Ag nanocubes through a polyol method based
on ethylene glycol (EG).[40] It was proposed that PVP could preferentially adsorb onto Ag(100)
surface to reduce its growth rate, facilitating the formation of nanocubes enclosed by {100} facets.
In the original report, the edge lengths of the Ag nanocubes could be tuned in the range of 80−175
nm by adjusting the reaction time, temperature, and AgNO3 concentration. To produce smaller Ag
nanocubes with sizes ranging from 30−70 nm, the PVP-assisted polyol method was modified by
11

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

introducing trace amounts of NaSH (for the formation of single-crystal seeds) and HCl (for the
selective oxidative etching and dissolution of twinned seeds) and by switching the precursor from
AgNO3 to CF3COOAg.[113] Furthermore, by replacing the most commonly used EG with
diethylene glycol (DEG), the size of the Ag nanocubes was further reduced to 18−32 nm.[53] When
the size drops below 15 nm, however, the long-chain polymer may not be able to effectively
stabilize the small {100} facets anymore,[49] and halide ions have to be introduced (Section 4.2).

Accepted Manuscript
In addition to nanocubes, it was shown that 1D nanostructures such as Ag penta-twinned
nanowires could be prepared using a modified polyol method (Figure 6B).[41-44] In this case, one
needed to reduce the concentration of AgNO3 to slow down the reduction kinetics while keeping
the molar ratio of PVP to AgNO3 the same as what was used for nanocube synthesis. Specifically,
it was feasible to attain the right kinetics for the formation of decahedral seeds. In the growth
process, the Ag atoms were preferentially deposited along the twin boundaries and the decahedral
seed was uniaxially elongated into a penta-twinned nanorod and then nanowire covered by {100}
facets on the side faces. PVP could selectively bind to the {100} facets, facilitating the longitudinal
growth of nanorods into nanowires with aspect ratios as high as ca. 1000. In recent studies, similar
growth mechanisms were also observed in the syntheses of Cu and Pd penta-twinned rods/wires
from decahedral seeds, albeit halide ions were also typically involved (Section 4.2).[73,114,115] In
contrast, PVP selectively passivated the Au(111) surface and ultimately induced the formation of
Au nanocrystals enclosed by {111} facets, such as octahedra and icosahedra.[65] This argument is
also consistent with the result of a recent density functional theory (DFT) study, in which the PVP-
covered Au(111) surface was found to be thermodynamically more stable relative to Au(100).[116]

4.2. Halides
In general, PVP is too bulky to effectively cap the surface of metal nanocrystals with sizes
below 15 nm. As an alternative, halide ions (e.g., Cl–, Br–, and I–) can form a close-packed adlayer
on the face of small nanocrystals via strong covalent or coordination binding.[117] Most importantly,
they are able to preferentially chemisorb onto the {100} facets of nanocrystals made of Ag, Pd, Pt,
Rh, and Pd-Pt alloys. As shown in Figure 7A, Pd nanocubes encased by {100} facets can be made
by employing Br– as a capping agent.[70] In a typical synthesis, aqueous Na2PdCl4 (precursor) was
injected into an aqueous mixture containing L-ascorbic acid (reducing agent), PVP (in this case,
stabilizer), and KBr (capping agent). Similarly, Ag, Pt, and Rh nanocubes were also obtained in
12

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

the presence of halide ions (Figure 7, B-D).[49,81,82] In addition, it was shown that the presence of
Br− ions at an adequate concentration could promote the formation Pd-Pt alloy nanocrystals with
a cubic shape and an edge length of ca. 6 nm.[97] This is because the Br− ions could alter the ratio
between the initial reduction rates of the PdCl42− and PtCl42− precursors through a ligand exchange
process while capping the {100} facets of the alloy nanocrystals.
If the concentration of halide ions is below a critical value, the resultant nanocubes are likely

Accepted Manuscript
covered by a submonolayer of halides ions.[48,68,72] Under this condition, oxidative etching (i.e.,
removal of metal atoms from the surface) may selectively occur on one of the side faces, making
this face most active for the subsequent deposition of atoms. As a result, anisotropic growth is
initiated to elongate a nanocube into a nanobar whose surface is still bound by {100} facets. This
phenomenon is often observed in the Pd and Ag systems (Figure 7, E and F).[48] When switched to
I− ions, Pd penta-twinned nanowires were obtained with an average diameter of 7.8 nm and aspect
ratios up to 100 (Figure 7G).[73] In this case, I− ions not only tuned the reduction rate of the Pd(II)
precursor into the proper regime favorable for the formation of decahedral seeds,[118] but also acted
as a capping agent for {100} facets. Using a similar strategy, Br− ions were used with PVP as
capping agents for the Ag(100) surface to generate Ag penta-twinned nanowires with diameters
below 20 nm, together with aspect ratios over 1000.[50] The nanowires could be formed rapidly
(within 35 min) and in high (>85%) morphology purity (Figure 7H). In situ liquid-cell transmission
electron microscopy (TEM) was also used to analyze the evolution of Au decahedral seeds into
penta-twinned nanorods in the presence of Br− ions, as shown in Figure 7I.[119] The sequential
TEM images revealed that the formation of Au penta-twinned nanorod started from the growth of
a regular decahedron and proceeded to reentrant groove-induced anisotropic growth of a truncated
decahedron and, eventually, to penta-twinned nanorods with increasing aspect ratios.

4.3. Citric Acid (CA) and Citrate


CA and citrate are commonly used as reducing agents for the syntheses of metal colloids.
They can also serve as capping agents for some metals. For example, CA possesses a hydroxyl
and three carboxylic acid groups that match well with the hexagonal symmetry of (111) surface of
fcc metals.[76,77,94] As a result, CA can selectively passivate {111} facets on Pd or Rh nanocrystals
through chemisorption. For the Pd system, the ratio of PdCl42– (1.2−17.4 mM) to CA (10−84 mM)
could be manipulated for the production of Pd polyhedral nanocrystals (Figure 8, A-D), including
13

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

octahedra, icosahedra, decahedra, and plates, that are all dominated by {111} facets.[76,77] In this
protocol, CA acts as a mild reducing agent and a capping agent for Pd(111) surface. Taking the
synthesis of Pd octahedra as an example, the product contained cuboctahedral particles enclosed
by a mix of {100} and {111} facets at t = 1 h. The product became octahedra bound by {111}
facets at t = 20 h. Such shape evolution suggests that CA favors the formation of octahedra due to
its strong binding towards {111} instead of {100} facets during the growth stage. Using a similar

Accepted Manuscript
strategy, Rh tetrahedra enclosed by four {111} facets were also synthesized.[94]
Like CA, citrate was found to preferentially bind towards Ag(111).[56-58] A possible
explanation can be attributed to the match in symmetry and dimensions between this compound
and Ag(111) surface. In one study, it was shown that citrate was critical to the formation of Ag
nanoplates due to its ability to passivate the two {111} basal planes and thus limit the increase in
plate thickness.[56] By adjusting the ratio of Ag(I) precursor to citrate, Ag nanoplates with
controllable lateral dimensions and thickness were obtained.[57,58] A similar protocol was also
reported for the synthesis of Au nanoplates.[59,60]

4.4. Carbon Monoxide (CO)


Recently, CO from a variety of sources, including the direct use of CO gas or decomposition
of W(CO)6 or Fe(CO)5, has been adopted to control the shape of Pd, Pt, or Pt-M (M = Pd, Co, Cu,
Ni, and Rh) alloy nanocrystals owing to its facet-selective of CO binding.[79,80,84-86,101-103] The CO
adsorbed on metal nanocrystals could be effectively removed by exposing the sample to air,
offering a major advantage for catalytic applications. An early study demonstrated the synthesis
of ultrathin, hexagonal Pd nanoplates using CO as a capping agent for Pd(111) surface (Figure
9A).[79] In the synthesis, Pd(acac)2, cetyl trimethyl ammonium bromide (CTAB) , and PVP were
dissolved in dimethylformamide (DMF) to initiate homogeneous nucleation, and the solution was
then transferred into a glass pressure vessel under 1 bar of CO for the formation of nanoplates.
Due to the strong adsorption of CO molecules on the basal (111) planes, the resulting Pd nanoplates
were only 1.8 nm in thickness and 96% of the exposed surface was terminated in {111} facets. By
introducing H2 to vary the adsorption strength and coverage of CO on Pd(111) planes, single-
crystal Pd tetrapods and tetrahedra enclosed by {111} facets were obtained (Figure 9, B and C).[80]
In contrast to Pd, CO molecules form a much stronger chemical bond with Pt(100) rather
than Pt(111), and this selectivity led to Pt nanocrystals with a cubic shape (Figure 9D).[84-86] A
14

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

number of research groups also found that the introduction of a second metal such Ni, Co, and Rh
could result in the formation of octahedral Pt-Ni (Figure 9, E and F), Pt-Co, and Pt-Rh alloy
nanocrystals covered by {111} facets.[86,101-103] When the second metal was switched to Cu,
however, the Pt-Cu alloy nanocrystals took a cubic shape similar to that of Pt.[101] The formation
of cubic or octahedral shape can be attributed to the facet selectivity of CO. The binding between
CO and Cu is much weaker than that between CO and Pt and thus the inclusion of Cu had

Accepted Manuscript
essentially no impact on the growth behavior of the nanocrystals. In contrast, CO binds strongly
to Co, Ni, and Rh atoms on the {111} facets and the inclusion of such a metal will compete with
that of Pt, leading to the formation of octahedral nanocrystals when its content is high enough to
dominate the adsorption of CO. This speculation was supported by the experimentally observed
shape transformation from cubic Pt to cuboctahedral Pt6Ni and finally to octahedral Pt3Ni when
the Ni content was gradually increased.[101] Computational modeling will help us understand the
binding selectivity of CO towards an alloy surface, and the results are expected to enable rationally
designed synthesis of bimetallic nanocrystal with controlled shapes.

4.5. Peptides
Peptides with various sequences and conformations have also been explored as capping
agents to control the shapes of metal nanocrystals.[88-90] Figure 10A summarizes a study based on
Pt.[89] The authors examined two peptides: Acyl–Thr–Leu–Thr–Thr–Leu–Thr–Asn–CONH2
(sequence: TLTTLTA; termed peptide T7) for {100} facets and Acyl–Ser–Ser–Phe–Pro–Gln–
Pro–Asn–CONH2 (sequence: SSFPQPN; termed peptide S7) for {111} facets. These peptides
could guide the formation of Pt cubic and tetrahedral nanocrystals, respectively, with their sizes
controlled below 10 nm. In the presence of peptide T7, the Pt nanocrystals evolved from
cuboctahedra covered by a mix of {100} and {111} facets to cubes encased by {100} facets,
suggesting the selective stabilization of {100} facets by peptide T7. When switching to peptide S7,
the nanocrystals showed a cuboctahedral shape in the early stage and then evolved into tetrahedral
shape with {111} facets on the surface, indicating the selectivity of peptide S7 towards Pt{111}.
The correlations between the conformational structures of peptides T7 and S7 and the different
atomic arrangements on (100) and (111) planes might be responsible for the facet selectivity. The
highly hydroxylated (−OH) “TLT” end of peptide T7 makes it possible for oxygen atoms to bind
to the Pt atoms on (100), generating a monolayer to stabilize this surface. In comparison, peptide
15

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

S7 has the phenyl group (−C6H5) in the “F” unit and it prefers (111) surface due to the hexagonal
array of atoms.
It is worth noting that the peptide concentration also has a profound impact on the nanocrystal
synthesis. For example, the synthesis of Pt cubes in the presence of peptide T7 could only be
achieved at intermediate peptide concentrations (10−20 µg/mL), which can be attributed to the
attraction and repulsion resulting from the competition between the peptide and water molecules

Accepted Manuscript
on the Pt(100) surface at different concentrations of peptide T7.[90] As shown in Figure 10B, the
authors were able to apply the principle to the design of small organic molecules with the phenyl
ring (e.g., 3-hydroxy-2-phenylpropanoic acid) for the synthesis of tetrahedral and octahedral Pt or
Rh nanocrystals covered by {111} facets by leveraging the epitaxial coordination between phenyl
ring and {111} facets.[88]

4.6. One-pot versus Seed-mediated Growth


The examples shown in Sections 4.1 to 4.5 all involve the one-pot approach. In recent years,
capping agents have also been combined with seed-mediated growth to obtain nanocrystals with a
variety of complex shapes or structures that are otherwise difficult to achieve via homogeneous
nucleation or one-pot synthesis. By choosing the right capping agent to passivate a specific type
of facet on the preformed seeds, different growth rates for different sites on the seeds can induce
specific growth pattern and eventually result in the evolution of seeds into predictable shapes. In
one study, it was demonstrated that single-crystal, spherical Ag seeds could be directed to grow
into nanocrystals with octahedral and cubic shapes by using CA and PVP as the capping agents,
respectively.[120] This method has also been applied to investigate growth patterns of Ag nanoplates
with different capping agents.[121] The nanoplate seeds are enclosed by two {111} basal planes as
the top and bottom faces and contain twin planes and stacking faults on the side faces along the
vertical direction. With the use of CA as a capping agent, lateral growth was more favorable,
leading to increase in edge length for the nanoplates. This growth pattern can be attributed to the
selective adsorption of CA on the Ag{111} facets. In contrast, when CA was substituted with PVP,
thicker plates were obtained due to the vertical growth caused by the relatively stronger binding
of PVP to {100} than {111}. It is worth pointing out that the overall growth rate was significantly
faster for the case of CA than PVP. Such difference can be explained by the fact that twin planes
and stacking faults on the side faces should have higher surface energies than those located on
16

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

both the top and bottom faces. Similar results were also observed in the growth of Pd nanoplates
into enlarged plates during seed-mediated growth in the presence of a low concentration of Br− as
a capping agent for Pd{100}. When Br− ions were used at a higher concentration, all three of the
{100} side faces of the triangular plates would be completely blocked by Br− ions, forcing the Pd
nanoplate to grow from three of the corners for the generation of tripods.[122]
In addition to the monometallic system, seed-mediated growth has been used to generate a

Accepted Manuscript
number of bimetallic nanocrystals. In general, the products are limited to a core-shell structure due
to a thermodynamic control that forces the deposition of the second metal to occur in a conformal
fashion.[24,108] Compared to the core-shell structure, site-selected deposition and thus formation of
an incomplete shell is highly desired for catalytic applications.[123] In 2012, our group reported the
synthesis of Pd-Rh bimetallic nanocubes with a “core-frame” structure and concave faces through
confined deposition of Rh atoms onto the corners and edges of Pd cubic seeds, as shown in Figure
11A.[124] Through a combination of selective capping of Pd{100} facets by Br− and kinetic control,
site-selective overgrowth could be readily achieved (Figure 11, B and C). This strategy has been
successfully applied to achieve site-selected deposition of Pt and Ru on Pd cubic seeds in recent
years, as well as to the synthesis of star-like Au-Cu nanocrystals referred to as pentacles.[108,125,126]
It is worth noting that site-selected growth also allows one to fabricate a “nanoframe” with a highly
open structure when the deposited metal was more resistant to oxidative etching than the core
metal.[124,125,127-129] As shown in Figure 11, D and E, the Pd-Rh core-frame nanocubes could be
readily converted to Rh cubic nanoframes with ridges as thin as 2 nm by selectively etching away
the Pd cores.[124,127-129] Using a similar approach, Rh octahedral, Au triangular, and Au decahedral
nanoframes were also fabricated (Figure 11F-H).[127-129] We believe site-selected growth enabled
by surface capping can be combined with post-synthesis etching to generate nanoframes with
tunable compositions, morphologies, and properties for a variety of applications.

5. Computational Investigation
It is still difficult to gather information about the binding energy or strength of a capping
agent, as well its adsorption/desorption dynamics, through experimental analyses. Alternatively,
DFT calculations and other advanced theoretical studies can offer insightful and quantitative trends
with regard to surface capping.[90,130] It has been shown that the efficacy of surface capping is
dependent on the structural conformations and chemical functional groups of the capping ligand,
17

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

as well as the crystallographic plane and electronic structures of the metal.[130] Here we focus on
the studies that use computational methods to understand the interaction between a capping agent
and a metal surface in terms of facet selectivity and shape evolution.

5.1. PVP on Ag(100)


For the synthesis of Ag nanocrystals encased by {100} facets, PVP with different molecular

Accepted Manuscript
weights are commonly used as a capping agent. To resolve the role of PVP in the synthesis while
understanding the differences between small molecules and macromolecules, DFT calculation was
conducted with an aim to elucidate the complex interactions between PVP and Ag surface.[131] The
repeating unit of PVP was represented by two-component sub-units: ethane and 2-pyrrolidone (2P).
Because of the inert and nonpolar character of ethane, it was expected that the interaction between
ethane and Ag surface would be dominated by vdW attraction. Different from ethane, the oxygen
and possibly nitrogen atoms in the 2P ring could bind to an Ag surface through direct chemical
bond formation. The results suggest that, in the case of macromolecular capping agent, both vdW
and covalent bonding should be considered when analyzing the binding configurations of PVP on
Ag and calculating the total binding energy, different from the case of small capping molecules.
When the authors simulated the binding conformations of ethane and 2P on Ag(100) and
Ag(111) surfaces with no vdW interactions, there was a slight preference for 2P to bind to Ag(111),
inconsistent with experimental observations. In contrast, when the vdW forces were included in
the binding interactions between PVP and Ag, the angles between the 2P rings and the Ag surface
were in the ranges of 10−20o, indicating that the rings of PVP were slightly tilted relative to the
surface when binding to (100) and (111) planes. The results suggest that the vdW interactions
between PVP and Ag can dramatically affect and “flatten” the binding conformations on the
surface. To understand the facet selectivity of PVP binding, the authors divided the total binding
energy (Ebind) into three components: long-range vdW attraction (EvdW), short-range direct bonding
(Edirect), and Pauli repulsion (EPauli, i.e., the repulsive interaction between the electrons of metal
atoms and the adsorbed molecules):

𝐸bind = 𝐸vdW + 𝐸direct + 𝐸Pauli (Eq. 2)

The results indicate that the total binding energy of PVP on Ag(100) is larger than that on
18

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Ag(111), along with a significant contribution from the long-range vdW interactions. Comparing
the binding energies of PVP on Ag(100) and Ag(111), there was 80 meV preference for PVP to
bind to Ag(100). Importantly, the binding energy (Ebind) derived from the simulation can be used
to calculate the density of PVP on a given surface. Based on the Boltzmann-Gibbs statistics at a
finite temperature T, the ratio of the densities on Ag(100) and Ag(111) can be expressed as:[131]

Accepted Manuscript
(100)
𝑁(100) Exp[𝐸bind /𝑘𝐵 𝑇]
= (111) (Eq. 3)
𝑁(111) Exp[𝐸bind /𝑘𝐵 𝑇]

where kB is the Boltzmann constant. The ca. 80 meV difference in binding energy on these two Ag
planes leads to a 10 times higher density of PVP on Ag(100) than on Ag(111), consistent with
experimental observation. Take together, it is necessary to consider the surface-sensitive balance
between direct binding and vdW interactions, which determines the energetic preference of
capping towards a particular facet, especially in the case of macromolecules.

5.2. Bromide on Pd(100)


Small inorganic species such as halides (Cl–, Br–, and I–) can be advantageous over the
commonly used macromolecular capping agents whose long chains may induce a steric hindrance
effect. Halide ions tend to form strong covalent or coordination bonds with the metal atoms on the
surface. Importantly, they can form a close-packed adlayer structure on the metal surface because
of their simple monatomic structure. In an adlayer, two competing interactions (halide−metal and
halide−halide) should be taken into account when discussing the role of halides in shape-controlled
synthesis of nanocrystals. One of the notable cases is the selective adsorption of Br– on Pd{100}
facets, leading to the formation of nanocubes. Recently, an ab initio study based on DFT simulation
was conducted to examine the energetics involved in the morphological transformation of Pd
nanocrystals under different surface coverages of Br atoms.[132] They calculated the binding
energies of Br atoms at the most stable sites on various types of Pd surfaces (including low-and
high-index planes) as a function of Br coverage, as shown in Figure 12A. At a coverage of 0.25
monolayer (ML), the binding energy (-2.33 eV) of Br to the (100) plane was the lowest among all
crystallographic planes. With the increase of Br coverage beyond 0.25 ML, the binding energies
on all planes quickly became less favorable and Br atom bond most strongly to the most open
19

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Pd(110). This trend can be attributed to the repulsive, adsorbate-adsorbate interaction on the more
closely packed Pd (100) surface.
The calculation results are, however, not consistent with the experimental observations. This
could be due to i) the adsorbate they chose was neutral Br atoms, rather than Br– ions and ii) the
simulation environment is vacuum instead of aqueous solution. Both of these factors deviate from
the conditions of a real synthesis. It still requires additional efforts to achieve a good understanding

Accepted Manuscript
of the interaction between the halide ions and a clean metal surface.

5.3. CA on Ag(111)
CA, containing one hydroxyl and three carboxyl groups, is widely used as a capping agent
to block {111} facets while allowing {100} facets to grow for Ag, Au, and Pd nanocrystals. To
understand the origin of this facet selectivity, ab initio DFT calculation was applied to establish
the binding model and evaluate the binding energy of CA on Ag surface.[133] As an fcc metal, the
(100) plane of Ag has square-packing symmetry, while the (111) plane shows hexagonal symmetry.
On the other hand, the approximate symmetry of CA molecule has three- and six-fold rotation axes
that interchange the three carboxyl groups, as well as the reflection through a plane that relates the
methylene-carboxyl groups. The authors hypothesized that the optimal binding structures of CA
on Ag(100) and Ag(111) were determined by their matching in geometry, as shown in Figure 12B.
The binding of CA to these two Ag surfaces involves different numbers of surface-ligand bonds
(i.e., Ag-O bonds). For the CA on square-packed (100) surface, the mismatch between the surface
and molecule and the interatomic distances only result in two Ag-O bonds with the (100) surface.
On the other hand, the hexagonal symmetry of (111) surface matches the approximate three-fold
symmetry of CA, and thus four Ag-O bonds are formed. The four bonds between CA and Ag(111)
induces a large binding energy of 13.84 kcal/mol, which is significantly greater than that of Ag(100)
(3.69 kcal/mol). According to Boltzmann distribution (Eq. 3), this large difference results in a
higher density of CA on Ag(111) than on Ag(100), explaining why CA is able to retard the growth
of Ag(111).[133] This study offers a quantitative description of the facet selectivity of CA towards
Ag{111}. A similar trend was also observed for CA and Pd nanocrystals.[134]

5.4. CO on Pt(100)
The triple bond (C≡O) in CO is made of two normal covalent bonds and one dative covalent
20

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

bond in which the two electrons come from O. As such, the C atom carries a negative fractional
charge although it is less electronegative than O atom. Owing to this electronic feature and thus a
dipole moment, the CO molecules chemisorbed on a metal tend to be oriented perpendicular to the
surface, with the C atom pointing down. This feature has been proven by both theoretical and
experimental studies. In addition, the adsorption energy of CO on a metal surface strongly depends
on the crystallographic plane and electronic structure of the metal, as well as the repulsion among

Accepted Manuscript
the adsorbed molecules. Based on DFT study, the adsorption energy of CO at the most stable “top”
site on a Pt(100) plane (1.95 eV) is larger than that of on a (111) plane (1.72 eV), suggesting the
preferential binding of CO towards Pt(100) surface.[135]
It is worth pointing out that the preferable adsorption structure for CO on a surface depends
on its coverage.[136,137] For example, with the increase of CO coverage from 0.25 to 0.5 ML, the
most stable structure of CO on Pt(100) switches from the top to the bridge configuration.[137] In
another study, when the coverage of CO increased from 0.06 to 1.0 ML on Pt(111), four stable
ordered structures, including (√3×√3)R30°–CO, c(4×2)–4CO, c(√3×5)rect–6CO, and c(√3×

3)rect–4CO, with various bridge to top ratios were identified.[136] Interestingly, the trend in the
bridge to top ratio first increased and then decreased with the increase in coverage, which can be
ascribed to the change in Pt surface charge upon CO adsorption at the top and bridge sites. These
predictions indicate that the site preference and the stable structure for CO on various types of Pt
surfaces at low, intermediate, and high coverages are totally different, which is expected to create
different surface free energies for different Pt facets.

5.5. Peptides S7 on Pt(111)


The discovery of facet-specific biomolecules (e.g., peptides) has enriched the pool of capping
agents for the shape-controlled synthesis of metal nanocrystals. As shown in Figure 12, C and D,
experimental and simulation studies have been integrated to understand the preferential and non-
preferential adsorptions on Pt{111} when the compositions and sequences of peptide S7 were
altered (Section 4.5).[88] The authors designed a set of S7 variants to identify the key factor in
controlling peptide capping. The variables they investigated including the sequence of peptide S7,
the peptide conformation, the individual functional motifs, and the detailed molecular effect of the
phenyl ring in the amino acid. The experimental results suggest amino acid F with phenyl ring
(−C6H5) plays the key role in inducing the formation of Pt{111}.
21

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Surprisingly, based on the computational results, all peptides have large negative adsorption
energies (Eads) on Pt(111) surface, suggesting effective binding interactions between them but there
was no direct correlation between the yield of tetrahedra and Eads(111) value, as shown in Figure
12D.[88] On the other hand, the adsorption energies of the peptides on Pt(100) surface strongly
correlates with the yield of tetrahedra. The F-containing (S7, S7-1, S7-2, SSF, and FPQPN) and
F-free (S7-G, PQPN, S7-Y, and SSY) peptides gave positive and negative Eads(100), respectively.

Accepted Manuscript
Thus, it was concluded that the binding contrast between (111) and (100) surfaces was the origin
of selectivity. Based on the molecular simulations, the “lie-flat” configuration was observed for F-
containing peptides on Pt(111) due to the hexagonal epitaxial coordination between the phenyl
group and the surface, which contributed to stronger attraction among the capping agents (Figure
12C). In contrast, because of the mismatch between the phenyl ring and the square packing of (100)
surface, the “stand-up” configuration dominated the binding landscape of F-containing peptides
on (100) surface. Such configuration creates much less steric hindrance, reducing the overall
peptide binding affinity towards Pt(100) when compared to the lie-flat configuration. These results
reveal that the phenyl group in a peptide is responsible for the high binding contrast between Pt(111)
and Pt(100), suggesting that the binding contrast enabled by the conformation of a capping agent
is critical to facet specificity.
Taken together, computation provides a powerful means to reveal the role played by a
capping agent in carving out the shape of a nanocrystal at an atomic/molecular level, as well as to
identify their binding selectivity towards various types of surfaces. Although the computational
studies have reached some agreement with the experimental observations made in the shape-
controlled synthesis of metal nanocrystals, they all involved assumptions with regard to the
experimental conditions.[130] These assumptions may produce biased simulation results. As such,
there is still an urgent need to develop computational methods that take into account the real
conditions of a synthesis to examine the roles of various experimental parameters.

6. Characterization Tools
Despite their successful use in shape-controlled synthesis, a quantitative measure of most of
the capping agents at the nanocrystal/solution interface is still elusive. As a result, the explicit roles
played by most capping agents are still under debate. In addressing issue, more and more analytical
tools are being applied to quantify the adsorption of capping agent. For instance, zeta potential[138]
22

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

measurements can indicate the overall charge of the particle, giving information on the presence
of positively/negatively charged capping agent adsorbed at the interface. Ultraviolet-visible (UV-
vis) spectroscopy[139] can provide information about the electronic structure of the nanocrystals
and the chemical environment at the interface, both of which can change depending on the capping
agent. Fourier transform infrared spectroscopy (FTIR)[138,140] can be used to track the characteristic
absorption peaks of different functional groups associated with a specific capping agent. To gain

Accepted Manuscript
more knowledge on the distribution of elements appeared in the capping agent, energy dispersive
X-ray spectroscopy (EDX)[47] has been implemented, while X-ray photoelectron spectroscopy
(XPS)[141-143] has been used to investigate the presence of certain element, as well as its chemical
state. Also, nuclear magnetic resonance spectroscopy (NMR)[144] has been used to study the
binding mode of a capping agent on the metal surface. Thermogravimetric analysis (TGA)[145,146]
can detect the presence of a capping agent based on its thermostability. The high sensitivity of
inductively coupled plasma mass spectrometry(ICP-MS)[141,147] makes it a powerful technique to
quantify even a trace amount of a capping agent adsorbed on the surface of nanocrystals. Similarly,
high performance liquid chromatography (HPLC)[54] has recently been employed to derive the
adsorption equilibrium constant of a capping agent and determine the facet-binding preference.
Finally, isothermal titration calorimetry (ITC)[148] has been used to investigate the thermodynamics
involved in the binding of an adsorbent to the surface of nanoparticles. Figure 13 shows a brief
summary of the characterization techniques mentioned above, and more detailed discussion can
be found in the Supplementary Information.
Quantification of surface capping can also be accomplished by simply monitoring the shape
of nanocrystals during growth in the presence of a certain concentration of capping agent in the
reaction solution. As shown in Figure 14, this approach was used to quantitatively analyze the
coverage density of PVP on Ag nanocubes.[149] With Ag nanocubes enclosed by {100} facets
serving as seeds for growth, two sets of experiments were carried out: i) with a fixed initial PVP
concentration (C1) to determine the critical size at which Ag{111} facets started to appear; and ii)
with the initial concentration of PVP reduced to a critical point (C2) so Ag{111} facets started to
appear at the very beginning of seed-mediated growth. The coverage density (ϕ) of PVP on Ag{100}
can be derived from the difference in concentration, C1 − C2, and total surface area of Ag cubes
between samples 3 and 5 in Figure 14A using the following equation:

23

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

𝜙 = (𝐶1 − 𝐶2 ) ∙ 𝑉 ∙ 𝑁𝐴 /∆𝑆𝐴𝑔 (Eq. 4)

where V is the total volume of the solution, NA stands for the Avogadro’s number, and ΔSAg is the
increase in surface area between samples 3 and 5. It was found that the coverage densities were
140 and 30 repeating units per nm2 for PVP of 55000 and 10000 g/mol, respectively, when 40 nm
Ag cubes were used as the seeds. The values dropped slightly to 100 and 20 repeating units per

Accepted Manuscript
nm2, respectively, for cubic seeds of 100 nm in size. These results suggested that PVP with a
relatively lower molecular weight was more effective in reducing the free energy of Ag(100) via
surface adsorption than PVP with a higher molecular weight. On a quantitative basis, this approach
is able to determine the coverage density of a capping agent and thus elucidate its effect on the
shape evolution and growth habit of metal nanocrystals.

7. Nanocrystals Free of Capping Agent


The capping agent remaining on the surface of metal nanocrystals after a synthesis might
compromise their performance in heterogeneous catalysis and other applications. For example, it
was reported that the electrochemically active surface area of the Pt-Ni octahedral catalyst without
PVP capping layer was much greater than that of the same catalyst but with PVP. [150] This issue
can be attributed to the blocking effect of PVP in restricting the free access of reactants to catalytic
sites on a Pt-Ni octahedral nanocrystal. As such, nanocrystals free of capping agent are generally
more favored in heterogeneous catalysis. As discussed in Section 2 using the Langmuir isotherm,
the adsorption of a capping agent is determined by both its concentration in the solution and the
equilibrium constant K. In principle, repeated washing of the nanocrystals with a good solvent for
the capping agent should be able to remove all the adsorbed molecules. However, for a capping
agents that bind to the surface with strong affinity, the desorption kinetics could be very slow,
making the washing process highly inefficient. In this section, we summarize recent progress in
the development of metal nanocrystals with a clean surface by either removing the capping agent
after the synthesis or conducting a synthesis without involving any capping agent.

7.1 Post-synthesis Removal of Capping Agent


The interaction between a capping agent and a metal surface can involve either physical or
chemical adsorption.[130,151-154] For the physisorbed capping agent, it can be removed from the
24

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

surface through simple washing with a proper solvent. It is more difficult to remove chemisorbed
capping agent due to the involvement of chemical bonding. Although the chemisorbed capping
agent can be removed through post-synthesis treatment such as thermal annealing, it often induces
changes to the shape, in addition to irreversible aggregation or fusion of the nanocrystals.[151] Here
we focus on the strategies for removing the capping agent without altering the original shape.
Halides are commonly used as capping agents for the synthesis of metal nanocubes, nanobars,

Accepted Manuscript
and penta-twinned rods/wires mainly enclosed by {100} facets since they preferentially adsorb on
(100) surface. It was shown that Br− could be removed from Pd surface via desorption at 100 oC
under a reductive environment.[141] Due to the involvement of chemisorption, the Pd atoms on the
surface should be partially oxidized. If the oxidized atoms are reduced to the zero oxidation state,
the adsorbed Br− ions can be easily removed according to the following half-reaction:

Pd(II) − Br − + 2e− → Pd(0) + Br − (Eq. 5)

This concept was proven by conducting a set of experiments in which aqueous or EG suspensions
containing CA (reducing agent), PVP (colloidal stabilizer), and Br−-capped Pd nanocubes were
aged at different temperatures. As shown in Figure 15A, the Pd 3d XPS spectrum of Br−-capped
Pd nanocubes displayed two peaks at 340.4 and 335.1 eV, corresponding to elemental Pd 3d3/2 and
Pd 3d5/2, respectively, while shoulders at higher energies were also observed, implying the
existence of Pd(II) on the surface. The shoulders gradually decreased in intensity and disappeared
for the EG sample aged at 100 °C, revealing the reduction of Pd(II) to Pd(0) and thus the desorption
of Br− ions. This result was also consistent with the quantitative analysis based on ICP-MS (Figure
15B). The amount of Br− ions dropped to 12% of the original value for the EG sample aged at 100
°C. Altogether, the chemisorbed Br− ions can be removed without changing the shape of the Pd
nanocrystals in the presence of a mild reducing agent.[141]
Other methods such as UVO (UV-ozone) irradiation have also been adopted to remove the
capping agent on nanocrystals.[155-157] Unlike thermal treatment, UVO irradiation at room
temperature can yield clean surface while preserving the shape of the nanocrystals. It involves the
simultaneous action of UV light and ozone, which are responsible for the oxidation of an organic
capping agent into CO2 and H2O. In one study, UVO treatment was used to remove organic
capping agents such as tetradecyl tributylammonium bromid (TTAB) from Pt nanocubes.[156]
25

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Figure 15C shows FTIR spectra of the TTAB-coated Pt nanocubes as a function of time for UVO
treatment, revealing the complete disappearance of the alkyl chain, while the quaternary
ammonium moiety remained after 120 min of treatment. In another case, PVP-capped Pd
nanocubes were subjected to UVO cleaning, as shown in Figure 15D.[155] The removal of PVP was
analyzed by ATR-FTIR spectra. After 3 h of treatment, PVP was removed and only a few bands
(1900 and 1650 cm-1) were still visible. The broad band at 1650 cm-1 can be assigned to the H2O

Accepted Manuscript
bending mode while the two new bands at 1930 and 2045 cm-1 suggested that a small amount of
CO was formed during the decomposition of PVP.
For some applications, post-synthesis ligand exchange is used to replace the original capping
agent with a new one having the desired functional group. For example, it is necessary to replace
the toxic CTAB or CTAC on Au nanocrystals with biocompatible ligands for biomedical
applications. To this end, an indirect method was reported for effectively replacing the ligand on
Au nanocrystals, as shown in Figure 15E.[138] In this method, an ultrathin layer (ca. 1.5 nm) of Ag
was deposited on Au nanocrystals to help remove the original, toxic capping agent such as CTAC,
followed by etching of the Ag layer in the presence of the second ligand such as biocompatible
citrate ions. This strategy offers a facile and robust means to completely replace a strong ligand
such as CTAC with an oppositely charged, weak ligand such as citrate, without inducing changes
to the shape of the nanocrystals.

7.2. Synthesis without Capping Agent


Compared to the conventional synthesis that needs additional cleaning procedures to remove
the capping agent, synthesis conducted in the absence of any capping agent is more straightforward
in obtaining nanocrystals with a clean surface. Without the stabilization by a capping agent (or a
colloidal stabilizer), it will be difficult to keep the nanocrystals dispersed in the reaction medium.
This issue can be addressed by directly depositing the nanocrystals on a solid support. [150,158,159]
The control of shape is typically achieved through the use of a template.[24] For example, a simple
method was reported for the preparation of Pd@Pt3–4L core-shell nanocrystals with well-defined
{111} facets, as shown in Figure 16A.[159] In the first step, 21-nm Pd octahedral seeds were
synthesized and then loaded onto carbon black for use as a template. Upon optimization of reaction
conditions, ultrathin shells of Pt could be conformally deposited on the octahedral seeds in a layer-
by-layer fashion without involving self-nucleation (Figure 16, B and C). XPS analysis was used
26

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

to examine the surface species for the samples prepared in the absence and presence of PVP (a
stabilizer in this case), as shown in Figure 16D. Even after the sample had been washed eight times
with water, the sample prepared in the presence of PVP still exhibited a noticeable N1s peak at ca.
398 eV associated with PVP. In contrast, the N1s signal of PVP was negligible in the XPS
spectrum for the sample prepared in the absence of PVP, implying that the sample was free of
PVP.[159] It is worth noting that some synthetic systems involve the use of organic reducing agents

Accepted Manuscript
such as L-ascorbic acid, formaldehyde, and hydrazine, which could be oxidized to generate CO
during a synthesis.[160] The CO would strongly adsorb onto the surface of noble metals, blocking
the active sites and thus negatively impacting the catalytic activity.
In addition to wet chemical synthesis carried out in the presence of solid support, physical
vapor deposition (PVD) has been explored to produce metal nanocrystals free of capping agent or
colloidal stabilizer.[161-164] As demonstrated in a series of studies, both monometallic and bimetallic
nanocrystals made of Au, Ag, Pd, and Pt were fabricated on amorphous SiO2 or (0001)-oriented
sapphire substrates through the PVD-based method. In brief, a thin metal film was first deposited
on the substrate by sputtering or evaporation techniques, followed by the formation of nanocrystals
with desired twin structures through careful heating and then cooling. In one example, Ag films
with different thicknesses were deposited on SiO2 and then subjected to heating/cooling to obtain
capping-free Ag nanocrystals with icosahedral, decahedral, or single-crystal structures under a
thermodynamic control, as shown in Figure 16, E-J.[162] Another study showed that a periodic array
of PVD-derived Au nanocrystals on (0001)-sapphire could serve as seeds for the kinetically
controlled growth of Ag to generate either Au-Ag nanocrystals with either a Janus or a core-shell
structure.[163] Capping agent-free synthesis could also be achieved in gas phase by incorporating a
trace amount of foreign metal powders as the shape-directing agent, followed by evaporative
dealloying.[165] A recent report shows that tetrahexahedral (THH) nanocrystals with a range of
sizes and compositions could be synthesized using this method. The solid metal precursor or metal
particles, together with a foreign metal (Sb, Te, Pb, or Bi) were heated togher under Ar/H2
atmosphere. The {210} facets exposed on the surface of THH nanocrystals could be stabilized in
the presence of the foreign metal atoms before they were removed by dealloying. These
demonstrations suggest that gas-phase synthesis offers an attractive platform for the preparation
of metal nanocrystals with a clean surface.

27

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

8. Concluding Remarks
Maneuvering the shape of metal nanocrystals and thus the types of facets exposed on their
surface offers a powerful means for tailoring their physicochemical properties to realize various
applications. Under the assistance of capping agents, significant progress has been made in recent
years with regard to the colloidal synthesis of metal nanocrystals with controlled shapes. It is worth
pointing out that, in addition to their role in directing the shape evolution of nanocrystals, the

Accepted Manuscript
capping ligands can also serve as a colloidal stabilizer to prevent the suspended nanocrystals from
aggregation during synthesis and/or storage.[166] Despite the incredible accomplishments, many
scientific problems or technological challenges still need to be addressed before we are able to use
the as-obtained nanocrystals for a broad spectrum of applications.
One of the challenges is a quantitative analysis and understanding of the roles played by a
capping agent in a typical synthesis of metal nanocrystals in the context of reduction, deposition,
and diffusion. In the seed-mediated growth, it was shown that the deposition rate of atoms relative
to the surface diffusion rate of adatoms dominates the evolution of a seed into nanocrystals with
diverse shapes.[108] However, we still have no quantitative measure of how a capping agent and its
coverage density would affect these kinetics processes due to the lack of experimental tools. For
the deposition of atoms from precursor reduction in the solution phase, they will be prohibited if
the atoms strike a region on the seed capped and thus passivated by a capping agent. However, if
the coverage density of the capping agent is not high enough to passivate the entire surface, atom
deposition can still occur on the spared regions to evolve into specific shapes depending on the
coverage density.[54,141,149] In addition, recent studies also suggested that the precursor can first
adsorb onto the surface of a growing nanocrystal, followed by chemical reduction to atoms.[167-169]
In this case, the interaction between the precursor and the capping agent on the surface will play a
key role in determining the reduction and thus deposition processes. For instance, the deposition
is expected to preferentially take place on the capped region if there is an attractive interaction
between the precursor and capping agent, giving a growth pattern complementary to the one arising
from surface passivation by the capping agent. On the other hand, the adsorbed capping agent is
also believed to play a crucial role in influencing surface diffusion of adatom on the seed’s
surface.[170] The capping layer with a specific binding strength would present as a kinetic barrier,
weakening the ability of adatoms to diffuse to more energetically favored sites. Taken together, it
is highly desired to have a quantitative measure and understanding of the capping agent involved
28

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

in a synthesis. This can be achieved through the development of in situ tracking methods such as
synchrotron X-ray diffraction[171,172] and surface-enhanced Raman scattering (SERS) while
including computational modeling with an emphasis on the dynamic aspects.[173]
Another major challenge is how to screen and identify new capping agents for the syntheses
of nanocrystals covered by various types of facets, in particular, the high-index ones such as {221},
{310}, {411}, and {720}, that contain high densities of low-coordination surface atoms. As shown

Accepted Manuscript
in a one-pot synthesis, CTAC molecules could serve as a capping agent in the formation of Au
trisoctahedrons enclosed by {221} facets because the long alkyl (CTA+) chains have a size
comparable to the atomic spacing on the high-index facets and can thus adsorb onto the less closely
packed, high-index facets rather than the low-index counterparts.[67] In another study, concave Pt
nanocrystals with high-index {411} facets were prepared in the presence of amine compounds as
the capping agent.[93] Overall, these capping agents were discovered randomly. Before rational
design becomes available, it should be more productive by systematically screening a library of
compounds for their selectivity towards various combinations of metals and facets. This can be
achieved by conducting a synthesis with the pre-synthesized nanocrystals as seeds.[88,120] Upon the
introduction of additional metal precursor and reducing agent, together with a specific compound
of interest, the growth can be initiated and continued at an elevated temperature to facilitate surface
diffusion for the achievement of thermodynamic control. By monitoring the shape evolution of the
seeds either ex situ or in situ, one should be able to elucidate the capping ability and facet
selectivity of the compound. This rapid and simple approach can be easily extended to a broad
range of compounds to evaluate their potential for the preparation of a variety of metal nanocrystals
bound by diverse facets.
As for the application in heterogeneous catalysis, it is generally accepted that the capping
molecules on a metal nanocrystal tend to block the active sites and compromise its catalytic activity.
However, several groups demonstrated an unexpected benefit of capping agent in serving as an
activity promoter and/or selectivity modifier towards a specific catalytic reaction.[153,174,175] This
benefit can be attributed to the creation of a metal-ligand interface, through which geometric (steric)
and electronic factors can be leveraged to generate a favorable change to the local environment.
The influence of these factors on the catalytic reaction was found to be dependent on the
composition, coverage density, spatial arrangement, and interaction strength between the capping
agent and other chemical species such as reactants and solvents. In one study, it was reported that
29

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

the Au nanocrystals stabilized by PVP at a proper coverage density showed higher activity towards
aerobic oxidation of alcohol than that of Au nanocrystals stabilized by poly(allylamine).[174] As
illustrated in Figure 17A, the Au nanocrystals were negatively charged due to the donation of
electrons from PVP to Au and thus alteration to the electronic structure. During the oxidation
catalysis, the electrons flew from the anionic Au on PVP-capped Au nanocrystals into the LUMO
(π*) of O2, generating superoxo (O2−)- or peroxo (O22−)-like species. As a result, the catalytic

Accepted Manuscript
activity of the PVP-capped Au nanocrystals was significantly enhanced. In another study, UVO-
treatment (Section 7.1) was used to control the degree of removal and thus coverage density of
PVP on the surface of Au nanocrystals to assess the impact of PVP on the hydrogenation of p-
chloronitrobenzene (p-CNB) and cinnamaldehyde (CAL), as shown in Figure 17B.[175] The
presence of PVP was found to be beneficial to the activity of Au towards the hydrogenation of p-
CNB while compromising the hydrogenation of CAL. In the case of p-CNB, p-chloroaniline (p-
CAN) was the sole product, and the corresponding turnover frequency (TOFp-CNB, the number of
moles of p-CNB transformed into p-CAN by one mole of active sites per hour) decreased rapidly
from 178 to 100 h−1 within the first hour of UVO-treatment. This result was attributed to the
difference in orientation for the activated complex that interacted differently with the metal surface
depending on the presence or absence of the capping agent. The benzene ring of p-CNB in the
activated complex on the PVP-capped Au nanocrystals was oriented perpendicular to the surface,
and its interaction with the catalyst would be remarkably weaker than its flat-on orientation parallel
to the surface of the clean Au nanocrystals. The mobility of the activated complex would be
facilitated by a weakened interaction of the activated complex with the Au in the presence of PVP,
leading to a higher activity towards the hydrogenation of p-CNB. In contrast, the significantly
lower TOFCAL of the as-prepared, PVP-capped Au catalyst than that of the UVO-cleaned Au-
catalyst indicated an adverse impact of PVP on the catalytic reaction, which can be attributed to
the partial blockage of catalytic sites by the organic residues, either through direct interaction with
or steric hindering of active sites from being accessed by the reactant. Taken together, these
experimental results demonstrated that the capping agent could serve as a “poison” to limit the
accessibility of active sites, or a “promoter” to improve the activity and selectivity. Owing to the
advantage of diverse capping agents with different functionalities, there is a lot of room for further
development in this field, where the synergetic effect between metal nanocrystals and capping
agents can be used to manipulate the activity and selectivity. To this end, it is an urgent challenge
30

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

to integrate ex situ and in situ characterization techniques for probing catalytic processes at metal-
ligand interface under reaction conditions to distinguish the electronic and geometric changes
induced by capping molecules on the catalytic performances of metal nanocrystals.
When these challenges are fully addressed, it will become possible to rationally synthesize
and engineer nanocrystals with desired properties by choosing the suitable capping agents together
with other experimental parameters. It is hoped that the concepts and case studies presented in this

Accepted Manuscript
article will not only assist researchers in understanding the roles played by surface capping but
also offer the insights and guidelines for designing their own syntheses.

Acknowledgements
This work was supported in part by research grants from the NSF (DMR-1505400, CHE-1505441,
and CHE-1804970) and startup funds from the Georgia Institute of Technology.

References
[1] Y. Xia, Y. Xiong, B. Lim, S. E. Skrabalak, Angew. Chem. Int. Ed. 2009, 48, 60.
[2] C. Burda, X. Chen, R. Narayanan, M. A. El-Sayed, Chem. Rev. 2005, 105, 1025.
[3] L. M. Liz-Marzán, Mater. Today 2004, 7, 26.
[4] C. J. Murphy, T. K. Sau, A. M. Gole, C. J. Orendorff, J. Gao, L. Gou, S. E. Hunyadi, T.
Li, J. Phys. Chem. B 2005, 109, 13857.
[5] Y. Xia, K. D. Gilroy, H.-C. Peng, X. Xia, Angew. Chem. Int. Ed. 2017, 56, 60.
[6] K. D. Gilroy, A. Ruditskiy, H. C. Peng, D. Qin, Y. Xia, Chem. Rev. 2016, 116, 10414.
[7] S. Lal, S. Link, N. J. Halas, Nat. Photonics 2007, 1, 641.
[8] S. Ye, A. R. Rathmell, Z. Chen, I. E. Stewart, B. J. Wiley, Adv. Mater. 2014, 26, 6670.
[9] A. Ruditskiy, H. C. Peng, Y. Xia, Annu. Rev. Chem. Biomol. Eng. 2016, 7, 327.
[10] X. Huang, Z. Zeng, S. Bao, M. Wang, X. Qi, Z. Fan, H. Zhang, Nat. Commun. 2013, 4,
1444.
[11] Z. Zhao, C. Chen, Z. Liu, J. Huang, M. Wu, H. Liu, Y. Li, Y. Huang, Adv. Mater. 2019,
31, 1808115.
[12] Z.-Y. Zhou, N. Tian, J.-T. Li, I. Broadwell, S.-G. Sun, Chem. Soc. Rev. 2011, 40, 4167.
[13] G. Konstantatos, E. H. Sargent, Nat. Nanotechnol. 2010, 5, 391.
[14] K. A. Willets, R. P. Van Duyne, Annu. Rev. Phys. Chem. 2007, 58, 267.
[15] K. M. Mayer, J. H. Hafner, Chem. Rev. 2011, 111, 3828.
[16] Y. C. Cao, R. Jin, C. A. Mirkin, Science 2002, 297, 1536.
[17] E. C. Dreaden, A. M. Alkilany, X. Huang, C. J. Murphy, M. A. El-Sayed, Chem. Soc.
Rev. 2012, 41, 2740.
[18] X. Yang, M. Yang, B. Pang, M. Vara, Y. Xia, Chem. Rev. 2015, 115, 10410.
[19] B. Sepúlveda, P. C. Angelomé, L. M. Lechuga, L. M. Liz-Marzán, Nano Today 2009, 4,
244.
[20] Y. Zheng, X. Zhong, Z. Li, Y. Xia, Part. Part. Syst. Charact. 2014, 31, 266.
31

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

[21] L. Scarabelli, M. Coronado-Puchau, J. J. Giner-Casares, J. Langer, L. M. Liz-Marzán,


ACS Nano 2014, 8, 5833.
[22] Y. Xia, W. Li, C. M. Cobley, J. Chen, X. Xia, Q. Zhang, M. Yang, E. C. Cho, P. K.
Brown, Acc. Chem. Res. 2011, 44, 914.
[23] Z. Fan, H. Zhang, Acc. Chem. Res. 2016, 49, 2841.
[24] K. D. Gilroy, X. Yang, S. Xie, M. Zhao, D. Qin, Y. Xia, Adv. Mater. 2018, 30, 1706312.
[25] J. Qian, M. Shen, S. Zhou, C.-T. Lee, M. Zhao, Z. Lyu, Z. D. Hood, M. Vara, K. D.
Gilroy, K. Wang, Y. Xia, Mater. Today 2018, 21, 834.
[26] Z. Wang, G. Yang, Z. Zhang, M. Jin, Y. Yin, ACS Nano 2016, 10, 4559.

Accepted Manuscript
[27] X. Zhang, H. Yin, J. Wang, L. Chang, Y. Gao, W. Liu, Z. Tang, Nanoscale 2013, 5,
8392.
[28] K. M. Bratlie, H. Lee, K. Komvopoulos, P. Yang, G. A. Somorjai, Nano Lett. 2007, 7,
3097.
[29] P. Christopher, S. Linic, J. Am. Chem. Soc. 2008, 130, 11264.
[30] I. Lee, R. Morales, M. A. Albiter, F. Zaera, Proc. Natl. Acad. Sci. U. S. A. 2008, 105,
15241.
[31] B. Piqueras-Fiszman, Z. Laughlin, M. Miodownik, C. Spence, Food Qual. Prefer. 2012,
24, 24.
[32] P. Milan, S. Mordechay, Fundamentals of electrochemical deposition, John Wiley &
Sons, 2006.
[33] S. S. Kruglikov, N. T. Kudriavtsev, G. F. Vorobiova, A. Y. A. Antonov, Electrochim.
Acta 1965, 10, 253.
[34] M. Giulietti, M. M. Seckler, S. Derenzo, L. H. Schiavon, J. V. Valarelli, J. Nyvlt, Cryst.
Res. Technol. 1999, 34, 959.
[35] G. Wulff, Z. Kristallogr. 1901, 34, 449.
[36] G. D. Barmparis, Z. Lodziana, N. Lopez, I. N. Remediakis, Beilstein J. Nanotechnol.
2015, 6, 361.
[37] T. Wang, C. Lee, L. D. Schmidt, Surf. Sci. 1985, 163, 181.
[38] D. A. Jefferson, P. J. F. Harris, Nature 1988, 332, 617.
[39] T. S. Ahmadi, Z. L. Wang, T. C. Green, A. Henglein, M. A. El-Sayed, Science 1996, 272,
1924.
[40] Y. Sun, Y. Xia, Science 2002, 298, 2176.
[41] Y. Sun, B. Gates, B. Mayers, Y. Xia, Nano Lett. 2002, 2, 165.
[42] Y. Sun, Y. Xia, Adv. Mater. 2002, 14, 833.
[43] Y. Sun, Y. Yin, B. T. Mayers, T. Herricks, Y. Xia, Chem. Mater. 2002, 14, 4736.
[44] Y. Sun, B. Mayers, T. Herricks, Y. Xia, Nano Lett. 2003, 3, 955.
[45] D. F. Evans, H. Wennerström, The Colloidal Domain: Where Physics, Chemistry,
Biology, and Technology Meet, Wiley, 1999.
[46] S. Zhou, J. Li, K. D. Gilroy, J. Tao, C. Zhu, X. Yang, X. Sun, Y. Xia, ACS Nano 2016,
10, 9861.
[47] Z. Chen, T. Balankura, K. A. Fichthorn, R. M. Rioux, ACS Nano 2019, 13, 1849.
[48] B. J. Wiley, Y. Chen, J. M. McLellan, Y. Xiong, Z.-Y. Li, D. Ginger, Y. Xia, Nano Lett.
2007, 7, 1032.
[49] A. Ruditskiy, Y. Xia, J. Am. Chem. Soc. 2016, 138, 3161.
[50] R. R. da Silva, M. Yang, S. I. Choi, M. Chi, M. Luo, C. Zhang, Z. Y. Li, P. H. Camargo,
S. J. Ribeiro, Y. Xia, ACS Nano 2016, 10, 7892.
32

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

[51] F. Wu, W. Wang, Z. Xu, F. Li, Sci. Rep. 2015, 5, 10772.


[52] S.-W. Hsu, A. R. Tao, Chem. Mater. 2018, 30, 4617.
[53] Y. Wang, Y. Zheng, C. Z. Huang, Y. Xia, J. Am. Chem. Soc. 2013, 135, 1941.
[54] Z. Chen, J. W. Chang, C. Balasanthiran, S. T. Milner, R. M. Rioux, J. Am. Chem. Soc.
2019, 141, 4328.
[55] Y. Gao, P. Jiang, D. F. Liu, H. J. Yuan, X. Q. Yan, Z. P. Zhou, J. X. Wang, L. Song, L. F.
Liu, W. Y. Zhou, G. Wang, C. Y. Wang, S. S. Xie, J. M. Zhang, D. Y. Shen, J. Phys.
Chem. B 2004, 108, 12877.
[56] R. Jin, Y. Cao, C. A. Mirkin, K. L. Kelly, G. C. Schatz, J. G. Zheng, Science 2001, 294,

Accepted Manuscript
1901.
[57] G. S. Métraux, C. A. Mirkin, Adv. Mater. 2005, 17, 412.
[58] Q. Zhang, N. Li, J. Goebl, Z. Lu, Y. Yin, J. Am. Chem. Soc. 2011, 133, 18931.
[59] J. E. Millstone, S. Park, K. L. Shuford, L. Qin, G. C. Schatz, C. A. Mirkin, J. Am. Chem.
Soc. 2005, 127, 5312.
[60] W.-L. Huang, C.-H. Chen, M. H. Huang, J. Phys. Chem. C 2007, 111, 2533.
[61] T. H. Ha, H.-J. Koo, B. H. Chung, J. Phys. Chem. C 2007, 111, 1123.
[62] L. Chen, F. Ji, Y. Xu, L. He, Y. Mi, F. Bao, B. Sun, X. Zhang, Q. Zhang, Nano Lett.
2014, 14, 7201.
[63] J. Xie, J. Y. Lee, D. I. C. Wang, J. Phys. Chem. C 2007, 111, 10226.
[64] L. Au, B. Lim, P. Colletti, Y.-S. Jun, Y. Xia, Chem. Asian J. 2010, 5, 123.
[65] F. Kim, S. Connor, H. Song, T. Kuykendall, P. Yang, Angew. Chem. Int. Ed. 2004, 43,
3673.
[66] G. Wang, S. Tao, Y. Liu, L. Guo, G. Qin, K. Ijiro, M. Maeda, Y. Yin, Chem. Commun.
2016, 52, 398.
[67] Y. Ma, Q. Kuang, Z. Jiang, Z. Xie, R. Huang, L. Zheng, Angew. Chem. Int. Ed. 2008, 47,
8901.
[68] Y. Xiong, H. Cai, B. J. Wiley, J. Wang, M. J. Kim, Y. Xia, J. Am. Chem. Soc. 2007, 129,
3665.
[69] W. Niu, Z.-Y. Li, L. Shi, X. Liu, H. Li, S. Han, J. Chen, G. Xu, Cryst. Growth Des. 2008,
8, 4440.
[70] M. Jin, H. Liu, H. Zhang, Z. Xie, J. Liu, Y. Xia, Nano Res. 2011, 4, 83.
[71] Y.-L. Zhang, X.-L. Sui, L. Zhao, D.-M. Gu, G.-S. Huang, Z.-B. Wang, Int. J. Hydrog.
Energy 2019, 44, 6551.
[72] B. Lim, M. Jiang, J. Tao, P. H. C. Camargo, Y. Zhu, Y. Xia, Adv. Funct. Mater. 2009, 19,
189.
[73] H. Huang, A. Ruditskiy, S. I. Choi, L. Zhang, J. Liu, Z. Ye, Y. Xia, ACS Appl. Mater.
Interfaces 2017, 9, 31203.
[74] X. Xia, S.-I. Choi, J. A. Herron, N. Lu, J. Scaranto, H.-C. Peng, J. Wang, M. Mavrikakis,
M. J. Kim, Y. Xia, J. Am. Chem. Soc. 2013, 135, 15706.
[75] N. Lu, W. Chen, G. Fang, B. Chen, K. Yang, Y. Yang, Z. Wang, S. Huang, Y. Li, Chem.
Mater. 2014, 26, 2453.
[76] B. Lim, Y. Xiong, Y. Xia, Angew. Chem. Int. Ed. 2007, 46, 9279.
[77] L. Figueroa-Cosme, Z. D. Hood, K. D. Gilroy, Y. Xia, J. Mater. Chem. C 2018, 6, 4677.
[78] Z. Niu, Q. Peng, M. Gong, H. Rong, Y. Li, Angew. Chem. Int. Ed. 2011, 50, 6315.
[79] X. Huang, S. Tang, X. Mu, Y. Dai, G. Chen, Z. Zhou, F. Ruan, Z. Yang, N. Zheng, Nat.
Nanotechnol. 2011, 6, 28.
33

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

[80] Y. Dai, X. Mu, Y. Tan, K. Lin, Z. Yang, N. Zheng, G. Fu, J. Am. Chem. Soc. 2012, 134,
7073.
[81] A. Demortière, P. Launois, N. Goubet, P. A. Albouy, C. Petit, J. Phys. Chem. B 2008,
112, 14583.
[82] C. Kim, M. Min, Y. W. Chang, K. H. Yoo, H. Lee, J. Nanosci. Nanotechnol. 2010, 10,
233.
[83] Y. Zhang, M. E. Grass, J. N. Kuhn, F. Tao, S. E. Habas, W. Huang, P. Yang, G. A.
Somorjai, J. Am. Chem. Soc. 2008, 130, 5868.
[84] Y. Kang, X. Ye, C. B. Murray, Angew. Chem. Int. Ed. 2010, 49, 6156.

Accepted Manuscript
[85] Z. Peng, C. Kisielowski, A. T. Bell, Chem. Commun. 2012, 48, 1854.
[86] S.-I. Choi, S. Xie, M. Shao, J. H. Odell, N. Lu, H.-C. Peng, L. Protsailo, S. Guerrero, J.
Park, X. Xia, J. Wang, M. J. Kim, Y. Xia, Nano Lett. 2013, 13, 3420.
[87] I. A. Safo, M. Werheid, C. Dosche, M. Oezaslan, Nanoscale Adv. 2019, 1, 3095.
[88] L. Ruan, H. Ramezani-Dakhel, C. Y. Chiu, E. Zhu, Y. Li, H. Heinz, Y. Huang, Nano
Lett. 2013, 13, 840.
[89] C. Y. Chiu, Y. Li, L. Ruan, X. Ye, C. B. Murray, Y. Huang, Nat. Chem. 2011, 3, 393.
[90] H. Ramezani-Dakhel, L. Ruan, Y. Huang, H. Heinz, Adv. Funct. Mater. 2015, 25, 1374.
[91] J. M. Gisbert-González, J. M. Feliu, A. Ferre-Vilaplana, E. Herrero, J. Phys. Chem. C
2018, 122, 19004.
[92] M. Moglianetti, J. Solla-Gullón, P. Donati, D. Pedone, D. Debellis, T. Sibillano, R.
Brescia, C. Giannini, V. Montiel, J. M. Feliu, P. P. Pompa, ACS Appl. Mater. Interfaces
2018, 10, 41608.
[93] X. Huang, Z. Zhao, J. Fan, Y. Tan, N. Zheng, J. Am. Chem. Soc. 2011, 133, 4718.
[94] S. Xie, H. Zhang, N. Lu, M. Jin, J. Wang, M. J. Kim, Z. Xie, Y. Xia, Nano Lett. 2013, 13,
6262.
[95] K. Jang, H. J. Kim, S. U. Son, Chem. Mater. 2010, 22, 1273.
[96] H. Duan, N. Yan, R. Yu, C.-R. Chang, G. Zhou, H.-S. Hu, H. Rong, Z. Niu, J. Mao, H.
Asakura, T. Tanaka, P. J. Dyson, J. Li, Y. Li, Nat. Commun. 2014, 5, 3093.
[97] M. Zhou, H. Wang, M. Vara, Z. D. Hood, M. Luo, T. H. Yang, S. Bao, M. Chi, P. Xiao,
Y. Zhang, Y. Xia, J. Am. Chem. Soc. 2016, 138, 12263.
[98] A. X. Yin, X. Q. Min, Y. W. Zhang, C. H. Yan, J. Am. Chem. Soc. 2011, 133, 3816.
[99] X. Huang, Y. Li, Y. Li, H. Zhou, X. Duan, Y. Huang, Nano Lett. 2012, 12, 4265.
[100] J. Wu, A. Gross, H. Yang, Nano Lett. 2011, 11, 798.
[101] C. Zhang, S. N. Oliaee, S. Y. Hwang, X. Kong, Z. Peng, Nano Lett. 2016, 16, 164.
[102] J. Zhang, H. Yang, J. Fang, S. Zou, Nano Lett. 2010, 10, 638.
[103] C. Zhang, S. Y. Hwang, A. Trout, Z. Peng, J. Am. Chem. Soc. 2014, 136, 7805.
[104] Y. Xia, X. Xia, H.-C. Peng, J. Am. Chem. Soc. 2015, 137, 7947.
[105] Y. Wang, J. He, C. Liu, W. H. Chong, H. Chen, Angew. Chem. Int. Ed. 2015, 54, 2022.
[106] P. Liu, R. Qin, G. Fu, N. Zheng, J. Am. Chem. Soc. 2017, 139, 2122.
[107] M. I. Vesselinov, Crystal Growth For Beginners: Fundamentals Of Nucleation, Crystal
Growth And Epitaxy (Third Edition), World Scientific Publishing Company, 2016.
[108] X. Xia, S. Xie, M. Liu, H. C. Peng, N. Lu, J. Wang, M. J. Kim, Y. Xia, Proc. Natl. Acad.
Sci. U. S. A. 2013, 110, 6669.
[109] M. Chi, C. Wang, Y. Lei, G. Wang, D. Li, K. L. More, A. Lupini, L. F. Allard, N. M.
Markovic, V. R. Stamenkovic, Nat. Commun. 2015, 6, 8925.
[110] Q. Fu, X. Bao, Chem. Soc. Rev. 2017, 46, 1842.
34

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

[111] K. B. Eisenthal, Chem. Rev. 2006, 106, 1462.


[112] T. H. Yang, K. D. Gilroy, Y. Xia, Chem. Sci. 2017, 8, 6730.
[113] Q. Zhang, W. Li, L. P. Wen, J. Chen, Y. Xia, Chem. Eur. J. 2010, 16, 10234.
[114] M. Luo, M. Zhou, R. Rosa da Silva, J. Tao, L. Figueroa-Cosme, K. D. Gilroy, H.-C.
Peng, Z. He, Y. Xia, ChemNanoMat 2017, 3, 190.
[115] M. Luo, A. Ruditskiy, H.-C. Peng, J. Tao, L. Figueroa-Cosme, Z. He, Y. Xia, Adv. Funct.
Mater. 2016, 26, 1209.
[116] S.-H. Liu, W. A. Saidi, Y. Zhou, K. A. Fichthorn, J. Phys. Chem. C 2015, 119, 11982.
[117] S. Ghosh, L. Manna, Chem. Rev. 2018, 118, 7804.

Accepted Manuscript
[118] Y. Wang, H. C. Peng, J. Liu, C. Z. Huang, Y. Xia, Nano Lett. 2015, 15, 1445.
[119] B. Jin, M. L. Sushko, Z. Liu, X. Cao, C. Jin, R. Tang, J. Phys. Chem. Lett. 2019, 10,
1443.
[120] J. Zeng, Y. Zheng, M. Rycenga, J. Tao, Z.-Y. Li, Q. Zhang, Y. Zhu, Y. Xia, J. Am. Chem.
Soc. 2010, 132, 8552.
[121] J. Zeng, X. Xia, M. Rycenga, P. Henneghan, Q. Li, Y. Xia, Angew. Chem. Int. Ed. 2011,
50, 244.
[122] L. Zhang, S.-I. Choi, J. Tao, H.-C. Peng, S. Xie, Y. Zhu, Z. Xie, Y. Xia, Adv. Funct.
Mater. 2014, 24, 7520.
[123] H. Zhang, M. Jin, Y. Xia, Angew. Chem. Int. Ed. 2012, 51, 7656.
[124] S. Xie, N. Lu, Z. Xie, J. Wang, M. J. Kim, Y. Xia, Angew. Chem. Int. Ed. 2012, 51,
10266.
[125] H. Ye, Q. Wang, M. Catalano, N. Lu, J. Vermeylen, M. J. Kim, Y. Liu, Y. Sun, X. Xia,
Nano Lett. 2016, 16, 2812.
[126] R. He, Y. C. Wang, X. Wang, Z. Wang, G. Liu, W. Zhou, L. Wen, Q. Li, X. Wang, X.
Chen, J. Zeng, J. G. Hou, Nat. Commun. 2014, 5, 4327.
[127] S. Xie, H. C. Peng, N. Lu, J. Wang, M. J. Kim, Z. Xie, Y. Xia, J. Am. Chem. Soc. 2013,
135, 16658.
[128] M. McEachran, D. Keogh, B. Pietrobon, N. Cathcart, I. Gourevich, N. Coombs, V.
Kitaev, J. Am. Chem. Soc. 2011, 133, 8066.
[129] M. M. Shahjamali, M. Bosman, S. Cao, X. Huang, X. Cao, H. Zhang, S. S. Pramana, C.
Xue, Small 2013, 9, 2880.
[130] H. Heinz, H. Ramezani-Dakhel, Chem. Soc. Rev. 2016, 45, 412.
[131] W. A. Al-Saidi, H. Feng, K. A. Fichthorn, Nano Lett. 2012, 12, 997.
[132] S. H. Yoo, J. H. Lee, B. Delley, A. Soon, Phys. Chem. Chem. Phys. 2014, 16, 18570.
[133] D. S. Kilin, O. V. Prezhdo, Y. Xia, Chem. Phys. Lett. 2008, 458, 113.
[134] M. Shao, T. Yu, J. H. Odell, M. Jin, Y. Xia, Chem. Commun. 2011, 47, 6566.
[135] S. E. Mason, I. Grinberg, A. M. Rappe, Phys. Rev. B 2004, 69, 161401.
[136] G. T. K. K. Gunasooriya, M. Saeys, ACS Catal. 2018, 8, 10225.
[137] S. Yamagishi, T. Fujimoto, Y. Inada, H. Orita, J. Phys. Chem. B 2005, 109, 8899.
[138] S. Zhou, D. Huo, S. Goines, T.-H. Yang, Z. Lyu, M. Zhao, K. D. Gilroy, Y. Wu, Z. D.
Hood, M. Xie, Y. Xia, J. Am. Chem. Soc. 2018, 140, 11898.
[139] S. K. Ghosh, S. Nath, S. Kundu, K. Esumi, T. Pal, J. Phys. Chem. B 2004, 108, 13963.
[140] S. K. Balavandy, K. Shameli, D. R. B. A. Biak, Z. Z. Abidin, Chem. Cent. J. 2014, 8, 11.
[141] H. C. Peng, S. Xie, J. Park, X. Xia, Y. Xia, J. Am. Chem. Soc. 2013, 135, 3780.
[142] G. Collins, M. Schmidt, G. P. McGlacken, C. O’Dwyer, J. D. Holmes, J. Phys. Chem. C
2014, 118, 6522.
35

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

[143] P. Jiang, S.-Y. Li, S.-S. Xie, Y. Gao, L. Song, Chem. Eur. J. 2004, 10, 4817.
[144] M. Wu, A. M. Vartanian, G. Chong, A. K. Pandiakumar, R. J. Hamers, R. Hernandez, C.
J. Murphy, J. Am. Chem. Soc. 2019, 141, 4316.
[145] R. Z. Li, A. Hu, D. Bridges, T. Zhang, K. D. Oakes, R. Peng, U. Tumuluri, Z. Wu, Z.
Feng, Nanoscale 2015, 7, 7368.
[146] B. Nikoobakht, M. A. El-Sayed, Langmuir 2001, 17, 6368.
[147] M. R. Langille, M. L. Personick, J. Zhang, C. A. Mirkin, J. Am. Chem. Soc. 2012, 134,
14542.
[148] M. De, C.-C. You, S. Srivastava, V. M. Rotello, J. Am. Chem. Soc. 2007, 129, 10747.

Accepted Manuscript
[149] X. Xia, J. Zeng, L. K. Oetjen, Q. Li, Y. Xia, J. Am. Chem. Soc. 2012, 134, 1793.
[150] X. Huang, Z. Zhao, Y. Chen, E. Zhu, M. Li, X. Duan, Y. Huang, Energy Environ. Sci.
2014, 7, 2957.
[151] Z. Niu, Y. Li, Chem. Mater. 2013, 26, 72.
[152] L. M. Rossi, J. L. Fiorio, M. A. S. Garcia, C. P. Ferraz, Dalton Trans. 2018, 47, 5889.
[153] S. Campisi, M. Schiavoni, C. Chan-Thaw, A. Villa, Catalysts 2016, 6, 185.
[154] H. Heinz, C. Pramanik, O. Heinz, Y. Ding, R. K. Mishra, D. Marchon, R. J. Flatt, I.
Estrela-Lopis, J. Llop, S. Moya, R. F. Ziolo, Surf. Sci. Rep. 2017, 72, 1.
[155] M. Crespo-Quesada, J. M. Andanson, A. Yarulin, B. Lim, Y. Xia, L. Kiwi-Minsker,
Langmuir 2011, 27, 7909.
[156] C. Aliaga, J. Y. Park, Y. Yamada, H. S. Lee, C.-K. Tsung, P. Yang, G. A. Somorjai, J.
Phys. Chem. C 2009, 113, 6150.
[157] Y. Borodko, S. E. Habas, M. Koebel, P. Yang, H. Frei, G. A. Somorjai, J. Phys. Chem. B
2006, 110, 23052.
[158] H. Cheng, Z. Cao, Z. Chen, M. Zhao, M. Xie, Z. Lyu, Z. Zhu, M. Chi, Y. Xia, Nano Lett.
2019, 19, 4997.
[159] S. Bao, M. Vara, X. Yang, S. Zhou, L. Figueroa-Cosme, J. Park, M. Luo, Z. Xie, Y. Xia,
ChemCatChem 2017, 9, 414.
[160] S. Bao, X. Yang, M. Luo, S. Zhou, X. Wang, Z. Xie, Y. Xia, Chem. Commun. 2016, 52,
12594.
[161] R. A. Hughes, E. Menumerov, S. Neretina, Nanotechnology 2017, 28, 282002.
[162] K. D. Gilroy, J. Puibasset, M. Vara, Y. Xia, Angew. Chem. Int. Ed. 2017, 56, 8647.
[163] K. D. Gilroy, A. Sundar, M. Hajfathalian, A. Yaghoubzade, T. Tan, D. Sil, E. Borguet, R.
A. Hughes, S. Neretina, Nanoscale 2015, 7, 6827.
[164] A. S. Preston, R. A. Hughes, T. B. Demille, V. M. Rey Davila, S. Neretina, Acta Mater.
2019, 165, 15.
[165] L. Huang, M. Liu, H. Lin, Y. Xu, J. Wu, V. P. Dravid, C. Wolverton, C. A. Mirkin,
Science 2019, 365, 1159.
[166] K. M. Koczkur, S. Mourdikoudis, L. Polavarapu, S. E. Skrabalak, Dalton Trans. 2015,
44, 17883.
[167] S. Zhou, T.-H. Yang, M. Zhao, Y. Xia, Chin. J. Chem. Phys. 2018, 31, 370.
[168] T.-H. Yang, S. Zhou, K. D. Gilroy, L. Figueroa-Cosme, Y.-H. Lee, J.-M. Wu, Y. Xia,
Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 13619.
[169] T. H. Yang, H. C. Peng, S. Zhou, C. T. Lee, S. Bao, Y. H. Lee, J. M. Wu, Y. Xia, Nano
Lett. 2017, 17, 334.
[170] M. Liu, K. D. Gilroy, H. C. Peng, M. Chi, L. Guo, Y. Xia, Chem. Commun. 2016, 52,
13159.
36

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

[171] S. Peng, J. S. Okasinski, J. D. Almer, Y. Ren, L. Wang, W. Yang, Y. Sun, J. Phys. Chem.
C 2012, 116, 11842.
[172] Q. Liu, M. R. Gao, Y. Liu, J. S. Okasinski, Y. Ren, Y. Sun, Nano Lett. 2016, 16, 715.
[173] Y. Wu, D. Qin, J. Am. Chem. Soc. 2018, 140, 8340.
[174] H. Tsunoyama, N. Ichikuni, H. Sakurai, T. Tsukuda, J. Am. Chem. Soc. 2009, 131, 7086.
[175] R.-Y. Zhong, K.-Q. Sun, Y.-C. Hong, B.-Q. Xu, ACS Catal. 2014, 4, 3982.

Accepted Manuscript

37

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Table 1. Summary of shapes that have been successfully obtained for noble-metal nanocrystals
using facet-selective capping agents in the setting of one-pot synthesis.

Capped
Composition Metal Capping agent Shape of the nanocrystal References
facet
Cl− {100} cube [46,47]

Ag Br− {100} cube, bar, penta-twinned rod/wire [48-52]

Accepted Manuscript
PVP {100} cube, penta-twinned rod/wire [40-44,53-55]
citrate {111} plate [56-58]
citrate {111} plate [59,60]
I− {111} plate [61,62]
Au
bovine serum albumin {111} plate [63,64]
PVP {111} tetrahedron, octahedron, plate [65,66]
CTA+ {221} trisoctahedron [67]
Br− {100} cube, bar, penta-twinned rod/wire [68-72]
penta-twinned rod/wire, right
I− {100} [73-75]
bipyramid

Pd tetrahedron, octahedron,
citric acid {111} [76,77]
decahedron, icosahedron, plate
Monometallic tetrahedron, octahedron,
formaldehyde {111} [78]
System decahedron, icosahedron, plate
CO {111} tetrahedron, tetrapod, plate [79,80]
Br− {100} cube [81-83]
CO {100} cube [84-86]
PVP {100} cube [87]
3-hydroxybutyric acid {100} cube [88]
Pt
peptide T7 {100} cube [88-90]
citrate {111} octahedron, tetrahedron [91,92]
peptide S7 {111} tetrahedron [88-90]
sodium polyacrylate {111} tetrahedron [39]
methylamine {411} octapod [93]
Br− {100} cube [83]

Rh citric acid {111} tetrahedron [94]


oleylamine {111} plate [95]
PVP {111} plate [96]
Br− {100} cube [97]
Pd-Pt
Br− and trace I− {100} cube [98]
Alloy[a]
I− {100} cube (Pt51Pd49) [99]
oleic acid {100} cube (Pt3Pd) [100]

38

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Cl− {111} octahedron (Pt52Pd48) [99]


2− and
C2O4 HCHO {111} tetrahedron [98]
Pt-Cu CO {100} cube (Pt4Cu)
[101]
CO {111} octahedron (Pt3Co)
Pt-Co diphenyl ether {111} octahedron (Pt3Co)
oleic acid {100} cube (Pt3Co) [100]
Pt-Fe oleic acid {100} cube (Pt3Fe)
Pt-Rh CO {111} octahedron (PtRh) [101]

Accepted Manuscript
octahedron (Pt3Ni) [86,101,102]
Pt-Ni CO {111}
octahedron (Pt1.5Ni) [103]
diphenyl ether {111} octahedron (Pt3Ni, PtNi)
[100]
oleic acid {100} cube (Pt3Ni, PtNi, PtNi3)

[a] Depending on the composition of the alloy, the same capping agent may have selectivity towards different facets.

39

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 1. Electroplating offers a simple and versatile method for coating a metal substrate with a
thin layer made of a different metal. (A) Metal spoons electroplated with Ag, Cu, Au, Pt, Pd, Rh
and Ni (left to right). (B) A substrate with a V-shaped recess on the surface and (C–E) outcomes
of electroplating for this substrate: (C) retention of roughness when the deposition rate is uniform
across the surface; (D) in reality, the higher current density at ridges leads to a thicker deposition
at these sites than in the recessed region; and (E) formation of a uniform coating by adding a
“leveling agent” to selectively reduce the deposition rate at the ridges. Figure (A) was Reprinted
from Ref. [31] with permission. Copyright 2012 Elsevier.

40

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 2. Langmuir isotherms for adsorbates differing in adsorption equilibrium constant K.

41

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 3. Schematic illustration showing the thermodynamic and kinetic products in terms of total
free energy, and the activation energy barrier to the transformation from a kinetic product to a
thermodynamic product (indicated by the arrow).

42

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 4. (A) Models of the (111), (100), and (110) planes of a fcc metal and the corresponding
numbers of dangling bonds per surface unit cell (NB). (B) Schematic illustration of the role of a
capping agent in directing the growth of a single-crystal seed into nanocrystals with different
shapes. Reprinted from Ref. [104] with permission. Copyright 2015 American Chemical Society.

43

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 5. Schematic illustration of a 2-D square seed whose side faces are covered by a capping
agent and the products obtained under different conditions. Reprinted from Ref. [5], copyright
2017 Wiley-VCH.

44

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 6. Use of PVP as a capping agent for the selective stabilization of {100} facets on Ag
nanocrystals. Schematic illustrations and TEM images of (A, C–E) Ag nanocubes with different
sizes and (B) penta-twinned nanowires. The images in (A–E) were reprinted from Ref. [40, 44,
113, 53, 49], respectively, with permission. Copyright 2002 AAAS, 2003 American Chemical
Society, 2010 Wiley-VCH, 2013 American Chemical Society and 2016 American Chemical
Society.

45

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 7. Halide ions as capping agents for the selective stabilization of {100} facets on metal
nanocrystals. (A–D) TEM images of (A) Pd, (B) Ag, (C) Pt, and (D) Rh nanocubes that were
obtained in the presence of Br− ions. (E, F) TEM images of (E) Pd and (F) Ag nanobars that were
obtained in the presence of Br− ions. (G, H) TEM images of (G) Pd and (H) Ag penta-twinned
nanowires obtained by introducing I− and Br− ions, respectively. (I) Time-lapsed in situ TEM
images showing the evolution of a Au decahedral seed to truncated decahedron and finally to
penta-twinned nanorod in the presence of Br− ions. The images in (A–I) were reprinted from Ref.
[70, 49, 82, 83, 72, 48, 73, 50, 119], respectively, with permission. Copyright 2009 Springer, 2016
American Chemical Society, 2010 American Scientific Publishers, 2008 American Chemical
Society, 2009 Wiley-VCH, 2007 American Chemical Society, 2017 American Chemical Society,
2016 American Chemical Society and 2019 American Chemical Society.

46

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 8. CA as a capping agent for the selective stabilization of {111} facets on Pd nanocrystals.
Schematic illustrations and TEM images of (A) octahedra, (B) icosahedra, (C) decahedra, and (D)
plates obtained in the presence of CA. The images in (A–C) were reprinted from Ref. [76] with
permission. Copyright 2007 Wiley-VCH. The image in D was reprinted from Ref. [77] with
permission. Copyright 2018 Royal Society of Chemistry.

47

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 9. CO as a capping agent for the selective stabilization of {111} or {100} facets on metal
nanocrystals. (A–C) TEM images of (A) Pd plates, (B) Pd tetrahedra, and (C) Pd tetrapods
enclosed by {111} facets. (D) TEM image of Pt nanocubes enclosed by {100} facets. (E) TEM
and (F) High-resolution TEM (HRTEM) images of Pt-Ni alloy octahedra enclosed by {111} facets.
The image in (A) was reprinted from Ref. [79] with permission. Copyright 2010 Nature Publishing
Group. The images in (B) and (C) were reprinted from Ref. [80] with permission. Copyright 2012
American Chemical Society. The images in (D–F) were reprinted from Ref. [86] with permission.
Copyright 2013 American Chemical Society.

48

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 10. Peptides serving capping agents for the selective stabilization of {100} or {111} facets
on Pt nanocrystals. (A) Schematics and TEM images of nanocubes and octahedra that were
obtained by introducing peptides T7 and S7, respectively. (B) Schematics and TEM images of T7-
capped Pt nanocrystals at different concentrations of T7. The images in (A) were reprinted from
Ref. [89] with permission. Copyright 2011 Nature Publishing Group. Copyright 2013 American
Chemical Society. The images in (B) were reprinted from Ref. [90] with permission. Copyright
2015 Wiley-VCH.

49

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 11. Influence of capping agent on the growth pattern of nanocrystals. (A) Schematic
illustration of the synthesis of Rh cubic nanoframes. (B) TEM image and (C) HRTEM image and
EDX mapping of the Pd-Rh core-frame nanocubes. (D) TEM image and the (E) 3D atomic model
of Rh cubic nanoframes. (F-H) TEM images of (F) Rh octahedral nanoframes, (G) Au triangular
nanoframes, and (H) Au decahedral nanoframes. The plots and images in (A–E) were reprinted
from Ref. [124] with permission. Copyright 2012 Wiley-VCH. The image in (F) was reprinted
from Ref. [127] with permission. Copyright 2013 American Chemical Society. The image in (G)
was reprinted from Ref. [129] with permission. Copyright 2013 Wiley-VCH. The image in (H)
was reprinted from Ref. [128] with permission. Copyright 2011 American Chemical Society.

50

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 12. Computational simulation of capping agent on the surface of metal nanocrystals. (A)
The calculated binding energies of Br atoms on the various Pd facets as a function of surface
coverage density. (B) Binding configurations (top and side views) of CA on Ag(100) and (111)
surfaces. (C) Binding configurations (top and side views) of peptide S7 on Pt(100) and (111)
surfaces, and their corresponding phenyl rings highlighted in yellow. (D) Correlation of Pt
tetrahedra yields with calculated peptide adsorption energies for Pt(100) and (111) surfaces. The
plots in (A) were reprinted from Ref. [132] with permission. Copyright 2014 Royal Society of
Chemistry. The plots in (B) was reprinted from Ref. [133] with permission. Copyright 2008
Elsevier. The plots in (C) and (D) were reprinted from Ref. [88] with permission. Copyright 2013
American Chemical Society.

51

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 13. Spectroscopy analysis of capping agent on the surface of metal nanocrystals. (A) UV-
vis analysis of the as-prepared Au nanospheres and after modification with different. (B) FTIR
analysis of the species on the surface of Au nanospheres and reference spectra of the corresponding
compounds. (C) XPS analysis (N 1s spectra) of the PVP on Pd nanocubes and concave nanocubes.
(D) ICP-MS analysis of the coverage density (ϕ) of Br− ions on the surface of Pd nanocubes with
different edge lengths (L). The plot in (A) was reprinted from Ref. [139] with permission.
Copyright 2004 American Chemical Society. The plot in (B) was reprinted from Ref. [138] with
permission. Copyright 2018 American Chemical Society. The plot in (C) was reprinted from Ref.
[142] with permission. Copyright 2014 American Chemical Society. The data in (D) was reprinted
from Ref. [141] with permission. Copyright 2013 American Chemical Society.

52

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Figure 14. (A) Schematic illustrations of the growth of a Ag cubic seed with an edge length of a
in the presence of PVP at a high concentration C1 and a critical concentration C2 (C1 > C2). (B–G)
Accepted Manuscript
SEM images of Ag polyhedra grown from 40-nm cubic seeds in the presence of (A–C) 1.0 mM
and (D–F) 0.1 mM PVP (55000 g/mol), respectively. The samples were collected from the
synthesis at different time points: (B, E) 5 min, (C, F) 10 min, and (D, G) 20 min. Reprinted from
Ref. [149] with permission. Copyright 2012 American Chemical Society.

53

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 15. Removal of capping agent and ligand exchange without altering the shape of the
nanocrystals. (A) Pd 3d XPS spectra and (B) concentration of Br− ions on Pd nanocubes subjected
to aging in aqueous solution or EG containing CA at different temperatures for 18 h. (C) FTIR
spectra of TTAB-capped Pt nanocubes as a function of duration for UVO treatment. (D) ATR-
FTIR spectra of PVP-capped Pd nanocubes after UVO treatment (3 h) and the other controlled
samples without UVO treatment. (E) Schematic illustration and TEM images of indirect ligand
exchange on a Au nanosphere as assisted by the deposition and etching of Ag. The spectra in (A)
and plot in (B) were reprinted from Ref. [141] with permission. Copyright 2013 American
Chemical Society. The spectra in (C) was reprinted from Ref. [156] with permission. Copyright
2009 American Chemical Society. The spectra (D) were reprinted from Ref. [155] with permission.
Copyright 2011 American Chemical Society. The images in (E) were reprinted from Ref. [138]
with permission. Copyright 2018 American Chemical Society.

54

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 16. Capping-free synthesis of nanocrystals. (A-D) Synthesis of capping-free Pd@Pt3-4L
core-shell octahedral nanocrystals. (A) Schematic of the procedure that includes loading of Pd
octahedral nanocrystals onto a carbon support, followed by the conformal deposition of Pt ultrathin
shells to produce the Pd@Pt3-4L nanocrystals. (B) TEM and (C) HRTEM images of the Pd-Pt3-4L
nanocrystals. (D) N1s XPS spectra of the Pd@Pt3-4L nanocrystals synthesized with and without
PVP. (E–J) TEM images of capping-free (E) Ag icosahedral, (G) Ag decahedral, and (I) Ag
cuboctahedral nanocrystals supported on SiO2 substrates. (F-J) HRTEM images of an individual
(F) icosahedral, (H) decahedral, and (J) cuboctahedral particle. Red lines indicate the twin planes.
The spectra and image in (A–D) were reprinted from Ref. [159] with permission. Copyright 2017
Wiley-VCH. The images in (E–J) were reprinted from Ref. [162] with permission. Copyright 2017
Wiley-VCH.
55

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Figure 17. Influence of capping agent on the catalytic activity and selectivity of metal nanocrystals.
(A) The effect of PVP on the catalytic activity of Au nanocrystals towards an aerobic oxidation
reaction. (B) Product selectivity of Au nanocrystals with different degrees of PVP removal towards
the hydrogenation reactions of p-CNB and CAL. The schemes in (A) was reprinted from Ref. [174]
with permission. Copyright 2009 American Chemical Society. The plot and schemes in (B) were
reprinted from Ref. [175] with permission. Copyright 2014 American Chemical Society.

56

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Biographical sketch:

Accepted Manuscript
Younan Xia studied at the University of Science and Technology of China (B.S., 1987) and
University of Pennsylvania (M.S., 1993), and received his Ph.D. from Harvard University in 1996
(with George M. Whitesides). He started as an assistant professor of chemistry at the University
of Washington (Seattle) in 1997 and joined the department of biomedical engineering at
Washington University in St. Louis in 2007 as the James M. McKelvey Professor. Since 2012, he
holds the position of Brock Family Chair and GRA Eminent Scholar in Nanomedicine at Georgia
Tech.

Tung-Han Yang received his B.S. in chemical engineering from National Cheng-Kung University,
Taiwan, in 2009, He then studied at National Tsing-Hua University, Taiwan, and received his M.S.
and Ph.D. in materials science and engineering in 2011 and 2017, respectively. In 2015–2017, he
was a visiting student in the Xia group. Currently, he is a postdoctoral fellow Georgia Tech under
the supervision of Prof. Dong Qin. His research interest includes a quantitative analysis of the
shape-controlled synthesis of noble-metal nanocrystals.

57

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.201911135

Accepted Manuscript
Yifeng Shi received her B.S. degree in chemical engineering from Sichuan University, China, in
2017. She joined the Xia group in 2017 as a graduate student. Her current research focuses on the
surface science of noble-metal nanocrystals and their applications in heterogeneous catalysis.

Annemieke Janssen received her B.S. in chemistry from University College Utrecht in 2016, and
her M.S. in materials science from Imperial College, London in 2017. She joined the Xia group in
2018 as a graduate student. Her current research focuses on the development of new methods for
the synthesis of noble-metal nanocrystals.

58

This article is protected by copyright. All rights reserved.

You might also like