You are on page 1of 15

Chem Soc Rev

View Article Online

TUTORIAL REVIEW

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Cite this: Chem. Soc. Rev., 2013,


42, 2497

View Journal | View Issue

Controlled synthesis of colloidal silver nanoparticles in


organic solutions: empirical rules for nucleation
engineering
Yugang Sun*
Controlled synthesis of colloidal nanoparticles in organic solutions is among the most intensely studied
topics in nanoscience because of the intrinsic advantages in terms of high yield and high uniformity in
comparison with aqueous synthesis. However, systematic studies on the formation mechanism of

Received 28th July 2012

nanoparticles with precisely tailored physical parameters are barely reported. In this tutorial review, we

DOI: 10.1039/c2cs35289c

take the synthesis of dierent Ag nanoparticles as an example to rule out the general principles for
controlling the nucleation process involved in the formation of colloidal Ag nanoparticles in organic
solutions, which enables the synthesis of high-quality nanoparticles.

www.rsc.org/csr

1. Introduction
Silver (Ag) is a ductile, malleable coinage metal that exhibits the
highest electrical and thermal conductivity among all metals

Center for Nanoscale Materials, Argonne National Laboratory, 9700 South Cass
Avenue, Argonne, Illinois 60439, USA. E-mail: ygsun@anl.gov
Part of the chemistry of functional nanomaterials themed issue.

Yugang Sun received his BS and


PhD degrees in chemistry from
the University of Science and
Technology of China (USTC) in
1996 and 2001, respectively. He
is currently a sta scientist for the
Center for Nanoscale Materials at
Argonne National Laboratory. He
is the 2007 recipient of The
Presidential
Early
Career
Awards for Scientists and
Engineers (PECASE) and the
2008 recipient of DOEs Oce of
Yugang Sun
Science Early Career Scientist
and Engineer Award. His current research interests focus on the
synthesis of a wide range of nanostructures, including metal
nanoparticles with tailored properties, the development of in situ
synchrotron X-ray techniques for real-time probing of nanoparticle
growth, and the application of these nanomaterials in energy
storage, photocatalysis, and sensing.

This journal is

The Royal Society of Chemistry 2013

and high optical reflectivities, resulting in Ag being a widely


used material in many areas such as electric contacts and
conductors, mirrors, and catalysis of chemical reactions. As
sizes of Ag particles decrease down to the nanometer scale, they
exhibit many unique properties that cannot be observed in bulk
Ag. For example, the high ductility of Ag dramatically reduced in
Ag nanowires with fivefold twinning structures.1 Synthesis of Ag
nanoparticles boomed in the past decade and their corresponding
properties and applications were extensively studied.25 This
progress has advanced the commercialization of manmade Ag
nanomaterials that represent the most widely used materials in
nanotechnology consumer products (i.e., 313 Ag-based products
as analyzed on March 10, 2011).6 For instance, Ag nanoparticles
have been used as a class of broad-spectrum antimicrobial
reagents in medical and consumer products such as household
antiseptic sprays and antimicrobial coating for medical devices.7,8
Water filters incorporating Ag nanowires have been demonstrated
to be very ecient for cleaning water that is polluted with
bacteria.9 Due to the large surface-to-volume ratios of the Ag
nanoparticles in comparison with their bulk counterparts, Ag
nanoparticles have been used as classic catalysts for important
industrial reactions including oxidation of ethylene to ethylene
oxide, propylene to propylene oxide, and methanol to formaldehyde.10,11 Heterocyclizations, addition of nucleophiles to
alkynes (or allenes, or olefins), cycloaddition reactions (e.g.,
enantioselective [2+3]-cycloaddition of azomethine and nitrilimine),
[4+2]-cycloaddition of imines, and acetylenic CspH and CspSi bond
transformations can also be achieved through Ag-catalyzed
processes.12,13 The high electrical and thermal conductivities
of Ag make Ag nanoparticles to be widely used in electronics

Chem. Soc. Rev., 2013, 42, 2497--2511

2497

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Tutorial Review
industry as conductive fillers in conductive adhesives14 and thermal
interfacial materials.15 Most recently, two-dimensional (2D)
random networks of Ag nanowires have been exploited to
serve as transparent conductive films due to the fact that the
low percolation threshold for Ag nanowires assures a high
percentage of open areas in the conductive networks.16 In
combination with the thin diameters of Ag nanowires that are
responsible for the mechanical flexibility of the nanowires,
such 2D networks are very promising to replace the traditional
rigid doped metal oxide conductive films, such as the most
commonly used tin-doped indium oxide (ITO).17
In addition to these commercial applications, Ag nanoparticles also represent an important class of optical materials
related to a recent hot research field, i.e., plasmonics.18 Ag
nanoparticles exhibit strong surface plasmon resonances
(SPRs) under illumination of light due to strong coherent
oscillation of free surface electrons in the nanoparticles, resulting
in strong absorption and scattering of incident light. As a consequence, dispersions of Ag nanoparticles always exhibit a colorful
appearance. The evanescent electrical fields near the surface of an
Ag nanoparticle are usually very high, providing hot spots to
enhance Raman scattering1923 and fluorescence24 of molecules or
emitters (such as quantum dots and upconversion nanocrystals)
adjacent to the Ag nanoparticle. The unique SPRs in Ag nanoparticles can benefit their traditional use such as in catalytic
oxidation reactions (e.g., ethylene epoxidation, CO oxidation, and
NH2 oxidation) because excitation of SPRs on the surfaces of the
Ag nanoparticles can form energetic electrons that are transferrable to chemical species adsorbed on the nanoparticle surfaces.25
For example, in the commercially important partial oxidation of
ethylene to form ethylene oxide O2-dissociation represents the ratelimiting elementary step that requires a large thermal energy
(corresponding to a high temperature) to drive this reaction.
Illumination of the Ag nanoparticles can excite plasmons on the
Ag surface to populate O2 antibonding orbitals and so form a
transient negatively ionic state, which thereby facilitates the ratelimiting O2-dissociation reaction. As a result, the thermal energy
and temperature can be lowered to drive this oxidation reaction,
leading to an increase in energy eciency and long-term stability
of catalysts and product selectivity. These results imply that
continuous study of the unique properties of Ag nanoparticles
can help us exploit their novel applications.
Intensive studies in the past decade clearly show that the
physical parameters including size, shape, surface coating, and
surrounding environment of an Ag nanoparticle strongly influence its properties and thus its performance in applications.
For example, unpromoted, Ag3 clusters and B3.5 nm Ag
nanoparticles on alumina supports can catalyze the direct
propylene epoxidation by O2 to selectively form propylene oxide
with high activity at low temperatures.26 In contrast, using
commercial industrial catalysts containing non-selected Ag
nanoparticles the reaction selectivity and activity at low temperatures
dramatically decreased. In another example, cubic Ag nanoparticles
bounded with {100} facets exhibit much higher catalytic capability
toward oxidation of styrene with tert-butyl hydroperoxide than
Ag nanoplates mainly bounded with {111} facets,27 indicating

2498

Chem. Soc. Rev., 2013, 42, 2497--2511

Chem Soc Rev


that enhanced catalytic performance can be achieved by
carefully choosing nanoparticles with appropriate shapes as
catalysts. Controlling the shape of Ag nanoparticles can also
change their optical properties over a broader spectral range,2830
thus their optoelectronic applications such as solar cells.31,32
From these examples, it is clear that controlled synthesis of
colloidal Ag nanoparticles is critical to tailor their properties as
well as optimize their performance in applications. Material
scientists have witnessed great successes in the synthesis of
various Ag nanoparticles in the past decade.3,5,28,33 For example,
the shapes of the synthesized Ag nanoparticles include spheres,
spheroids, cubes, cuboctahedrons, octahedrons, tetrahedrons,
decahedrons, icosahedrons, thin plates, rods or wires. Although
significant progress has been made and a number of very good
reviews are available, there is still lack of review articles focusing
on the controlled synthesis of Ag nanoparticles in organic
solutions. The advantages for synthesizing Ag nanoparticles in
organic solvents include high yield, narrow size distribution, and
ease in assembly of the synthesized particles into superlattices in
comparison with the nanoparticles synthesized in aqueous
solutions.34,35 In this tutorial review, the empirical principles
for controlling the synthesis of colloidal Ag nanoparticles in
organic solvents are discussed by summarizing the work done by
our group. The controllability relies on chemically engineering
the nucleation processes involved in the formation of Ag nanocrystals. In Section 2 the classic nucleation theory and the
corresponding classic LaMer model for the formation of colloidal
nanoparticles are briefly discussed to highlight that engineering
nucleation processes can be an ecient strategy for tuning the
parameters of final Ag nanoparticles. Exemplar syntheses of Ag
nanoparticles with dierent sizes, shapes, and composites are then
discussed with details in Section 3 to demonstrate how to manipulate the nucleation processes by tuning the chemistry of the
synthetic reactions. A brief conclusion and personal perspectives
are provided in the final section to wrap up the review.

2. Classical nucleation theory


In general, colloidal Ag nanoparticles are synthesized through
either reduction of Ag+ ions with reducing reagents (or reductive
solvents) or thermal decomposition of organometallic compounds
in the presence of surfactant molecules that can attach to the
nanoparticles surfaces to stabilize them. The basic model used to
describe the formation of colloidal nanocrystals in a solution
phase was presented by LaMer and Dinegar in 1950 and the
model is summarized in Fig. 1a.36 According to this model, zerovalence Ag (Ag0) should be continuously provided to maintain a
sustainable growth of Ag nanoparticles. As a result, an appropriate
chemical reaction is first chosen to continuously generate Ag0 in
the solution. As long as more Ag0 are produced, the solution is
saturated with Ag0 quickly. Even at the saturation concentration
(Cs), the Ag0 still cannot spontaneously condense into solid nuclei
because forming a new solid phase in the homogeneous liquid
environment is an energy-consuming process (Fig. 1b). As a result,
only when the concentration of the Ag0 species reaches a critical
value, i.e., critical concentration (Ccrit), the Ag0 can condense to

This journal is

The Royal Society of Chemistry 2013

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Chem Soc Rev

Fig. 1 (a) LaMer model describing nucleation and growth of nanocrystals as a


function of reaction time and concentration of precursor atoms. Adapted with
permission from ref. 36. (b) Classical nucleation model showing the free energy
diagram for nucleation. Adapted with permission from ref. 41.

form nuclei. Once stable nuclei are formed, they can grow larger at
a lower concentration of Ag0 that is slightly above Cs because this
process is a less energy-consuming process or an energy-saving
process. As a result, the nucleation and growth steps are two
relatively separated processes: formation of nuclei occurs only at a
concentration of Ag0 much higher than Cs, otherwise growing
the existing nuclei dominates. Therefore the two individual steps
(i.e., nucleation and growth) can be reasonably engineered by
tuning the concentration of Ag0, leading to a controlled synthesis
of Ag nanoparticles with appropriate parameters.
Intensive studies on the synthesis of colloidal nanoparticles
have proven that nucleation is critical to determine the properties
of the final nanoparticles. Crystal nucleation can be considered
as a chemical reaction that takes solvated precursor atoms or
molecules (e.g., Ag0 for the synthesis of Ag nanoparticles) into a
solid-state crystalline product. As a chemical reaction, one can
understand the nucleation process from both thermodynamic and
kinetic aspects. In the classical nucleation theory (Fig. 1b), the
driving force for spontaneous phase transition is the exothermicity
of lattice formation. In this thermodynamic aspect, the free energy
change required for the formation of nuclei (DG) is determined by

This journal is

The Royal Society of Chemistry 2013

Tutorial Review
the sum of the free energy change for the phase transformation
(DGv) and the free energy change for the formation of a solid
surface (DGs). As the solid-state crystals are more stable than the
solvated precursors, DGv is negative to decrease the total Gibbs free
energy of the system. In contrast, the introduction of solid/liquid
interfaces generally increases the free energy with the increase in
the surface area of the nuclei. As a result, the evolution of nuclei
depends on the competition between a decrease in DGv, which
favors condensation of solvated precursors into nuclei, and an
increase in DGs, which destabilizes the nuclei toward solvation in
proportion to the crystals surface area. When the radii (R) of the
nuclei are very small, the positive surface free energy DGs term
dominates the total free energy change, leading the small nuclei to
be dissolved. When the size of the nuclei increases, the total free
energy change reaches a maximum (DG*) at a critical size (R*)
and then turns over and continuously decreases to favor the
stabilization and growth of the nuclei. From this thermodynamic
aspect, one can change the pathway for the formation of nuclei by
modulating the function of surface free energy and/or volume free
energy to change the dependence of the total free energy on the
size of the nuclei. As a result, controlling the nucleation process for
the synthesis of colloidal Ag nanoparticles can be realized through
the possible strategies: (i) varying surfactants that can change the
surface free energy of Ag nuclei; (ii) forming nuclei of dierent
materials that exhibit DGv and DGs dierent from Ag nuclei,
followed by their chemical transformation to Ag nuclei; (iii)
changing the reaction environment that can influence the stability
of the nuclei, such as etching and dissolving the nuclei.
According to the Arrhenius reaction rate equation, kinetics
of the nucleation reaction can be described by the steady-state

rate of nucleation, J A exp DG
kT , which equals the number
of nuclei formed per unit time per unit volume. In this
equation, k is the Boltzmanns constant and A is the preexponential factor. The theoretical value of the pre-exponential
factor is given as 1030 cm3 s1 although the value is very
dicult to measure in practice.37 This kinetic factor depends
on the mobility of precursor species (e.g., Ag0 for the synthesis
of Ag nanoparticles) that can influence the rate of attachment
of the precursor species to the critical nuclei. Since the mobility
of precursor species varies rapidly with temperature, the temperature dependence of the pre-exponential factor can be quite
significant. In addition, variation of temperature also changes
the value of the exponential term. As a result, from the kinetic
aspect we can change reaction temperature to influence the
kinetics of the nuclei formation. The value of DG* also plays an
important role in determining the nucleation kinetics. As
discussed in the previous paragraph, this thermodynamic
energy diagram can be tuned by controlling the chemical
environment of the synthetic reactions. As a consequence, the
nucleation kinetics can be tuned to control the synthesis of Ag
nanoparticles.
The classical nucleation theory indicates that supersaturated precursor species spontaneously condense into nuclei
with critical sizes (this is called self-nucleation) followed by
gradually enlarging the nuclei with continuous addition of
precursor species (Fig. 1a). However, recent studies using the

Chem. Soc. Rev., 2013, 42, 2497--2511

2499

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Tutorial Review

Chem Soc Rev

advanced in situ techniques reveal that the nucleation process


is usually complex and involves multiple steps that are not
reflected in the classical nucleation theory.38 For example, the
nuclei with small sizes may coalesce to form larger nuclei that
are more stable to support the gradual enlargement through
continuous attachment of precursor species. Two models, i.e.,
the LifshitzSlyozovWagner (LSW) model39,40 and the two-step
model,41 are proposed to describe the detailed nucleation process
for the formation of solid crystals from reaction solutions. Even
though these progresses are promising for comprehensively understanding the nucleation process, the principles for controlling the
nucleation process revealed by the classical nucleation theory can
still serve as the practical guidelines for design and synthesis of
colloidal nanoparticles with tailored parameters. Typical examples
for the synthesis of dierent Ag nanoparticles in organic solutions
are discussed in the next section.

3. Controlled synthesis of Ag nanoparticles


Two classic reaction systems, i.e., reduction of AgNO3 with hot
oleylamine (cis-1-amino-9-octadecene, OAm) and reduction of
AgNO3 with hot ethylene glycol (EG), are used as the model
systems to highlight the importance of nucleation engineering
in determining the final Ag nanoparticles. In the synthesis,
both OAm and EG are reaction media and are also used as the
solvents for dissolving AgNO3 and the possible additional
surfactant. Dissolution of AgNO3 in the solvents releases Ag+
ions that might coordinate with the solvent molecules or
surfactant molecules.42 In the following content, Ag+ ions is
used for simplicity regardless of the coordination states. Since
the reducing ability of OAm and EG highly depends on the
reaction temperature, the generation rate of the precursor Ag0
species and thus the following nucleation kinetics for the
formation of Ag nanocrystals can be tuned by controlling the
reaction temperature. However simply controlling the reaction
temperature is not enough to synthesize high-quality Ag nanoparticles. As highlighted in Section 2, controlling the nucleation
kinetics can also be realized by tuning the reaction thermodynamics. This review focuses on the aspect: controlled
synthesis of colloidal Ag nanoparticles in organic solutions
through manipulation of chemistries of the synthetic reactions
that influence the nucleation thermodynamics and kinetics.
3.1. Synthesis of icosahedral Ag nanoparticles through fast
reduction of Ag+ ions
Reactions in OAm have been widely used to synthesize colloidal
nanoparticles made of a broad range of materials including
magnetic materials,43 metals,44 and oxides.45 Fig. 2 shows
transmission electron microscopy (TEM) images of the Ag
nanoparticles with varying sizes that have been synthesized
through a fast reduction of Ag+ ions in hot OAm.46,47 In a typical
synthesis of 10 nm Ag nanoparticles (Fig. 2f), addition of
1 mmol AgNO3 to 20 ml OAm at room temperature forms a
suspension that is then heated up to 60 1C. The temperature is
maintained until the granular AgNO3 crystals are completely
dissolved. The solution is colorless, indicating that the OAm

2500

Chem. Soc. Rev., 2013, 42, 2497--2511

Fig. 2 TEM images of Ag nanoparticles with dierent diameters. The number in


front of the  sign in each frame is the average diameter of the nanoparticles
shown in the same frame and the number following the  sign represents the
three standard deviations (3s). Adapted with permission from ref. 46.

molecules are not active enough to reduce Ag+ ions. Quickly


heating the solution with a ramp Z10 1C min1 to 180 1C
dramatically enhances the reducing ability of OAm to reduce
Ag+ ions at a very high rate, which is confirmed by the
observation of quick appearance of a dark color corresponding
to the SPRs of Ag nanoparticles formed in the solution. Ag+ ions
are completely reduced within a couple of minutes although
the nanoparticles formed at this stage exhibit a broad size
distribution. By taking advantage of the Ostwald ripening
process, continuous incubation of the nanoparticles under
the reaction conditions for a longer time, such as 1 hour,
significantly decreases their size distribution, resulting in the
formation of Ag nanoparticles with spherical morphologies and
a very narrow size distribution. Post size-selection based on
simple centrifugation can further narrow their size distribution. As shown in Fig. 2f, the size distribution of the 10 nm
nanoparticles is only 5%. By controlling the reaction conditions
including temperature, growth time, and additives (e.g., oleic
acid), size of the Ag nanoparticles can be tuned in the range of
220 nm and the typical size distributions are controlled to be
lower than 10% (Fig. 2). In the synthesis, OAm serves as
both reducing reagent and surfactant that helps stabilize the
synthesized Ag nanoparticles. The surfaces of the nanoparticles
are coated with OAm molecules through interactions between
the amine groups (NH2) and surface Ag atoms in the nanoparticles. Fourier transform infrared (FTIR) spectroscopy of the
synthesized Ag nanoparticles shows major peaks essentially
similar to the characteristic peaks of pure OAm molecules
except a new peak at 1540 cm1. The appearance of this new
peak indicates the formation of AgN bonds. Meanwhile the
peak at about 1070 cm1 corresponding to the vibration mode

This journal is

The Royal Society of Chemistry 2013

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Chem Soc Rev


of CN slightly shifts to the position with a lower wave number.
Such dierences in FTIR spectra imply that the surfaces of the
Ag nanoparticles are primarily coated with OAm molecules
through the formation of chemical bonds between the surface
Ag atoms in the nanoparticles and the nitrogen atoms in
the OAm molecules. The long hydrocarbon chains of the
OAm molecules assist the Ag nanoparticles to well disperse in
non-polar and low-polar solvents, such as hexane, toluene, and
chloroform. In addition, the dense OAm capping layers on the
nanoparticles surfaces prevent the Ag nanoparticles from
being oxidized by air, leading the dispersions of the Ag nanoparticles to exhibit an excellent stability in the ambient
environment.
Due to the high growth rate, the Ag nanoparticles always
exhibit morphologies close to spheres that have the lowest
surface energy when they are small.48 Their exact morphologies
have been carefully studied by high-resolution TEM (HRTEM).
As shown in Fig. 3, regardless of the particle size each Ag
nanoparticle exhibits an icosahedral shape with the characteristic
co-existence of twofold, threefold, and fivefold symmetries (Fig. 3a).
Each icosahedral nanoparticle has twenty faces terminated with
{111} crystalline facets of face-centered cubic (f.c.c.) Ag, thirty edges
and twelve vertices. Formation of this unique morphology requires
the existence of 30 fivefold twin planes that connect 20 tetrahedral
subunits. Fig. 3be present the HRTEM images of dierently sized
Ag nanoparticles along dierent rotational axes, confirming their
icosahedral morphology with multiply twinned crystallinity. The
existence of twin planes is responsible for the inhomogeneous
contrast that is reflected by the randomness of dark spots in the
TEM image of individual Ag nanopartilces (Fig. 2). These
characterizations indicate that the synthesized Ag nanoparticles with dierent sizes shown in Fig. 2 have the consistent
morphology, surface coating, and narrow size distribution.
Such consistency makes these Ag nanoparticles to be an
ideal class of model materials for studying the size-dependent
properties. For example, the dependence of the absorption
peak position of the Ag nanoparticles on their particle size is
very interesting: as the particle size decreases from B20 nm the
absorption peak blue-shifts but then turns over near 12 nm and
strongly red-shifts. This exceptional size dependence is quite
dierent from large nanoparticles (with diameters >20 nm)
for which the peak position constantly blue-shifts as particle
size decreases.29 This turnover dependence is ascribed to the
significant eect of surface chemistry between the capping
molecules (i.e., OAm) and the surface Ag atoms in the nanoparticles that cannot be ignored for small nanoparticles.
3.2. Synthesis of Ag nanocubes mediated with the formation
of AgCl nanocrystals
Given the fact that the reduction of Ag+ ions with hot OAm is
very fast, it is dicult to control and manipulate the nucleation
process to grow nanoparticles with morphologies other than
icosahedron. One possibility is to introduce another reaction that
can also quickly form solid nanocrystals with dierent crystalline
structures (or shapes). This additional nucleation process has
a lower nucleation barrier (i.e., DG*) than the self-nucleation

This journal is

The Royal Society of Chemistry 2013

Tutorial Review

Fig. 3 (a) Schematic drawings and (be) HRTEM images of the individual Ag
nanoparticles shown in Fig. 2 viewed along dierent rotational axes: (left) twofold,
(middle) threefold, and (right) fivefold ones, revealing their icosahedral morphology. The red dashed lines in (a) highlight the twin planes corresponding to these
symmetries. The diameters of the nanoparticles shown in (be) are (b) 5.3 nm, (c)
7.3 nm, (d) 10.0 nm, and (e) 15.6 nm, respectively. Insets on the bottom left of the
images presented in (b, c) are the fast Fourier transforms (FTTs) of the corresponding HRTEM images showing the nanoparticles symmetries. Scale bars in (b),
(c), (d), and (e) represent 4, 5, 5, and 10 nm, respectively, and apply to all the
images in the corresponding rows. Reproduced with permission from ref. 47.

associated with direct reduction of Ag+ with OAm, leading to


a competition with the self-nucleation from Ag0 species. As
shown in Fig. 4a, halide ions, such as chloride ions, can be
added to the reaction system to quickly precipitate with Ag+
ions to form silver chloride (AgCl) nanoparticles. As a result, in
the reaction system silver species nucleate through two dierent ways to form two dierent types of particles. The Ag
particles derived from AgCl particles are usually polyhedral
single crystals49 while the Ag particles formed through selfnucleation are multiply twinned crystals (similar to those
shown in Fig. 2) with sizes smaller than the single-crystal
particles. Continuously heating the reaction system facilitates
an Ostwald ripening process to gradually dissolve the smaller
multiply twinned particles and grow the single-crystal polyhedral
particles into nanocubes.

Chem. Soc. Rev., 2013, 42, 2497--2511

2501

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Tutorial Review

Fig. 4 (a) Modified LaMer model including an additional nucleation process


besides the self-nucleation. (b) Schematic illustration of the major steps involved
in the formation of single-crystal Ag nanocubes. (c) TEM image of the synthesized
Ag nanocubes. The inset in (c) is the convergent beam electron diraction pattern
of an individual nanocube. (b, c) Adapted with permission from ref. 50.

Fig. 4b shows an example for the synthesis of Ag nanocubes


with the assistance of dimethyl distearyl ammonium chloride
(DDAC) in hot OAm mixed with octyl ether (OE).50 In a typical
synthesis, 8.0 mL of OE and 1.0 mL of OAm are sequentially
added to a 50 mL three-neck flask connected to a Schlenk line
purged with nitrogen. OE is desirable for dissolving DDAC and
OAm plays a role in reducing Ag+ ions and stabilizing the
synthesized Ag nanocubes. To the binary solvent (OEOAm)
are added 0.3 mmol DDAC powders. Heating the solvent to
60 1C and maintaining the temperature for 10 min completely
dissolves the DDAC powders. The resulting colorless solution is
then quickly heated up to 260 1C at a ramp of B10 1C min1. To
this hot DDAC solution is quickly injected 1.0 mL of OAm
solution of AgNO3 with a concentration of 0.2 M. The reaction
solution instantaneously turns milky yellowish, indicating the
quick formation of both AgCl and Ag nanoparticles. Continuous
reaction diminishes the milky color within 2 min, indicating the
disappearance of AgCl nanoparticles. Maintaining the reaction

2502

Chem. Soc. Rev., 2013, 42, 2497--2511

Chem Soc Rev


at 260 1C for 1 h completes the synthesis of pure Ag nanocubes
as shown in Fig. 4c. At elevated temperatures DDAC can release
free Cl ions to precipitate with Ag+ ions to form AgCl nanocrystals very quickly once the AgNO3 solution is injected. Meanwhile Ag+ ions are also reduced by OAm to form multiply
twinned Ag nanoparticles similar to those shown in Fig. 2
through the self-nucleation process. Continuous reaction
reduces the AgCl nanocrystals to single-crystal Ag particles with
polyhedral morphologies. An Ostwald ripening process then
facilitates the growth of the single-crystal polyhedrons to cubes
with consumption of the smaller multiply twinned particles.
Fig. 4c presents a typical TEM image of the synthesized Ag
nanoparticles through this DDAC-mediation reaction, clearly
showing their cubic morphology with slight truncation at the
corners and uniform size with an average edge length of 34 nm.
Each Ag nanocube exhibits a highly uniform contrast in the TEM
images, indicating the nanoparticles are free of twin defects. The
convergent beam electron diraction pattern (inset, Fig. 4c)
obtained by aligning the electron beam perpendicular to one
of the six surfaces of an individual Ag nanocube exhibits a
simple square symmetry, confirming that each nanocube is a
single crystal with its surfaces bounded by {100} facets. In this
synthesis, there are at least three dierent nucleation processes
involved in the formation of AgCl nanocrystals, multiply twinned
Ag nanocrystals, and solid phase transition from AgCl to singlecrystal Ag crystals.
The complex nucleation and growth processes involved in
the synthesis of Ag nanocubes have been probed in real time
with the time-resolved high-energy X-ray diraction (XRD).51
The use of high-energy synchrotron X-ray beam is advantageous
because of the strong penetration of high-energy X-ray into
liquid solutions and reaction vessels as well as weak absorption
of the X-ray in the solvents and reaction precursors. The weak
absorption of X-ray eliminates possible undesirable X-rayinduced reactions. Fig. 5a presents the 2D contour of the
XRD patterns recorded at dierent times during the synthesis
of Ag nanocubes. It clearly shows the appearance of AgCl and
Ag crystals as well as the transformation of AgCl to Ag at
dierent reaction stages: AgCl nanocrystals are formed first
once the reaction is initiated; Ag nanocrystals are then formed
through reduction of Ag+ ions with OAm; AgCl disappears due
to the reduction with OAm when the time is long enough. More
information on the reaction can be obtained from the variations in
the XRD peak area, which is approximately proportional to the mass
of crystalline materials, and peak width, which is related to the
lateral dimensions of nanocrystals. Fig. 5b plots the integrated peak
areas of the Ag(111) peak and the AgCl(200) peak that represent the
major peaks of these two crystalline materials. According to the
Scherrer equation, lateral dimensions of individual crystalline
domains can be calculated from the peak width of XRD patterns.
Fig. 5c compares the crystalline domain size in Ag nanoparticles
along the (111) direction and in AgCl nanoparticles along the (200)
direction. As shown in Fig. 5b, AgCl nanoparticles are formed within
the first 3 s through the fast precipitation between Ag+ and Cl ions.
Reducing Ag+ ions with OAm is initiated only after the complete
formation of AgCl (i.e., at 3 s). The reduction of Ag+ ions is also very

This journal is

The Royal Society of Chemistry 2013

View Article Online

Chem Soc Rev

Tutorial Review

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

quick and most of the Ag+ ions are reduced within several
seconds (period I). The peak areas of both Ag and AgCl exhibit
plateaus in the period II while their particle sizes slightly
increase, indicating that Ostwald ripening processes occur. Only
when the reaction time is long enough, i.e., at B60 seconds,
AgCl nanoparticles start to be reduced and transformed into
single-crystal Ag nanoparticles in the period III. During this
phase transition process, the reaction rate follows the Avrami
phase-boundary based nuclei growth model in a 3D fashion.52,53
The Avrami exponent is determined to be B4, indicating the
nucleation process with a constant nucleation rate. The size of
the AgCl nanoparticles calculated from the XRD patterns
remains essentially constant during this period, indicating that
once the phase transition of an AgCl nanoparticle is initiated it
can be quickly reduced to pure Ag before the phase transition of
another AgCl nanoparticle starts. This chemical transformation
process lasts B40 s. The increase in the crystalline size of Ag
during period III is ascribed to the fact that the sizes of the
single-crystal Ag nanoparticles derived from the AgCl nanoparticles are larger than the multiply twinned Ag nanoparticles
formed during period II. As more and more Ag nanoparticles
are formed through the chemical transformation of AgCl nanoparticles, the average crystalline size of Ag nanoparticles continuously increases until all of the AgCl nanoparticles are
reduced. In period IV, the mixture of single-crystal and multiply
twinned Ag particles in the reaction solution undergoes an
Ostwald ripening process. Because the multiply twinned Ag
particles exhibit smaller sizes than the single-crystal Ag particles
and contain twinning defects, continuous incubation of the
nanoparticles gradually dissolves the multiply twinned Ag particles
and drives the single-crystal Ag particles to grow into uniform Ag
cubes as shown in Fig. 4c. Apparently the time-resolved highenergy XRD studies provide more information on the complex
nucleation and growth processes involved in the synthesis of Ag
nanocubes than that obtained through the traditional sampling
strategy. Techniques with higher temporal resolutions are
expected in the future for better understanding the synthesis.
3.3. Controlled synthesis of Ag nanoparticles through
selectively etching defective nuclei

Fig. 5 Time-resolved XRD patterns recorded from the reaction solution for the
synthesis of Ag nanocubes shown in Fig. 4. (a) 2D contour plot of the XRD
patterns at dierent reaction times. The black and red sticks represent the peak
positions and relative intensities of the standard powder XRD patterns for f.c.c.
Ag and f.c.c. AgCl, respectively. Blue arrows highlight the time when the AgNO3
solution was injected to initiate the reaction. The wavelength of X-ray was
0.1771 . Data were collected on the X-ray Operations and Research beamline
1-ID at the Advanced Photon Source, Argonne National Laboratory. (b) Variation
in the integrated peak areas of the Ag(111) peak and the AgCl(200) peak as a
function of reaction time. (c) Dependence of the lateral dimensions of the
crystalline domains in the Ag nanoparticles along the {111} crystalline direction
and in the AgCl nanoparticles along the {200} direction as a function of the
reaction time. The dotted lines highlight the time periods (I, II, III, and IV) assigned
according to the important processes discussed in the text. Adapted with
permission from ref. 51.

This journal is

The Royal Society of Chemistry 2013

Another possible means to control the nucleation pathway is to


slow down the reaction rate for reducing Ag+ ions. In this case
one can have enough time to manipulate the reaction environment to select nuclei (i.e., seeds) with desirable crystalline
structures that determine the morphology of final nanoparticles. In a system with slow reaction rate, Ag atoms usually
self-nucleate into nuclei with crystalline structures that
fluctuate between single crystals and twined crystals. Such
structural fluctuation is consistent with the TEM observation
of small (o5 nm) metal particles made of Ag and Au showing
that a mild heating induced by the electron beam could force
fluctuations between single-crystal and twinned morphologies.54
The rate of such fluctuations decreases with the increase in
crystal size. As a result, one can find a way to defeat thermodynamics to selectively dissolve twinned nuclei to obtain high
yield of single-crystal Ag nanoparticles (Fig. 6a). In contrast,

Chem. Soc. Rev., 2013, 42, 2497--2511

2503

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Tutorial Review

Fig. 6 (a) Modified LaMer model describing the inclusion of an extra step for
selecting nuclei with appropriated crystal structures. (b) Schematic illustration of
the possible mechanism for the selective growth of single-crystal Ag nanoparticles (truncated cubes and tetrahedrons) through reduction of AgNO3 in
hot EG in the presence of PVP, NaCl, and oxygen. (c) TEM image of the Ag
nanoparticles formed at 44 h 10 min, showing the absence of twin planes in the
nanoparticles. (d) SEM image of the nanoparticles formed at 45 h containing
exclusively truncated cubes (indicated by a white octagon) and truncated
tetrahedrons (indicated by a white hexagon). (bd) Adapted with permission
from ref. 55.

twined particles dominate the product if the growth of single


crystal nuclei is prevented.
Fig. 6b shows an example that shaped Ag nanoparticles with
controlled crystalline structures can be synthesized through the
reduction of Ag+ ions with hot EG at a lower reduction rate in
comparison with the reaction systems with hot OAm. Silver
atoms formulate structures of nuclei with sizes less than 2 nm
at early stage of this reaction and these structures fluctuate
between twinned crystals and single crystals. As these nuclei
grow in size, the structural fluctuations slow down until the

2504

Chem. Soc. Rev., 2013, 42, 2497--2511

Chem Soc Rev


crystallites are locked in a specific morphology. The nuclei with
stable crystalline structures then serve as seeds to guide their
further growth into nanoparticles with appropriate shapes
and crystalline structures. In this reaction system, poly(vinyl
pyrrolidone) (PVP) is used as a surfactant and the EG solutions
are heated at 148 1C.55 When there are oxygen and trace
amounts of Cl ions in the reaction system, the initially formed
twinned structures could be dissolved because the twinning
defects provide active sites for the oxidation reaction between
Ag and oxygen. In this etching process, Cl ions may serve
as coordinate ligands to promote the oxidation reaction by
stabilizing the resultant smaller nanoclusters because no AgCl
crystals are observed during the synthesis. On the other hand,
when the structures are single crystalline, the nanoparticles
continue their growth with the assistance of PVP. As a result,
products consisting of pure single crystals are obtained by
adding a small amount of sodium chloride (0.06 mM NaCl) to
the reaction system to air. Fig. 6c and d show the electron
microscopic images of the single-crystal Ag nanoparticles
formed at dierent reaction times. The TEM image of the Ag
nanoparticles formed at 44 h 10 min (Fig. 6c) shows that each
nanoparticle has a quasi-spherical morphology free of apparent
facets. All the Ag nanoparticles are without twinning defects.
Growth of these quasi-spherical nanoparticles leads to the
development of well-defined {100} and {111} facets, resulting
in the formation of truncated cubes (highlighted by an white
octagon) and truncated tetrahedrons (highlighted by an white
hexagon) (Fig. 6d for the sample formed at 45 h). The insets
in Fig. 6d present the convergent beam electron diraction
patterns recorded by directing the electron beam perpendicular
to a (100) facet of a truncated cube (upper right) and a (111)
facet of a truncated tetrahedron (lower left), respectively.
The diraction patterns exhibit the standard symmetries of
single-crystal f.c.c. Ag, confirming the single crystallinity of the
synthesized Ag nanoparticles. As the reaction continues, the
size of Ag nanoparticles increases accordingly while their single
crystallinity remains.
When the reaction shown in Fig. 6b occurs in the absence of
oxygen, the oxidation reaction of Ag cannot be initiated to
selectively dissolve the twinned nuclei. For example, the polyol
reaction system including 0.06 mM NaCl produces uniform Ag
nanowires that are grown from the twinned particles formed at
the early stage when the reaction is performed under argon.55
Alternatively the concentration of oxygen can be controlled by
adding either Fe(II) or Fe(III) species to the reaction solution,
thus to select the crystallinity of the final Ag nanoparticles.
Due to the increased reducing activity of EG at the elevated
temperatures, the stable iron species is Fe(II) in hot EG. As
shown in Fig. 7a, in the reaction system molecular oxygen (O2)
dissolved in EG adsorbs on the Ag surfaces and dissociates to
atomic oxygen (Oa) for catalyzing oxidation reaction on the Ag
surfaces.56 Since Fe(II) species are more active than Ag atoms to
be oxidized, the adsorbed oxygen on the Ag surfaces can be
consumed by the Fe(II) species. The resulting Fe(III) species are
reduced back to Fe(II) by hot EG, leading to a continuous removal
of the adsorbed oxygen from the Ag surfaces. As a result, the

This journal is

The Royal Society of Chemistry 2013

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Chem Soc Rev

Fig. 7 Synthesis and characterization of Ag nanowires with fivefold twin planes.


(a) Illustration of the possible mechanism showing how the oxygen species
adsorbed on the Ag surfaces can be removed by Fe(II). Depletion of adsorbed
oxygen on the Ag surfaces is responsible for blocking the oxidative etching of the
twinned Ag nuclei with fivefold twinning structure that can grow to form Ag
nanowires. Reproduced with permission from ref. 57. (b) SEM image of the Ag
nanowires synthesized through a reduction of AgNO3 in hot EG in the presence
of PVP, NaCl, and Fe(acac)3. The inset is a TEM image of the cross section of a
nanowire that was viewed along the longitudinal axis of the nanowire. (c, d)
HRTEM images of a cross-sectioned Ag nanowire obtained by cutting it against
the planes that are perpendicular to the longitudinal axis of the nanowire. (e)
Schematic drawing of an Ag nanowire with a highly strained core including
lattice defects and a less strained sheath. The coreshell structure exposes
strained/defective lattices (highlighted by the random lines) only at the ends of
the nanowire to the surrounding environment. Due to the high reactivity of the
strained/defective surfaces and stability of the less strained side surfaces, the
short nanowires formed at the early stage tend to grow longer by preferentially
adding more Ag atoms to the strained/defective end surfaces. (c, d) Reproduced
with permission from ref. 58.

oxidative dissolution of twinned Ag nuclei can be eciently


prevented, leading to a preferential growth of twinned nuclei to
uniform Ag nanowires because the twinned particles are more
thermodynamically stable than the single-crystal particles.57 Fig. 7b
shows an SEM image of the Ag nanowires synthesized from the
reaction solution containing 2.2 mM tris(acetylacetonato)iron(III)
(Fe(acac)3), clearly highlighting their high aspect ratios. The cross
section of each Ag nanowire exhibits a pentagonal symmetry due to
the existence of five {111} twin planes that crossed along a line in
the center of the nanowire (inset, Fig. 7b). Generally, each nanowire

This journal is

The Royal Society of Chemistry 2013

Tutorial Review
can be considered to be composed of five single crystalline f.c.c.
subunits sharing their {111} crystallographic facets. However,
the five subunits cannot completely fill spaces as predicted by
the simple solid geometry model, leading to the formation of a
solid-angle deficiency. This angular deficiency leads to lattice
strains and/or defects in the nanowires to fill the 7.351 gap. The
high-resolution XRD patterns indicate that the lattice strains in
the Ag nanowires induce tetragonal distortions in the f.c.c.
lattices.58 Studies on the cross-sectional samples of individual
Ag nanowires with electron microscopy and electron diraction
reveal that the lattice strains distributed non-uniformly, i.e., the
lattice strains are concentrated in the central region of each
nanowire. The HRTEM images of a thick cross-sectional sample
that essentially retains the internal lattice strains and microstructured defects are presented in Fig. 7c and d. The image of
the central region (Fig. 7c) shows that the solid-angle gaps
induce lattice defects including stacking faults, associated
partial dislocations, slips, and possible additional small crystal
domains. These defects are responsible for partially releasing
the strong internal strains to stabilize the tetragonally distorted
nanowires. In contrast, the crystalline lattices near surfaces are
essentially free of defects except the {111} twin planes (highlighted by the red arrow), indicating much less strains in the
surface regions (Fig. 7d). Cross-sectional samples of dierent
nanowires exhibit the similar morphology and microstructures,
implying that each Ag nanowire is a coreshell structure with a
highly strained core that is responsible for the tetragonal
distortion and a thin less-strained sheath that protects the
strained core. The coreshell structure is responsible for the
enhanced stability of the strained Ag nanowires and might
provide the strong driving force for their anisotropic growth.
Because the fivefold twin planes do not twist or bend during
the growth of nanowires, the coreshell geometry and microstructured defects exist throughout the entire nanowires along
their longitudinal axes. As a result, the defects that represent
the most active sites for the addition of Ag atoms during
nanowire growth can be exposed only at the ends of the
nanowires (Fig. 7e). In contrast, the less-strained side surfaces
of the nanowires have lower reactivity towards the attachment
of Ag atoms for growing them thicker. The dierent reactivity
between the end surfaces and side surfaces of the nanowires
may be responsible for anisotropic growth of the nanowires.
The examples shown in Fig. 6 and 7 highlight the importance of trace amounts of additives, e.g., NaCl and Fe(acac)3, in
the selection of stable nuclei and the final nanoparticles. The
presence of Cl ions prompts the oxidative etching of twinned
nuclei due to the higher reactivity of the twinning defects
towards oxygen than the defect-free surfaces of the singlecrystal nuclei. In the reaction systems including iron species,
the reaction between Fe(II) and oxygen species adsorbed on Ag
surfaces can eectively prevent the oxidative etching of twinned
nuclei of Ag. Similar to the high reactivity towards oxidation,
the twinning defects and other crystalline lattice defects also
exhibit higher activity for the deposition of Ag atoms during
nanoparticle growth, leading to a preferential enlargement of
twinned nuclei when both twinned nuclei and single-crystal

Chem. Soc. Rev., 2013, 42, 2497--2511

2505

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Tutorial Review
nuclei coexist. As shown in Fig. 6b, decahedron and icosahedron represent the two typical morphologies of nanoparticles
with fivefold twinning structures. Experimental observations
and theoretical predictions indicate that icosahedral nanoparticles with large sizes (>30 nm) barely exist due to strong
three-dimensional (3D) constraints. In contrast, decahedral Ag
nanoparticles that exhibit tetragonal lattice distortions and
coreshell strain distributions58 are easily elongated along
their fivefold axes to form nanowires. The selective growth of
thermodynamically stable Ag nanowires exhibits a much faster
kinetics than the selective growth of single-crystal Ag nanoparticles, which is reflected from the dierence in reaction
times for the formation of the fivefold twinned nanowires
shown in Fig. 7b (40 min) and the single-crystal nanoparticles
shown in Fig. 6d (45 h).
3.4.

Synthesis of Ag nanoplates in N,N-dimethylformamide

Reduction of Ag+ ions with polyol solvents (e.g., ethylene glycol)


in the presence of PVP has been extensively explored for the
synthesis of Ag nanoparticles with single crystallinity and/or
fivefold twinning. When the polyol solvents are replaced with
N,N-dimethylformamide (DMF), the Ag+ ions can be reduced to
form Ag nanoplates with multiple twin planes parallel to the
basal surfaces of the nanoplates. Previous studies have proven
that DMF represents an organic solvent with powerful reducing
ability against metal ions in the synthesis of metal nanoparticles.59 The reduction can take place at room temperature,
but an increase in temperature can remarkably increase the
reaction rate. In addition, DMF slightly decomposes to a more
easily oxidized amine upon aging or upon catalytic decomposition with a solid base. The resulting amine can accelerate the
reduction of metal ions in particular during the formation of
metal nanoparticles that can provide the solid base to catalyze
the decomposition of DMF. Shortly after the first report of
photochemically synthesized Ag nanoplates with high quality
n and co-workers have demonstrated the
and yield,60 Liz-Marza
preparation of Ag nanoplates in boiled DMF containing AgNO3
and PVP.61 Control experiments indicate that increasing the
concentration of Ag+ ions relative to the concentration of PVP
changes the synthesized particles from isotropic spheres to
anisotropic nanowires and nanoplates. At a concentration of
AgNO3 that is higher than a critical value (i.e., 0.02 M), higher
concentration of PVP is beneficial for improving the yield of Ag
nanoplates. Since the co-existing Ag nanospheres are much
smaller than the nanoplates, the Ag nanoplates can be easily
purified by centrifugation. Time-dependent analysis reveals a
degree of size control based on the reaction time: longer
reaction time leads to larger nanoplates.
In addition to refluxing the reaction solutions with a heating
mantle, the thermal energy can also be delivered to the reaction
solutions with ultrasonication62 and microwave.63 Ag nanoplates have been observed as the major products in both
syntheses. He et al. have compared the reduction of AgNO3 in
dierent solvents (e.g., pyridine, ethanol, DMF, and N-methyl-2pyrrolidone) containing PVP when a microwave oven has been
used to drive the reaction. Ag nanoplates are formed only in

2506

Chem. Soc. Rev., 2013, 42, 2497--2511

Chem Soc Rev


DMF while pseudospherical nanoparticles and irregular nanoparticles are produced in other solvents, indicating that DMF
plays an important role in the formation of Ag nanoplates. Yang
et al. have used the solvothermal method to reduce AgNO3 in
DMF containing PVP to synthesize Ag nanoparticles.64 The
morphologies of the resulting Ag nanoparticles highly depend
on the molar ratio of PVP/AgNO3. Spherical Ag nanoparticles
and a small fraction of Ag nanorods with an aspect ratio of
B2 are formed at PVP/AgNO3 = 0.9. Upon increasing the molar
ratio of PVP/AgNO3 to 5, monodisperse triangular Ag nanoplates are formed in a very high yield and uniformity that are
superior to the Ag nanoplates in the previously reported work.
Larger triangular Ag nanoplates are obtained by continuously
increasing the molar ratio of PVP/AgNO3. The authors argue
that the higher pressure in the solvothermal process is helpful
for the formation and growth of triangular Ag nanoplates.
Compared to the products formed without PVP in DMF, the
authors also argue that PVP plays an important role in the
formation of Ag nanoplates due to its reducing power in
kinetically controlling the nucleation and growth of Ag nanoplates. Although the reducing ability of the end hydroxyl (OH)
groups of PVP has been extensively studied by Xia et al. to
synthesize metal nanoplates in aqueous solutions,65 a similar
role in the formation of Ag nanoplates in organic solutions has
not been confirmed. For example, Hupp and Schatz groups
have demonstrated the successful synthesis of Ag nanoplates by
using carboxylate-functionalized polystyrene (PS) spheres
instead of PVP in DMF solutions.66 As a result, in the synthesis
of Ag nanoplates PVP molecules mainly play similar roles as a
stabilizer of the nanoparticles and a coordination reagent
towards Ag+ ions to the reactions for synthesizing Ag nanocubes
and nanowires discussed in Section 3.3. With higher PVP/
AgNO3 ratios more Ag+ ions can be coordinated and the Ag
nanoparticles can be deeply passivated, leading to a change in
kinetics of nucleation and growth.
In all these examples, DMF is used as the solvent that also
serves as the reducing reagent regardless of other reaction
conditions, indicating the importance of DMF in determining
the anisotropic plate morphologies of Ag nanoparticles.
However, the exact mechanism for the formation of Ag nanoplates is not well understood yet unless the morphological and
structural evolutions can be in situ observed.
3.5. Synthesis of Ag/iron oxide hybrid nanoparticles through
hetero-nucleation
In addition to self-nucleation from Ag0 in an homogeneous
liquid environment, one can preload nanoparticles to a reaction system to provide nucleation sites for condensation of Ag
atoms (Fig. 8a). Due to the existence of foreign nanoparticles,
Ag atoms can more easily condense on the surfaces of the
nanoparticles in comparison with self-nucleation into freestanding Ag nuclei because of the thermodynamic energy
benefit. Nucleation on the existing nanoparticles can decrease
the Ag/solution interfacial surface areas thus lowering the
surface free energy (DGs) (Fig. 1b). The corresponding overall
energy barrier for nucleation of Ag on the preloaded

This journal is

The Royal Society of Chemistry 2013

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Chem Soc Rev

Fig. 8 (a) Modified LaMer model describing the formation of hybrid structures
through hetero-nucleation. (b) Schematic illustration showing the major steps
involved in the synthesis of hybrid nanostructures made of Ag nanodomains and
Fe/FexOy nanodomains. (c) Summary of the TEM images obtained from the
products formed through the synthetic reaction shown in (b) at dierent times
that was adjusted against the time when the AgNO3 solution was injected into
the dispersion of Fe/FexOy nanoparticles. From the left top to the bottom left
following the arrow direction, the reaction times were 0, 2, 180, and 300 s,
respectively. The sample shown in the center was the same as that shown in the
bottom left arc. The images were false colored and the scale bar applies to all the
images. (b, c) Adapted with permission from ref. 67.

nanoparticles decreases, resulting in that a relatively low


concentration of Ag0 species can drive the nucleation process.
This strategy is always called hetero-nucleation. With this
method nanoparticles decorated with Ag nanodomains can be
synthesized. If the original nanoparticles are made of materials

This journal is

The Royal Society of Chemistry 2013

Tutorial Review
dierent from Ag and have dierent properties, the synthesized
hybrid structures can exhibit multiple functionalities.
Fig. 8b shows an example for the synthesis of magnetoplasmonic bi-functional nanoparticles consisting of magnetic
iron oxide (FexOy) nanodomains and plasmonic Ag nanodomains by using amorphous iron nanoparticles (as the preloaded foreign nanoparticles) to mediate the nucleation and
growth of Ag nanodomains on their surfaces.67 In a typical
synthesis, amorphous Fe nanoparticles with uniform sizes are
first synthesized through a thermal decomposition of Fe(CO)5
in 1-octadecene (ODE) containing OAm.68 Separating the
synthesized Fe nanoparticles from the reaction solution followed
by washing them with hexane leads to a partial oxidation of
the nanoparticles surfaces forming thin iron oxide layers that
are also amorphous. Such oxidation is ascribed to the high
reactivity of metallic Fe with the trace amount of oxygen
dissolved in hexane. Formation of the thin FexOy shells passivates the Fe nanoparticles and significantly prevents the inner
Fe cores from quick oxidation.69 Once an OAm solution of
AgNO3 is injected into a hot ODEOAm solvent containing the
amorphous Fe/FexOy nanoparticles, Ag nanodomains quickly
deposit on the surfaces of the Fe/FexOy nanoparticles because
the amorphous FexOy surfaces provide the nucleation sites for
Ag. Due to the fast reduction of Ag+ with hot OAm and the high
density of nucleation sites on the amorphous FexOy surfaces,
this heterogeneous nucleation leads to the formation of multiple Ag domains (as many as eight) on the surface of each
Fe/FexOy nanoparticle. Continuously heating the reaction
system initiates the ripening process of the Ag nanodomains
because of the high mobility of Ag atoms on the FexOy surfaces
at high temperatures, resulting in a gradual decrease in the
average number of the Ag domains on each Fe/FexOy nanoparticle. The ripening process enlarges the most stable Ag
nanodomain on a single Fe/FexOy nanoparticle by consuming
the others until a dimer is formed. During the ripening process,
the iron nanoparticles are converted to hollow iron oxide
nanoshells through a complete oxidation of the iron with
nitrate ions dissociated from AgNO3.
Fig. 8c presents a series of typical TEM images of samples
formed at dierent reaction stages, agreeing well with the
growth mechanism highlighted in Fig. 8b. These samples are
obtained by injecting AgNO3 solution (0.05 M in OAm, 2.0 mL)
into hot (180 1C) ODEOAm (10 mL/0.5 mL) in the presence of
Fe/FexOy coreshell nanoparticles with an average diameter of
14 nm (top left, Fig. 8c), followed by a continuous heating
for dierent times. Mixing the AgNO3 solution with the hot
dispersion of Fe/FexOy nanoparticles leads to an instantaneous
appearance of intense yellow color within 1 s due to the
formation of Ag nanoparticles that exhibit strong SPRs. In
contrast, it takes a much longer time (>60 s) to develop a light
yellow color from a hot ODEOAm solvent without Fe/FexOy
nanoparticles after the AgNO3 solution is injected. The significant dierence in the reaction rate for the formation of Ag
nanoparticles highlights the role of the amorphous Fe/FexOy
nanoparticles in facilitating the nucleation and growth of Ag
nanocrystals from solutions. TEM images of the sample formed

Chem. Soc. Rev., 2013, 42, 2497--2511

2507

View Article Online

Chem Soc Rev

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Tutorial Review

Fig. 9 (a) Schematic illustration describing the synthesis of dumbbell nanostructure made of two dierent nanoparticles linked with amorphous FexOy layers. (bj)
TEM images of Au@FexOy coreshell nanoparticles (c, f, i) and Au@FexOyAg dumbbell nanoparticles (d, g, j) that were synthesized from Au nanoparticles with
dierent sizes (b, e, h). The scale bar shown in (h) applies to all the images.

at 2 s reveal that each Fe/FexOy nanoparticle is decorated with


multiple Ag nanodomains with an average number of 3.6 (top
right, Fig. 8c). As the reaction proceeds, the average number of
Ag domains on each Fe/FexOy particle continuously decreases,
for example, the average number lowers to 1.25 at 180 s (bottom
right, Fig. 8c). When the reaction time is suciently long, the
product is dominated by dumbbell-like dimers that are formed
at 300 s (bottom left and center, Fig. 8c). Each dimer is
consisted of a single Ag domain and a hollow FexOy shell.
During the reaction, the dimensions of the Ag domains and the
morphology of the Fe/FexOy seed nanoparticles also undergo
significant changes such as those highlighted in Fig. 8b.
The success in selective deposition of Ag nanodomains on
the Fe/FexOy nanoparticles is ascribed to that the amorphous
FexOy surfaces provide active sites to facilitate the nucleation
and growth of Ag. As a result, more complicated hybrid
structures can be synthesized by coating nanoparticles with

2508

Chem. Soc. Rev., 2013, 42, 2497--2511

amorphous FexOy layers followed by decoration with Ag nanodomains through the same strategy shown in Fig. 8b. For
instance as shown in Fig. 9a, one can first form a thin layer
of amorphous FexOy around nanoparticles made of varying
materials (e.g., metal, semiconductor, oxide, etc.) through a
decomposition of Fe(CO)5 in a hot ODEOAm solution containing these nanoparticles followed by controlled post-oxidation.
In the next step, the Ag nanodomains can be grown on the
FexOy surfaces, leading to the formation of structures more
complex than those shown in Fig. 8c. Fig. 9bj show the
formation of hybrid structures containing both Au and Ag
nanodomains that are separated by the amorphous FexOy
layers.
3.6.

Summary of the synthesis of Ag nanoparticles

The examples presented in Sections 3.13.5 clearly demonstrate


that the synthesis of colloidal Ag nanoparticles in organic

This journal is

The Royal Society of Chemistry 2013

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Chem Soc Rev


solvents can be controlled by appropriately controlling the reaction
solution chemistries that influence the thermodynamic energy
diagrams involved in the nucleation processes. In summary, fast
reduction of Ag+ ions with hot OAm results in a burst nucleation
and formation of Ag nanoparticles with icosahedral morphology
that represents the morphology with the lowest surface energy for
f.c.c. metal nanoparticles with small sizes. The reaction rate of this
reaction is too fast to be conveniently tuned for synthesizing Ag
nanoparticles with morphologies other than icosahedron. Two
dierent strategies have been demonstrated to control the nanoparticles morphologies. First, to the fast reaction system are added
high-concentration Cl ions that can quickly precipitate with Ag+
ions to form single-crystal AgCl nanocrystals to compete with the
formation of multiply twinned Ag nanoparticles formed from the
direct reduction of Ag+ ions with hot OAm. The single-crystal AgCl
nanoparticles are then chemically converted to single-crystal Ag
nanoparticles with polyhedral morphologies, which can grow into
Ag nanocubes with consumption of the multiply twinned Ag
nanoparticles through an Ostwald ripening process. Second, the
reaction for reducing Ag+ ions can be slowed down to enable the
selection of nuclei with desirable crystalline structures by adding
appropriate chemical additives. Single-crystal Ag nanoparticles can
be achieved through reduction of Ag+ ions with hot EG by
selectively dissolving the nuclei with twinning defects while the
product is mainly composed of fivefold twinned Ag nanowires if
the growth of single-crystal nuclei is not prompted. In addition to
chemical species (e.g., DDAC, Cl, Fe(acac)3, etc.), foreign nanoparticles can also be preloaded to the reaction solution to provide
nucleation sites for condensation of Ag atoms, resulting in hybrid
structures with multiple functionalities. Such hetero-nucleation is
preferential in comparison with the self-nucleation through which
freestanding Ag nanoparticles are formed because the formation
of interfaces between the Ag nuclei and the preloaded nanoparticles can lower the free energy barrier for nucleation. By
applying these rules, the nucleation process can be engineered
to synthesize high-quality Ag nanoparticles shown in Fig. 29 that
exhibit the well-controlled sizes, shapes, and compositions of
hybrids.

4. Conclusions and remarks


The examples discussed in this review clearly demonstrate that
chemically engineering the synthetic reactions can eectively
influence the thermodynamic energy diagram of the nucleation
process to kinetically control the formation of Ag nanocrystals
with tailored parameters including size, shape, crystallinity,
and composites. These strategies, in principle, can be extended
for controlled synthesis of nanoparticles made of materials
other than Ag. As discussed in Section 2, the nucleation process for
the formation of colloidal nanocrystals is usually complicated with
involvement of a number of chemical and physical events (e.g.,
formation of non-crystalline clusters with a magic number of atoms,
coalescence of clusters, crystallization of nuclei, ripening of nuclei,
etc.). Development of in situ techniques that are capable of noninvasively probing the complex nucleation process in real time is
highly demanded to help better understand the nucleation process.

This journal is

The Royal Society of Chemistry 2013

Tutorial Review
The understanding will in turn help us better design and synthesize
high-quality nanoparticles. Environmental transmission electron
microscopy with specially designed thin liquid cells70 and timeresolved synchrotron X-ray techniques (e.g., transmission X-ray
microscopy,71 wide-angle X-ray scattering,72 small-angle X-ray scattering,38 X-ray absorption fine structure,73 etc.) represent the major
advances emerged in the past several years.
The synthesized Ag nanoparticles with well-controlled parameters can be used as a class of physical templates to direct the
deposition of other materials on the surfaces of the Ag nanoparticles to form coreshell nanoparticles with multiple compositions and functionalities. The Ag nanoparticles can also
serve as chemical templates to react with appropriate reagents
to transform the Ag nanoparticles into nanoparticles made of
dierent materials while the resulting nanoparticles can inherit
the morphology and/or crystallinity of the Ag nanoparticles. For
example, galvanic replacement reactions between the Ag nanoparticles and precursors of more noble metals (e.g., Au, Pt, Pd)
result in the formation of hollow metal nanoparticles.74 Reaction with appropriate oxidizing reagents (e.g., S, FeCl3, etc.) can
transform the Ag nanoparticles into semiconductor nanoparticles (e.g., Ag2S, AgCl, etc.).75 Assembly of the synthesized
Ag nanoparticles and the derived nanoparticles through templated transformations into complex superlattices represents
another interesting direction for developing functional materials because coupling between neighboring nanoparticles may
lead to novel properties and applications.76

Acknowledgements
This work was performed at the Center for Nanoscale Materials,
a U.S. Department of Energy, Oce of Science, Oce of Basic
Energy Sciences User Facility under Contract No. DE-AC0206CH11357. Data discussed in this review were partially
obtained with the use of Advanced Photon Source and Electron
Microscopy Center for Materials Research at Argonne National
Laboratory that are supported by the U.S. Department of
Energy, Oce of Science, Oce of Basic Energy Sciences, under
Contract No. DE-AC02-06CH11357. Dr Sheng Pengs eorts on
the synthesis of Ag icosahedral nanoparticles, Ag nanocubes,
and Ag/FexOy hybrid nanoparticles are greatly appreciated.

Notes and references


1 B. Wu, A. Heidelberg, J. J. Boland, J. E. Sader, X. Sun and
Y. Li, Nano Lett., 2006, 6, 468472.
2 D. V. Talapin and Y. Yin, J. Mater. Chem., 2011, 21,
1145411456 and all articles in this special issue.
3 Y. Xia, Y. Xiong, B. Lim and S. E. Skrabalak, Angew. Chem.,
Int. Ed., 2009, 48, 60103.
4 M. Baghbanzadeh, L. Carbone, P. D. Cozzoli and
C. O. Kappe, Angew. Chem., Int. Ed., 2011, 50, 1131211359.
5 M. R. Jones, K. D. Osberg, R. J. Macfarlane, M. R. Langille
and C. A. Mirkin, Chem. Rev., 2011, 111, 37363827.
6 http://www.nanotechproject.org/inventories/consumer/analysis_
draft/.

Chem. Soc. Rev., 2013, 42, 2497--2511

2509

View Article Online

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

Tutorial Review
7 M. E. Quadros and L. C. Marr, Environ. Sci. Technol., 2011,
45, 1071310719.
8 Z.-M. Xiu, Q.-B. Zhang, H. L. Puppala, V. L. Colvin and P. J.
J. Alvarez, Nano Lett., 2012, 12, 42714275.
9 D. T. Schoen, A. P. Schoen, L. Hu, H. S. Kim, S. C. Heilshorn
and Y. Cui, Nano Lett., 2010, 10, 36283632.
10 A. Nagy and G. Mestl, Appl. Catal., A, 1999, 188, 337353.
11 J. G. Serafin, A. C. Liu and S. R. Seyedmonir, J. Mol. Catal. A:
Chem., 1998, 131, 157168.
lvarez-Corral, M. Mun
oz-Dorado and I. Rodrgue12 M. A
Garca, Chem. Rev., 2008, 108, 31743198.
13 Y. Yamamoto, Chem. Rev., 2008, 108, 31993222.
14 D. D. Lu, Y. G. Li and C. P. Wong, J. Adhes. Sci. Technol.,
2008, 22, 815834.
15 A. Munari, J. Xu, E. Dalton, A. Mathewson and K. M. Razeeb,
in 2009 IEEE 59th Electronic Components and Technology
Conference, IEEE, San Diego, CA, 2009, vol. 14, pp. 448452.
16 Y. Sun, Nanoscale, 2010, 2, 16261642.
17 Z. Yu, Q. Zhang, L. Li, Q. Chen, X. Niu, J. Liu and Q. Pei, Adv.
Mater., 2011, 23, 664668.
18 T. W. Odom and G. C. Schatz, Chem. Rev., 2011, 111,
36673668 and all articles in this special issue.
19 M. Moskovits, J. Raman Spectrosc., 2005, 36, 485.
20 A. Tao, F. Kim, C. Hess, J. Goldberger, R. He, Y. Sun, Y. Xia
and P. Yang, Nano Lett., 2003, 3, 1229.
21 Z.-Y. Li and Y. Xia, Nano Lett., 2010, 10, 243249.
22 Y. Sun and G. P. Wiederrecht, Small, 2007, 3, 19641975.
23 Y. Hu and Y. Sun, J. Phys. Chem. C, 2012, 116, 1332913335.
24 W. Feng, L.-D. Sun and C.-H. Yan, Chem. Commun., 2009, 4393.
25 P. Christopher, H. Xin and S. Linic, Nat. Chem., 2011, 3,
467472.
26 Y. Lei, F. Mehmood, S. Lee, J. Greeley, B. Lee, S. Seifert,
R. E. Winans, J. W. Elam, R. J. Meyer, P. C. Redfern,
gl, M. J. Pellin, L. A. Curtiss and
D. Teschner, R. Schlo
S. Vajda, Science, 2010, 328, 224228.
27 R. Xu, D. Wang, J. Zhang and Y. Li, Chem.Asian J., 2006, 1,
888893.
28 M. Rycenga, C. M. Cobley, J. Zeng, W. Li, C. H. Moran,
Q. Zhang, D. Qin and Y. Xia, Chem. Rev., 2011, 111,
36693712.
29 K. L. Kelly, E. Coronado, L. L. Zhao and G. C. Schatz, J. Phys.
Chem. B, 2003, 107, 668677.
30 Y. Sun and Y. Xia, Analyst, 2003, 128, 686691.
31 H. A. Atwater and A. Polman, Nat. Mater., 2010, 9, 205213.
32 A. P. Kulkarni, K. M. Noone, K. Munechika, S. R. Guyer and
D. S. Ginger, Nano Lett., 2010, 10, 15011505.
33 A. Tao, S. Habas and P. Yang, Small, 2008, 4, 310325.
34 E. V. Shevchenko, D. V. Talapin, N. A. Kotov, S. OBrien and
C. B. Murray, Nature, 2006, 439, 5559.
35 X. Ye, J. E. Collins, Y. Kang, J. Chen, D. T. N. Chen,
A. G. Yodh and C. B. Murray, Proc. Natl. Acad. Sci. U. S. A.,
2010, 107, 2243022435.
36 V. K. LaMer and R. H. Dinegar, J. Am. Chem. Soc., 1950, 72,
48474854.
37 V. M. Fokin and E. D. Zanotto, J. Non-Cryst. Solids, 2000,
265, 105112.

2510

Chem. Soc. Rev., 2013, 42, 2497--2511

Chem Soc Rev


38 H. Koerner, R. I. MacCuspie, K. Park and R. A. Vaia, Chem.
Mater., 2012, 24, 981995.
39 I. M. Lifshitz and V. V. Slyozov, J. Phys. Chem. Solids, 1961,
19, 3550.
r Elektrochemie, Berichte der Bunsen40 C. Wagner, Zeitschrift fu

gesellschaft fur physikalische Chemie, 1961, 65, 585591.


41 D. Erdemir, A. Y. Lee and A. S. Myerson, Acc. Chem. Res.,
2009, 42, 621629.
42 Z. Deng, M. Mansuipur and A. J. Muscat, J. Phys. Chem. C,
2009, 113, 867873.
43 D. Ho, X. Sun and S. Sun, Acc. Chem. Res., 2011, 44, 875882.
44 V. Mazumder and S. Sun, J. Am. Chem. Soc., 2009, 131, 4588.
45 H. Imagawa, A. Suda, K. Yamamura and S. Sun, J. Phys.
Chem. C, 2011, 115, 17401745.
46 S. Peng, J. M. McMahon, G. C. Schatz, S. K. Gray and Y. Sun,
Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 1453014534.
47 Y. Sun, S. K. Gray and S. Peng, Phys. Chem. Chem. Phys.,
2011, 13, 1181411826.
48 F. Baletto, R. Ferrando, A. Fortunelli, F. Montalenti and
C. Mottet, J. Chem. Phys., 2002, 116, 38563863.
49 C. Goessens, D. Schryvers, J. Van Landuyt and R. De Keyzer,
Ultramicroscopy, 1992, 40, 151162.
50 S. Peng and Y. Sun, Chem. Mater., 2010, 22, 62726279.
51 S. Peng, J. S. Okasinski, J. D. Almer, Y. Ren, L. Wang,
W. Yang and Y. Sun, J. Phys. Chem. C, 2012, 116,
1184211847.
52 M. Avrami, J. Chem. Phys., 1939, 7, 11031112.
53 S. F. Hulbert, J. Br. Ceram. Soc., 1969, 6, 1120.
54 S. Iijima and T. Ichihashi, Phys. Rev. Lett., 1986, 56, 616619.
55 B. Wiley, T. Herricks, Y. Sun and Y. Xia, Nano Lett., 2004, 4,
17331739.
56 B. A. Sexton and R. J. Madex, Chem. Phys. Lett., 1980, 76,
294297.
57 B. Wiley, Y. Sun and Y. Xia, Langmuir, 2005, 21, 80778080.
58 Y. Sun, Y. Ren, Y. Liu, J. Wen, J. S. Okasinski and D. J. Miller,
Nat. Commun., 2012, 3, 971.
n, Langmuir, 2002,
59 I. Pastoriza-Santos and L. M. Liz-Marza
18, 28882894.
traux, J. I. Cutler and
60 J. E. Millstone, S. J. Hurst, G. S. Me
C. A. Mirkin, Small, 2009, 5, 646664.
n, Nano Lett., 2002, 2,
61 I. Pastoriza-Santos and L. M. Liz-Marza
903905.
62 L.-P. Jiang, S. Xu, J.-M. Zhu, J.-R. Zhang, J.-J. Zhu and
H.-Y. Chen, Inorg. Chem., 2004, 43, 58775883.
63 R. He, X. Qian, J. Yin and Z. Zhu, J. Mater. Chem., 2002, 12,
37833786.
64 Y. Yang, S. Matsubara, L. Xiong, T. Hayakawa and
M. Nogami, J. Phys. Chem. C, 2007, 111, 90959104.
65 I. Washio, Y. Xiong, Y. Yin and Y. Xia, Adv. Mater., 2006, 18,
17451749.
66 E. Hao, K. L. Kelly, J. T. Hupp and G. C. Schatz, J. Am. Chem.
Soc., 2002, 124, 1518215183.
67 S. Peng, C. Lei, Y. Ren, R. E. Cook and Y. Sun, Angew. Chem.,
Int. Ed., 2011, 50, 31583163.
68 S. Peng, C. Wang, J. Xie and S. Sun, J. Am. Chem. Soc., 2006,
128, 10676.

This journal is

The Royal Society of Chemistry 2013

View Article Online

Chem Soc Rev

73 M. Harada and Y. Kamigaito, Langmuir, 2012, 28,


24152428.
74 Y. Yin, C. Erdonmez, S. Aloni and A. P. Alivisatos, J. Am.
Chem. Soc., 2006, 128, 1267112673.
75 Y. Sun, J. Phys. Chem. C, 2010, 114, 21272133.
76 A. Dong, J. CHen, P. M. Vora, J. M. Kikkawa and
C. B. Murray, Nature, 2010, 466, 474477.

Downloaded by University of New Hampshire on 13 March 2013


Published on 16 October 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35289C

69 C. M. Wang, D. R. Baer, L. E. Thomas, J. E. Amonette, J. Antony,


Y. Qiang and G. Duscher, J. Appl. Phys., 2005, 98, 094308.
70 H. Zheng, R. K. Smith, Y.-W. Jun, C. Kisielowski, U. Dahmen
and A. P. Alivisatos, Science, 2009, 324, 13091312.
71 Y. Sun and Y. Wang, Nano Lett., 2011, 11, 43864392.
72 Y. Sun, Y. Ren, D. R. Haener, J. D. Almer, L. Wang, W. Yang
and T. T. Truong, Nano Lett., 2010, 10, 37473753.

Tutorial Review

This journal is

The Royal Society of Chemistry 2013

Chem. Soc. Rev., 2013, 42, 2497--2511

2511

You might also like