You are on page 1of 15

Wet Chemical Synthesis

Wet chemical synthesis is a widely used technique for growing ZnO NP solutions,
in which zinc nitrate (Zn(NO3)2) and sodium hydroxide (NaOH) are generally used
as the precursor reactants and deionized water as the reactive solvent.

From: Multifunctional Photocatalytic Materials for Energy, 2018

Related terms:

ZnO, Nanostructure, Metal Cluster, Titanium Dioxide, Nanocrystals, Nanoparticles,


Metal Nanoparticles, Nanostructured Material, Sol-Gels, Surface Active Agent

View all Topics

Synthesis and Processing of Emerging


Two-Dimensional Nanomaterials
Yasir Beeran Pottathara, ... Sabu Thomas, in Nanomaterials Synthesis, 2019

1.3.2.2 Wet-Chemical Synthesis


Wet-chemical synthesis routes deal with chemical reactions in the solution phase
using precursors at proper experimental conditions. Each wet-chemical synthesis
method differs from the others, meaning that one cannot find a general rule for
these kinds of synthesis approaches. These synthesis strategies have been used for
the preparation of 2D nanomaterials which are unable to be prepared by top-down
approaches. Wet-chemical synthesis routes offer a high degree of controllability and
reproducibility for 2D nanomaterial fabrication. Solvothermal synthesis, template
synthesis, self-assembly, oriented attachment, hot-injection, and interface-mediat-
ed synthesis are the main wet-chemical synthesis routes for 2D nanomaterials.

The solvothermal/hydrothermal approach is a simple and scalable method which


has been widely used for the preparation of 2D inorganic nanomaterials [102]. In a
typical solvothermal process, water or organic solvent acts as the reaction medium
in a closed vessel. In this process, the reaction temperature would be higher than
the boiling point of solvent and, as a result, the solvent will be autogenerated at
high pressure, which improves the crystallinity of the as-synthesized nanocrystals
[103]. Two-dimensional nanosheets of metals [104], metal oxides [105], and metal
chalcogenides [106,107] have been reported by this method. The wet-chemical syn-
thesis of two-dimensional single- and multi layer transition metal dichalcogenides
are schematically illustrated by Fig. 1.7. However, the solvothermal method offers
low cost and high yield for 2D nanomaterial synthesis, it is difficult to figure out the
growth mechanism by this method since the reactions occur in a closed vessel. Also,
the 2D nanosheets synthesized by this method are few-layer rather than single-layer.

Figure 1.7. Schematic illustration of wet-chemical synthesis of two-dimensional sin-


gle- and multilayer transition metal dichalcogenides.Reproduced with permission
from D. Yoo, M. Kim, S. Jeong, J. Han, J. Cheon, J. Am. Chem. Soc. (2014) [107].
Copyright 2014 American Chemical Society.

Template synthesis is another wet-chemical approach which utilizes bulk or presyn-


thesized nanomaterials as templates for the growth of anisotropic nanostructures
[108]. Many reports have been published for the template-based synthesis of 2D
nanomaterials. Jeong et al. [109] reported the growth of hexagonal close-packed
(hcp) gold nanosheets using GO nanosheets as a template. Wei et al. established a
freestanding half-unit-cell -Fe2O3 nanosheet with CuO nanoplate as the template
[110]. Cu-based chalcogenide nanoplates via the cation exchange reaction were
reported using presynthesized CuS nanoplates as the template [111].

In a typical self-assembly process, presynthesized low-dimensional nanocrystals,


such as nanoparticles and nanowires, were spontaneously organized with each
other by noncovalent interactions, such as van der Waals/electrostatic/hydrogen
bonds for the synthesis of 2D nanomaterials [112]. This assembly mainly involves
the combination of low-dimensional nanocrystals to form larger crystals in two
dimensions. Many 2D nanomaterials, such as nanosheets of Au [113], Cu [114], PbS
[115], etc., were reported by the self-assembly approaches of wet-chemical synthesis.
Two-dimensional-oriented attachment is another wet-chemical synthesis method
which offers the fabrication of 2D nanocrystals with nonlayered structures having
well-defined morphology. In this process, adjacent nanocrystals are associated and
bonded together to form 2D nanosheets with a single-crystalline nature. Many 2D
nanosheets were prepared by this technique, such as PbS nanosheets [116], Bi2Se3
nanosheets [117], etc.

The hot-injection method is another wet-chemical synthesis method which is usually


used to prepare monodispersed colloidal nanocrystals with uniform shape and size
[118]. This method involves the rapid injection of highly reactive reactants into a
surfactant, usually oleyl amine or oleyl acid, contained hot solution. The prepared 2D
nanomaterials by this method offer high purity and uniform size and shape. Some
of the disadvantages of this method include its relatively high reaction temperature,
use of surfactants, and difficulties in large-scale production. For the synthesis
of 2D metal coordination polymers, the main wet-chemical synthesis strategy is an
interface-mediated approach [119]. In a typical process, organic ligands confined to
the water–air interface reacted with metal ions dissolved in water and formed sin-
gle-layer dense nanosheets. This process is also extended to polymers and inorganic
nanosheets in addition to metal coordination polymers for the preparation of 2D
nanomaterials.

> Read full chapter

Metal-based semiconductor nanomate-


rials for thin-film solar cells
Wenxi Guo, ... Teng Li, in Multifunctional Photocatalytic Materials for Energy, 2018

8.2.2.1 Fabrication of ZnO NPs


Wet chemical synthesis is a widely used technique for growing ZnO NP solutions, in
which zinc nitrate (Zn(NO3)2) and sodium hydroxide (NaOH) are generally used as
the precursor reactants and deionized water as the reactive solvent. In the first step,
the aforementioned mixed solution was stirred to obtain Zn(OH)2 and then heated
at 180°C for 20 h in an autoclave to grow ZnO NPs. In addition, organic or inorganic
additives can be added to control the morphology of ZnO NPs. Fig. 8.6 shows the
SEM images of the undoped ZnO NPs [27]. However, the solution method limits
the large-scale production of ZnO films, and to solve this problem, other methods
were exploited to grow ZnO NPs, such as a vapor deposition method (VD) and spray
pyrolysis technology to fabricate large-scale ZnO films.
Fig. 8.6. (A) SEM images of the undoped ZnO nanoparticles and (B, C) their corre-
sponding high-magnification images.Reprinted with permission from D. Shao, et
al., High responsivity, fast ultraviolet photodetector fabricated from ZnO nanopar-
ticle-graphene core-shell structures, Nanoscale 5 (2013) 3664–3667.

> Read full chapter

Nanostructured Materials
Maximilian Fichtner, in Frontiers of Nanoscience, 2009

7.2 Synthesis Methods


Wet chemical synthesis and reconditioning of the product can be carried out by
the so-called Schlenk technique, where a vacuum line (10−3 mbar) and a feed line
with purified nitrogen are used for evacuating and inertizing the Schlenk glassware,
which is equipped with plugs and stop cocks. Almost every chemical operation,
from heating under reflux to filtering and evaporation, extraction and drying, can
be carried out under inert conditions by using this technique.

Mechanical synthesis is performed by ball milling in a high-energy ball mill. To obtain


inert conditions, the milling vials must be sealed properly under argon atmosphere
in a glove box; otherwise, they are evacuated and subsequently pressurized by
hydrogen, nitrogen, argon or a reactive gas. As the transformation properties of
metal hydrides may be sensitive to transition metal dopants, it is recommended
to use milling vials and balls made of catalytically inert and hard silicon nitride or
tungsten carbide instead of using stainless or hardened steel. The ball mill may be
put up in an argon-filled glove box to ensure inert operation conditions.

Milling under argon atmosphere is used for a wide range of applications and
nanocomposites for hydrogen storage have been prepared from:
• different metals (to form a hydrogen-absorbing alloy);

• a metal hydride and a dopant (to alter the kinetic properties of the hydride);

• a metal and a metal hydride (to synthesize complex hydride after subsequent
hydrogen absorption).

Ball milling of a hydrogen-absorbing metal under elevated hydrogen pressure leads


to the formation of brittle hydride which can facilitate the milling process. Ball
milling with other reactive gases, such as NH3, has been used to synthesize hydrogen
storage compounds and composite materials.

> Read full chapter

Sol–Gel Synthesis of Metal Fluorides


G. Scholz, E. Kemnitz, in Modern Synthesis Processes and Reactivity of Fluorinated
Compounds, 2017

4.2 Up- and Downconversion Materials


In the past years, the wet chemical synthesis of rare-earth (RE)-doped calcium
fluoride by precipitation methods has been reported by several groups.7,51,54 The
fluorolytic sol–gel method offers great advantages in contrast to classical precipita-
tion routes since it can be carried out at room temperature in large batches using
standard glass vessels and usually no surfactants are required. Recently, we reported
the sol–gel synthesis of luminescent CaF2 NP sols by the reaction of calcium lactate
dissolved in methanol with anhydrous HF.38 Doping of up to 40 mol% with RE metal
cations can be achieved by addition of lanthanide acetate before fluorination. The
reaction yields fully transparent sols of crystalline RE-doped CaF2 particles with a
size in the lower nanometer range, as seen in Fig. 21.28A. Under illumination with
366 nm radiation, the Eu3+- and Tb3+-doped CaF2 sols appear in the characteristic
colors (Fig. 21.28B) of 5D0 → 7FJ and 5D4 → 7FJ transitions for Eu3+ and Tb3+, re-
spectively. Codoping of CaF2 with Eu and Tb exhibits a different emission spectrum
than the physical mixture of the pure sols which show the expected bright yellow
emission. This suggests that there is an interaction between both rare earth ions in
the co-doped particles that has not been observed for the mixture of singly doped
particles. Apparently, separate doping in two different particles prevents interaction
between two different rare earth metals, and hence, energy transfer from one rare
earth ion to another is not possible.44
Figure 21.28. Transparent Eu3+-and Tb3+-doped CaF2 sols (A) and the same sols
excited with 366 nm (B). From left to right: CaF2:Tb10, mixture of 52% of CaF2:Tb10-
 + 48% of CaF2:Eu10, CaF2:Tb5.2, Eu4.8, and CaF2:Eu10.Reproduced from Ritter, B.;
Krahl, T.; Rurack, K.; Kemnitz, E. Nanoscale CaF2 Doped With Eu3+ and Tb3+ Through
Fluorolytic Sol-Gel Synthesis. J. Mater. Chem. C 2014, 2, 8607–8613 with permission
of The Royal Society of Chemistry.

> Read full chapter

Aerogels as promising materials for en-


vironmental remediation—A broad in-
sight into the environmental pollutants
removal through adsorption and (pho-
to)catalytic processes
Hajar Maleki, Nicola Hüsing, in New Polymer Nanocomposites for Environmental
Remediation, 2018

2.1 Sol-gel process


Almost all aerogels are prepared by a wet chemical synthesis approach, a sol-gel
reaction, but through different starting precursors, as well as operating and pro-
vision requirements [68]. Sol defines a dispersed solution of colloidal primary
particles/monomers that, by themselves, are prepared from a mixture solution
of precursors, water, solvents, and catalysts as a consequence of hydrolysis and
polycondensation reactions. Usually by the addition of a chemical cross-linker or
by changing the physical conditions of the reaction (e.g., pH, temperature) these
colloidal particles connected to form a three-dimensional and interconnected net-
work. It is worth mentioning that the generation of the three-dimensional porous
network is the most critical and determinant aspect of aerogel fabrication. Herein,
the wet gel depending on its pore-filling solvent can be named in different ways.
For example, alcogel and acetogel refer to the gels in which the pore solvent is
exchanged for alcohol and acetone, respectively. The hydrogel refers to a gel that
is usually prepared from natural polymer-based precursors that are subjected to a
physical or chemical gelation in an aqueous solution. During the sol-gel process,
depending on the growth process and rate of cross-linking, network clusters either
being polymeric or colloidal in a size range of 10–100 Å will be obtained [28,68].

Although the chemistry beyond the sol-gel process is different from one type to
another type of gel, Fig. 3 indicates the sol-gel reaction pathway to prepare a silica
(SiO2)-type network as a representative example. Usually, the aerogel formation by
alkoxides can proceed by a single-stage acid or base catalysis or two-stage acid and
then base catalysis [69]. The hydrolysis reaction replaces the OR alkoxides with OH
hydroxyl groups and, subsequently, during polycondensation reactions, the silanol
groups produce siloxane bonds along with alcohol and water as the by-products.
In most cases, condensation initiates simultaneously with hydrolysis reactions and
continues during the whole sol-gel process.

Fig. 3. Sol-gel reaction for an alkoxysilane.

The sol-gel reaction is a versatile process that allows tailoring the gel nanostructure
by adjusting the key reaction parameters. There are a number of parameters that in-
fluence the sol-gel reactions and, therefore, significantly control the nanostructura-
tion of the network and final material properties, namely the concentration of pre-
cursors, relative concentrations of the precursors to the solvent, solvent type, ratio of
water to silica precursors, temperature and pH [70,71]. Also, the sol-gel reaction cre-
ates an opportunity to incorporate an extra phase or a molecular compound to confer
a special feature on the gel network on both molecular and nanoscales. Compositing
in aerogel can be conducted either chemically by using suitable organofunctional
alkoxide derivatives, or physically by introducing additives/dopants in the porous
network. In this regard, the incorporation of hydrophobic moieties (e.g., a methyl
functionality) in the network can improve the stability of aerogels against water
and, integration of polymeric networks increases the weak mechanical strength
of pristine aerogels [72,73]. Other possibilities in this respect are the inclusion of
nanocatalysts or catalytically active elements in the nanostructure of the gel, that is,
addition or dispersion of different essential oxides, noble metals, or titanium oxide
to add additional properties (e.g., a catalytic or photocatalytic functionality to the
filigree aerogel) [74–76].

> Read full chapter

Protected Metal Clusters


Tatsuya Tsukuda, Hannu Häkkinen, in Frontiers of Nanoscience, 2015

1.3.1 Synthesis
Current active research on the protected metal clusters relies on the wet chemical
synthesis. Chapters 2–4Chapter 2Chapter 3Chapter 4 focus on the state-of-the-art
methods of atomically precise synthesis of metal clusters whose surfaces are pro-
tected by monolayers of organic ligands. Key structural parameters for properties
and functions of the protected metal clusters, MnLm (M = metal, L = ligand), are
their chemical compositions (n, m), the nature of the ligands, and mixing mode
of metal elements. In this book, M includes Au, Ag, Cu, and their alloys whereas L
includes thiolates, selenonate, telluronate, phosphines, alkynes, and their mixture.
Chapter 2, by Tsukuda, focuses on atomically precise control of size of metal clusters.
Size-dependent evolution of electronic and geometric structures is demonstrated
as an example. Chapter 3, by Negishi, describes the methods for controlling com-
position of intermetallic clusters and protection of metal clusters by a variety of
organic ligands. The influence of the chemical composition of the metal core and
the bonding mode at the interface on the fundamental properties will be illustrated.
Chapter 4, by Zheng, surveys the synthetic strategies how the morphologies and
metal distributions of heterometallic nanoclusters are manipulated by the proper
choice of surface ligands.

> Read full chapter

Nanomaterials
F. Hubenthal, in Comprehensive Nanoscience and Technology, 2011
1.13.3.3.5 Extensions of photosynthesis
One of the most obvious extensions is to combine two or more of the previously de-
scribed wet chemical synthesis routes [331–333]. For example, Henglein [332] used
a stock solution of gold nanoparticles prepared by a standard chemical reduction
of a gold salt. Henglein emphasizes that the seed particles are without any strongly
bound stabilizers, which seems to be a key point for the stock solution. Later, he
mixed the stock solution with methanol, sodium citrate (as buffer and stabilizer),
and a solution of Au(CN)2. The mixture was irradiated with -rays for several hours
until all gold was reduced. After this process, the size of the gold nanoparticles was
doubled. It is emphasized that the nanoparticle density does not change during the
enlargement. In other words, all reduced gold attaches to the already existing seed
nanoparticles, which allows an effective increase of the nanoparticle size. After the
synthesis, an ion-exchange resin was added to the solution to remove the CN− ions
which had been formed due to the reduction of Au(CN)2. The enlarged nanoparticles
were subsequently used as a new stock solution and the process was repeated. With
this method, Henglein could increase the mean nanoparticle size, step by step, from
R  = 2.7 nm up to R  = 30 nm.

The above-mentioned two- or multistep preparation has also been used to gen-
erate core–shell nanoparticles. For example, Mulvaney et al. [333] prepared silver
seeds ( R  = 3.8 nm) by irradiation of a solution containing AgClO4, 2-propanol,
acetone, and sodium polyphosphate with -rays. Subsequently, this stock solution
was mixed with a KAu(CN)2 solution and again irradiated. Depending on the gold
concentration, the mean nanoparticle size was increased up to R  = 14.4 nm.
In addition, Mulvaney et al. investigated the optical properties of the core–shell
particles as a function of the shell thickness. They observed that, for increasing gold
shell thickness, the silver LSPPR quickly vanishes and a separate gold LSPPR evolves.

One intriguing example of light-induced nanoparticle synthesis has been intro-


duced by Jin et al. [334]. They have shown that conventionally synthesized spherical
silver nanoparticles can be transformed into single-crystal nanoprisms by light-in-
duced morphological changes. In brief, spherical silver nanoparticles were syn-
thesized by injection of a NaBH4 solution in an aqueous solution of AgNO3 and
trisodium citrate. The generated silver nanoparticles with a mean radius of R  =-
 8 nm were subsequently irradiated for 70 h with conventional 40 W fluorescence
light. Following this, the solution turned from initially yellow to blue, indicating
the formation of the nanoprisms. It turned out that the reaction can be initiated
with wavelengths between  = 350 nm and  = 700 nm. The generated nanoprisms
have a uniform thickness of (15.6 ± 1.4) nm and an edge length of 100 nm, with
a standard deviation of  = 15%. A better control of the nanoprisms growth has
been achieved by a dual-beam illumination of the nanoparticle solution and driven
by the excitation of the nanoprisms LSPPR [335]. With one fixed wavelength of
 = 340 nm and variable primary wavelength (from  = 450 nm to  = 700 nm),
triangular nanoparticles with sizes between 30 nm and 120 nm with a narrow size
distribution have been generated. The final nanoparticle size is determined by the
primary wavelengths; the shorter this wavelength, the smaller the nanoprisms. The
reason is that the primary wavelength acts as a barrier that cannot be crossed by the
LSPPR, which effectively hinders further growth of the nanoprisms.

Due to their extraordinary optical properties, the synthesis of triangular-shaped


nanoparticles in solution became more and more attractive in the past few years
and several routes have been found for their synthesis. A review that discusses these
routes has recently been published by Millstone et al. [336].

> Read full chapter

Epitaxial Systems Combining Oxides


and Semiconductors
Gang Niu, ... Bertrand Vilquin, in Molecular Beam Epitaxy (Second Edition), 2018

17.4.1.4 Sol-gel
The sol-gel method appeared shortly before the 20th century as a new method of wet
chemical synthesis of the glass. This technique is based on the simple polymerization
of molecular precursors in solution to obtain glassy materials without heating a
raw material up to the melting temperature. This method has been extended to
various nonglass materials (Mackenzie, 2003). It is particularly well suited to the
production of coatings such as thin oxide layers (Francis, 1999). This method has
the advantages of being inexpensive, perfectly able to control the stoichiometry
and produce materials with high purity and homogeneity. On the other hand, it is
possible to realize nanostructure composites by phase separation, e.g., in the system
silica-zirconia (Gaudon et al., 2005).

The sol-gel process is to achieve a stable solution (i.e., sol) containing the molecular
precursors and to initiate hydrolysis-condensation reactions to convert this sol into
a solid state network (i.e., the gel) in which resides the initial solvent. There are two
main types of precursors: metal salts or alkoxides. The aggregation or polymerization
of these precursors leads to the formation of the gel.

There are two main processes to realize films from the sol-gel method: centrifuga-
tion (spin coating), which consists of depositing droplets of precursors (sol) on a
rotating substrate and dipping (dip-coating) the substrate in the sol.
The realization of ferroelectric films by sol-gel technique is a succession of steeps
of spreading of the sol-gel and thermal treatments. The sol is usually spread over
a substrate by spin coating. The thickness of deposited film depends primarily on
the substrate rotation speed and on the sol physicochemical properties (viscosity,
concentration, evaporation rate). After spin coating, the sample is dried to evaporate
some solvent and get a solid deposit. The layer is then put on a hot plate to perform
calcinations (or pyrolysis) to remove organic residues. This step is crucial because it
determines significant structural quality of the ferroelectric film. The calcination
temperature must be high enough to remove all organic residues (between 300°C
and 400°C). Otherwise, trapped organic materials inside the film may lead to an
increase in its porosity. However, a too high temperature risks encouraging the
presence of a parasitic phase, as the pyrochlore phase in the case of PZT. Moreover,
the calcination temperature could also influence the crystallographic orientation
of PZT films (Gong et al., 2004). After the pyrolysis step, the amorphous layer is
crystallized and densified at a temperature between 600°C and 800°C. This step is the
crystallization usually performed by rapid thermal annealing (RTA). RTA is selected
rather than annealing in a conventional furnace to limit the thermal budget. Its
main advantage is that the thermal power is transmitted by radiation and not by
convection, allowing to extremely fast temperature increase up to 300°C/s by using
of halogen lamps. Then the annealing proceeds within a few tens of seconds. As the
thickness of a layer after crystallization is usually 70 nm, according to the desired
final thickness all the steps described above have to be repeated several times.

The sol-gel process has few entered now in mainstream production. However, it
tends to be more and more used. Especially since the best piezoelectric properties of
a PZT film were obtained with a film prepared by sol-gel (Calame and Muralt, 2007).
In comparison with the deposition system previously described, this technique is
easy to perform since it just needs a spinner and a furnace and allows deposition on
a large surface. Moreover, complex compounds can be achieved by playing on the
composition of the precursor’s mix. The main disadvantage is the thickness of the
layer, which can be proceed in each time, about 70 nm. As a consequence, for a thick
film of 1 µm, full process has to be performed several times.

As these various deposition methods are the most used for complex oxide growth
and have shown successful results for perovskite film epitaxy, they are used to
investigate the growth of functional complex oxides, such as SrTiO3 or PZT, on
silicon. For example, the growth of SrTiO3 on Si/SiO2 (~1 nm) substrate was realized
by MOCVD at 700°C in order to achieve a high crystalline quality and to avoid the
formation of strontium carbonates (Dubourdieu et al., 2005). However, at such a
high temperature, interface reactions may occur, and a silicate layer containing Sr
and Ti was evidenced by X-ray photoelectron spectroscopy.
Even at a lower deposition temperature, such as 500°C, a PZT film grown on
silicon wafer can easily react with Si to form silicate, and interdiffusion of Pb and
Si occurs at the PZT/Si interface, avoiding any possible application for such type of
heterostructure (Shichi et al., 1994).

Contrary to the MBE process, because of the process high temperature or oxygen
pressure and the use of reactive chemical components, these deposition methods
lead to the formation of polycrystalline and/or an interfacial parasitic phase. One of
the best ways to integrate complex functional oxides is to combine MBE and a sec-
ond deposition technique. On the one hand, as previously shown, high quality STO
film can be achieved on silicon thanks to MBE with abrupt interface. On the other
hand, PLD, sputtering, MOCVD, and sol-gel have become important techniques,
which can be used advantageously to produce high quality multicomponents’ thin
films, including ferroelectrics and high-temperature superconductors. With this
approach, SrTiO3 on Si was used as a buffer layer for the epitaxy of La0.7Sr0.3MnO3
(LSMO) (Dubourdieu et al., 2010). LSMO films are ferromagnetic and metallic at
room temperature. The epitaxial growth of complex oxides on Si wafers opens
up the route to the integration of a wide variety of functionalities and devices in
nanoelectronics.

> Read full chapter

Metal-semiconductor core–shell
nanostructured photocatalysts for envi-
ronmental applications and their recy-
cling process
Kunal Mondal, Pallabi Moitra, in Metal Semiconductor Core-Shell Nanostructures
for Energy and Environmental Applications, 2017

6.4 Metal–metal oxide core–shell nanoparticle synthesis


Usually, metal-semiconductor core–shell nanostructured photocatalyst particles are
produced by means of wet chemical synthesis approaches. A common fabrication
scheme includes the synthesis of seed nanocrystals of the core material and using
them as a template to grow up the shell material on them [55–59]. The width of the
metal oxide semiconductor shell is generally determined by the quantity of precursor
material introduced into the synthesis scheme. However, this is a bit challenging to
control the thickness of the shell.
Current progresses in colloidal nanoparticle production permitted manageable
preparation of a range of hybrid metal-semiconductor nanoparticles, for example
metal-ornamented anisotropic semiconductor nanoparticles, nanoparticle dumb-
bell-like trimers and heterodimers, and core–shell heterostructures, with several
elemental arrangements [60–62]. A chain of fascinating phenomena rising from
the nanoscopic interfaces between the metal and metal oxide semiconductor con-
stituents have been perceived in these hybrid nanoheterostructured materials. It has
been described that the existence of metallic constituents may considerably rise the
photocatalytic and light-gathering efficacies either by refining the charge carrier
separation at the metal–metal oxide semiconductor-interfaces or by improving the
light absorption properties [63–65]. Furthermore, metal nanoparticles can greatly
modify the photocatalytic and photoluminescence performances and alter the non-
linear optical responses of the metal oxide constituents [66]. Also, the attendance of
a semiconductor counterpart adjacent to a metal nanostructured material can also
lead stimulating reforms in plasmonic properties of the hybrid nanostructures over
and done with plasmon–exciton coupling or modification of the native dielectric sur-
roundings. In the recent past, Wang’s group has established a healthy wet chemistry
method, which includes the manageable growth of a nanostructured polycrystalline
Cu2O metal oxide shell enclosing a core of gold nanoparticle [29].

Also, electrochemical route of metal-semiconductor nanoscale particle synthesis


is attractive due to various reasons, including the capability to consume aqueous
solvent instead of organic solvents, deposition at room-temperature, low cost,
and specific control of conformation and thickness of coating of semiconductor
on metal surfaces. In this context, Shannon and coworkers have been reported
for the first time the synthesis of metal-semiconductor core–shell nanoparticle
using codeposition and electrochemical atomic layer deposition approach [59]. Ting
has proposed a synthesis method for nanostructured gold-polyaniline core–shell
particles via electrochemical polymerization of aniline through 4-aminothiophenol
capped Au nanoparticles [67].

Hydrothermal synthesis is another interesting bottom-up way of fabricating met-


al-metal oxide core–shell nanostructures since this process has a huge advantage
over the selection of precursor materials and this can be done even in low tempera-
ture and mild reaction condition. It is worth mentioning that Pal’s group synthesized
a binary metal–metaloxide semiconductor Zn–ZnO core–shell nanorod structure
through an ethylenediamine-mediated low-temperature hydrothermal process and
also they proposed that their process can be extended for creating other hybrid
nanostructures [68].

Interestingly, in case of the noble metal–metal oxides core–shell nanoparticle es-


tablishment needs control over nucleation and growth mechanisms of the shell
formation on top of the seeds of the metals. These processes are heavily governed
by the lattice mismatch between core and shell materials. The other various para-
meters, such as temperature, concentration, interfacial energy, and surfactant are
also responsible. If there occur a lattice mismatch between constituent materials,
then it is challenging to produce their core–shell conformation. However, the use of
ligand and surfactant is very effective to adjust their interfaces. The synthesis of noble
metal–silica core–shell nanostructures are quite familiar, though, this approach
has not been comprehensively used to other oxides, since they have problems of
aggregation in solution while using metal salts as precursors [69]. On the other
hand, if the usage of metal salts is inevitable, the inclusion of appropriate surfactant
or ligand can be the finest approach to prepare core–shell structures. Unfortunately,
there are very few synthesis reports of metal–metal oxide core–shell nanostructures
reported as the choice of a friendly surfactant and ligand is very challenging.

The fabrication of the distinct TiO2 shell on metal core was first described by Li
et al. [70] The Au-TiO2 core–shell nanoparticles of 500–800 nm were produced in
two steps, which involved the deposition of 200–500 nm TiO2 shell on prefabricated
gold nanoparticle seed of 50–150 nm by hydrothermal technique. They have used
ascorbic acid and Cetyl trimethylammonium bromide (CTAB) which facilitated the
creation of anatase TiO2 shells and Au cores, respectively. Hollow Au-TiO2 core–shell
nanostructure has also been designed via Ostwald ripening which includes removal
of solid central TiO2. Furthermore, the core size of the Au nanoparticles was also
adjusted between 150–250 nm ranges using the reduction of gold in HAuCl4-satu-
rated solution with Au-TiO2 nanostructures. This involves slow interfacial diffusion
of HAuCl4 inside the Au-TiO2 along with the CTAB and ascorbic acid solutions
outer of the nanoscopic core–shell particles [70]. Also, Wang’s group has developed
a method to fabricate ZnO/NiO photosensor where they deposited metallic Ni
nanoshells onto the ZnO nanorods surface and after that annealed to transform
ZnO/Ni composite into ZnO/NiO core–shell nanorod arrays [71]. An innovative
synthetic direction for the preparation of Ag-ZnO hybrid core–shell nanoparticles
in precise conditions, consuming silver nitrate and zinc acetate precursor salts in
N, N-dimethylformamide solvent has been reported recently by Aguirre et al. [72]
The solvent concurrently acts as a reducing mediator for Ag+ ions and delivers
the easy medium for the hydrolysis of zinc acetate at room temperature, without
any stabilizers or additives. Also, Zhao and coworkers have produced core–shell
nanomorphology of Ag-ZnO metal–metaloxide by using excimer laser ablation in
liquid [73].

In another report, a nonorganic synthetic technique has been established to produce


the uniform pomegranate-like Pt-CeO2 multicore–shell nanospheres in a big scale
by Wang et al. [74] Under the argon atmosphere the redox reaction simply occurred
between Ce(NO3)3 and K2PtCl4 in an alkaline aqueous solution, in which no other
reducing agents or surfactants were added. The as achieved nanospheres showed
outstanding structure, stability even being heated at high temperature. Besides, the
obtained Pt-CeO2 multicore–shell nanospheres can be further reinforced on reduced
graphene oxide to form heterogeneous nanocatalyst, which has been effectively use-
ful in the chemical reduction reaction of nitrophenol by ammonia borane (NH3BH3)
as an alternative of using hazardous H2 or NaBH4 [74].

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like