You are on page 1of 9

Q118

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)
2162-8769/2013/2(7)/Q118/9/$31.00 The Electrochemical Society

A Review of Synthesis Techniques for Gallium-Zinc Oxynitride


Solar-Activated Photocatalyst for Water Splitting
Babak Adeli and Fariborz Taghipourz
Department of Chemical and Biological Engineering, University of British Columbia, Vancouver, BC, V6T 1Z3, Canada
Converting a practically limitless energy source, such as sunlight, into chemical energy with very little or no carbon footprint will
pose a major challenge in the coming decades. The technology exists to convert solar energy into chemical fuels, such as hydrogen,
through overall water splitting to produce hydrogen and oxygen. However, the photocatalytic efficiency is still below the feasibility
limit. Many photocatalyst materials have been developed for solar hydrogen generation through water splitting in the last four
decades. Gallium-zinc oxynitride (GaN:ZnO) solid solution is reported to be a suitable photocatalyst for overall water splitting and
to have the highest photocatalytic activity. The traditional synthesis method contains difficulties and inefficiencies, such as long
nitridation of starting materials at high temperatures and low Zn content of the synthesized photocatalyst. A number of experiments
have been conducted in recent years to develop new synthesis approaches. In this article, a comprehensive review of various synthesis
techniques of GaN:ZnO solid solution, along with their advantages and disadvantages, are presented. This information is essential
for improving the efficiency of the synthesis techniques of GaN:ZnO solid solution and its photocatalytic activity for overall water
splitting.
2013 The Electrochemical Society. [DOI: 10.1149/2.022307jss] All rights reserved.
Manuscript submitted March 26, 2013; revised manuscript received May 22, 2013. Published June 5, 2013.

Providing clean and renewable energy is arguably the most important challenge facing humanity in the 21st century.1 The worlds
energy demand is projected to more than double by mid-century and
more than triple by 2100.2 Holding the atmospheric CO2 levels to
their current values by mid-century would require the invention and
development of new technologies.3
Among the various clean and renewable energy sources, sunlight
is the largest. More energy from sunlight strikes the Earth in one hour
(4.3 1020 J) than all the energy consumed on the planet in a year
(4.1 1020 J).2 However, the use of sunlight as a clean energy source
depends on the capturing, conversion, storage and distribution of solar
energy.
There are three different ways to use solar energy: in the form
of heat (thermochemical), fuel (photo-electrochemical) or electricity
(electrolysis, and photovoltaics). Despite the enormous energy flux
supplied by the sunlight to the Earth, the contribution of all three conversion routes previously mentioned is only a fraction of our current
needs. The conversion of solar energy to usable forms of energy is
currently inefficient and expensive. Solar electricity generates 5 to
10 times that of the cost of electricity obtained from fossil fuels but
provides only 0.015% of the worlds current electricity demand. Solar
fuel, mainly biomass, supplies 11% of the worlds fuel needs, and solar
heat provides 0.3% of the energy used for heating spaces and water.1
The most attractive method of solar energy conversion and storage is
in the form of the chemical bond of an energy carrier, which does not
experience energy loss via thermal transformation (thermochemical)
and conversion of solar energy to electricity (electrolysis).4 The direct
production of fuel from sunlight is advantageous because it inherently
provides a method for extracting energy during the evening and for
dispatching and distributing energy costs effectively in the existing
infrastructure for use in the residential, industrial and transportation
sectors.5 Solar hydrogen makes solar energy as storable and transportable as fossil fuels without their negative environmental impacts.6
The conversion of sunlight into chemical fuels, through water splitting into hydrogen and oxygen, is an existing technology; however, the
current rate of solar water-splitting reaction is very low. Also, the photocatalysts that have been studied cannot readily provide the energy
levels that would be needed to support a terawatt-level implementation
of solar electricity use for hydrogen production.7
Thermodynamically, the overall water-splitting reaction is an uphill reaction with a large positive change in Gibbs free energy
(G = 238 kJ/mol).
2H2 O 2H2 + O2

E-mail: fariborz@chbe.ubc.ca

[1]

The reduction and oxidation half-reactions are described as


follows:
4H+ + 4e 2H2

[2a]

2H2 O + 4h+ O2 + 4H+

[2b]

When the change in Gibbs free energy is considered, photocatalytic water splitting is distinguished from photocatalytic degradation reactions, such as photo-oxidation of organic compounds using
oxygen molecules, that are generally downhill reactions. This photocatalytic degradation reaction is regarded as a photo-induced reaction and has been extensively studied using titanium dioxide, TiO2 ,
photocatalysts.8,9
In principle, overall water splitting can be achieved with both
visible and near-infrared light because the difference between the
potentials of the hydrogen and oxygen evolution half-reactions (Eq. 2a
and 2b) is only 1.23 eV that corresponds to the light with a wavelength
of approximately 1000 nm.6,8
Figure 1 shows a schematic illustration of the overall watersplitting reaction. Under irradiation at an energy level greater than
the bandgap of the semiconductor photocatalyst, electrons in the valence band (VB) are excited and migrate to the conduction band (CB),
leaving holes in the VB (Figure 1a). To achieve overall water splitting,
the semiconductor must have suitable band-edge positions. The CB
minimum energy level must be positioned at a more negative potential
than the reduction potential of H+ to H2 (0 eV vs. NHE). Furthermore,
the VB maximum must be located more positively than the oxidation
potential of water to oxygen (1.23 eV vs. NHE). Therefore, the overall water splitting can ideally be achieved by using the entire spectral
range of visible light (up to 1000 nm). However, there is an activation barrier in the charge-transfer process between photocatalysts
and water molecules that requires photon energy levels greater than
the bandgap of the photocatalyst, and band edge, with the over potential of water reduction/oxidation half-reactions (reactions 2a and
2b respectively) to drive the overall water-splitting reaction with reasonable reaction rates. Moreover, the formation of water molecules
from evolved hydrogen and oxygen, so called the backward reaction,
and, more importantly, the recombination of photogenerated electrons/holes must be strictly inhibited.
Fujishima and Honda initially triggered research in this field
back to 1972. They have reported hydrogen generation in a photoelectrochemical water-splitting cell using a TiO2 electrode as the photoanode (for water oxidation half-reaction) and a platinum as a cathode electrode (for water reduction half-reaction).10 Since then, many
semiconducting materials have been reported to be active for the photocatalytic water splitting. However, only a few of them demonstrated
stable and reproducible overall water splitting under visible-light

Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)

Q119

Table I. Unit cell parameters (a, b, and c) of wurtzite crystal


structures of GaN and ZnO.30,31

Figure 1. The principle of photocatalytic overall water splitting: (a) semiconductor excitation and migration of electron from VB to CB; (b) processes
occur in photocatalysts particle after light irradiation, including photoexcited
electrons/holes recombination and migration to the photocatalysts surface.

irradiation. The major obstacle in developing the desired photocatalyst is the lack of a suitable compound that meets the following
requirements: (I) a bandgap narrower than 3 eV to harvest the visible part of the solar spectrum, (II) a band-edge position suitable
for driving the overall water splitting reaction, (III) a highly crystalline structure to inhibit the photoexcited electrons/holes recombination and (IV) a stable chemical state in photoreaction condition.7
Figure 2 shows some of the photocatalyst materials previously studied for water splitting and the difficulties found in applying them for
overall water-splitting purposes.
Gallium-zinc oxynitride solid solution (GaN:ZnO; also known as
(Ga1x Znx ) (N1x Ox ), where x is the fraction of ZnO in GaN:ZnO
solid solutions crystal structure) is one of the materials that meets
all the previously stated conditions with highest photocatalytic activity for overall water splitting reported so far.1113 The efficiency
of the GaN:ZnO solid solution for overall water splitting has been
claimed to be the highest by Maeda et al.;14 which has also been
confirmed by several other articles1518 on this field. GaN:ZnO loaded
with Rh Cr2 O3 has been reported to achieve a hydrogen evolution
rate of 3.09 mmol/h.g,19 highest among other visible-light activated
photocatalysts reported in several review articles.8,20,21 It should be
noted that GaN:ZnO solid solution is among a few known photo-

Material

a = b (nm)

c (nm)

GaN
ZnO

0.319
0.325

0.519
0.521

catalysts that are capable of driving overall water splitting reaction


efficiently without using any sacrificial reagents.
GaN is a well-known material with a bandgap of 3.4 eV and has
been studied extensively for application in light-emitting diodes and
laser diodes.22,23 GaN has also been examined as a photoelectrode, and
it has been confirmed to have the potential for overall water splitting
under UV irradiation.24,25 The bandgap of ZnO is 3.23.4 eV, and
it has been studied for many applications such as in light-emitting
diodes and gas sensors.2629 GaN and ZnO have the wurtzite crystal
structures with almost the same lattice parameters; therefore, a solid
solution with the same crystal structure can be formed between the
two. Table I shows the crystal lattice parameters of ZnO and GaN.
The GaN:ZnO solid solution photocatalyst is typically synthesized
by nitridation of a mixture of Ga2 O3 and ZnO at high temperatures for
515 h13 via the solid-state reaction process, which is the traditional
method. Although the photocatalyst prepared through the traditional
method demonstrates the highest activity for overall water splitting,
the long synthesis time at high temperatures is considered a drawback
of this synthesis technique. Moreover, the Zn/Ga ratio of the solid solution is very low, even with low synthesis times (<0.28 at 5 h)13 due
to the volatilization of ZnO at high synthesis temperatures. Considering that the visible-light harvesting of the GaN:ZnO solid solution
photocatalyst is attributed to the Zn content of the prepared photocatalyst, preparing more efficient catalysts demands the development
of new synthesis strategies that result in higher Zn content. Various
attempts have been made to increase the Zn content without losing the
crystallinity of the synthesized photocatalyst, including the application of different precursors in different amounts, different annealing
times, as well as pre- and post-treatment techniques.
In addition to the traditional method of GaN:ZnO solid solution
synthesis previously described,11,12 there have been research reports
in the last few years concerning alternative synthesis routes. However, there is no review article published in the literature regarding
the various preparation techniques and their capabilities and limitations. Such information is essential for developing efficient visiblelight-driven photocatalysts for overall water splitting by emphasizing
the advantages and excluding the disadvantages. In this article, a

Figure 2. Band-gap structure of some semiconductors and their limitations for water splitting.5,8
Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Q120

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)

comprehensive review of different synthesis methods of GaN:ZnO


solid solution and the benefits and shortcomings of each method are
presented. In order to overcome the drawbacks of the traditional synthesis method, different strategies such as using mesoporous precursors, alternative nitrogen sources, preparation of the precursor in the
form of layered double hydroxides (LDHs) and different heating techniques have been addressed.
High Temperature Solid State Reaction
The solid-state reaction technique is the most widely used method
for synthesis of crystalline solids from a mixture of solid-starting
materials. Solids do not react at room temperature during normal time
scales, but solid-state reactions perform at high temperatures.
Domen and colleagues11 reported solid-state synthesis of
GaN:ZnO solid solution for the first time in 2005. Although the
method was proposed just a few years ago, it is known as the traditional method of GaN:ZnO solid solution photocatalyst preparation
because various research has been conducted since 2005 on the photocatalyst synthesis, characterization and application of this technique.
In this method, the atmospheric nitridation of a mixture of Ga2 O3
and ZnO powders under an ammonia (NH3 ) flow (100500 mL/min
flow rate) of 1123 K for 520 h results in a yellow powder,32 as
described in Reaction 3. At first sight, the yellow color of the prepared
photocatalyst indicates its visible-light absorption.32,33 The product
of the 5- to 20-h nitridation process is cooled to room temperature,
maintaining the ammonia flow throughout.11,13
NH3

Ga2 O3 + ZnO (Ga1x Znx ) (N1x Ox )

[3]

15

Chen et al. monitored the formation of GaN:ZnO solid solution during the solid-state reaction of ammonia with the Ga2 O3 /ZnO
mixture through an in situ time-resolved X-ray diffraction (XRD).
Figure 3 shows the different possible routes for the traditional solidstate reaction.
Considering the large band-gap energies of GaN and ZnO (3.4 eV
and 3.2 eV, respectively), it is assumed that any solid solution forms
from these materials should have a bandgap larger than 3.0 eV. However, making the solid solution shifts the VB maximum upward without affecting the CB minimum, as can be seen in Figure 4.
Similar results were reported by Lee et al.34 for zinc germanium
oxynitride solid solution (GeN:ZnO), which was synthesized through
the solid-state reaction, as well. It has been reported that p-d repulsion in II-VI semiconductors shifts the VB maximum upward without
affecting the CB minimum.35 Lee et al.34 suggested that the presence
of Zn3d and N2p (or O2p) electrons in the upper VB of GeN:ZnO
provides p-d repulsion and shifts the top of the VB, thus, narrowing the bandgap.34 Shi et al.36 also reported that the VB maximum
changes due to the p-d repulsion caused by the Zn3d orbital. Similarly, it can be hypothesized that p-d repulsion in the GaN:ZnO solid
solution may cause the formation of the VB maximum by N2p or

Figure 3. Different possible routes of traditional solid-state reaction with


Ga2 O3 /ZnO as starting materials.

Figure 4. Schematic illustration of band structures of GaN, ZnO and their


solid solution.

O2p atomic orbitals to shift it to the higher potential energy.37 The


previous discussion concluded that the bandgap of the synthesized
solid solution decreased with the increase of the Zn content in the
crystalline structure of the photocatalyst.
The XRD pattern in Figure 5 confirms that the resulting photocatalyst is a highly crystalline single-phase wurtzite structure material, similar to that of GaN and ZnO (for nitridation time 530 h;
Figure 5a). The positions of the diffraction peaks are shifted successively to lower angles with increasing Zn and O concentrations
(Figure 5b), confirming that the prepared samples are not physical mixtures of GaN and ZnO phases but rather their solid solutions. The same conclusion can be made with UV-vis spectroscopy

Figure 5. XRD patterns of samples prepared through the traditional method.


(a) Different nitridation time.32 Reprinted with permission from Chem. Mater.
2010, 22, 612623. Copyright 2010 American Chemical Society. (b) Various
Zn compositions.37 Reprinted with permission from J. Phys. Chem. C 2007,
111, 78517861. Copyright 2007 American Chemical Society.

Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)
for shifts in the absorption edge of the prepared photocatalyst to
longer wavelengths with increasing Zn and O concentrations (x) in
(Ga1x Znx ) (N1x Ox ) and to shorter wavelengths with increasing nitridation times. The band-gap energies of the solid solutions are calculated to be 2.62.8 eV based on the diffuse reflectance spectra.32 Considering the bandgap energy and band edge potentials of GaN:ZnO
solid solution (Figure 4), visible light degradation of toxic organic
pollutants (e.g. trichlorophenol, 2,4-dichlorophenol, and sodium benzoate, which have no absorption in the visible region)38 can be another
potential application of this photocatalyst.
The GaN:ZnO solid solution photocatalyst does not show any
activity for overall water splitting without co-catalyst loading.39,40
Typically, nanoparticles of co-catalysts deposited on the surface of
photocatalysts, as the hydrogen production sites, are used to enhance
the rate of water reduction half-reaction. The main role of co-catalysts
is to steal the photoexcited electrons from the bulk of the photocatalyst
and bring them to the water-photocatalyst interface. Noble metals or
transition-metal oxides (e.g., Pt or Rh) are often used as co-catalysts
to facilitate the water reduction half-reaction. Such co-catalysts are
typically applied as nanoparticles to the photocatalyst surface by different methods including in situ photodeposition. In situ photodeposition allows the co-catalysts to be located selectively at reaction sites
without the need for an activation treatment.39 The rate of hydrogen
evolution under visible light irradiation ( > 400 nm) for GaN:ZnO
solid solution was initially reported to be about 180 mol/h.g using RuO2 as co-catalyst.11 The evolution rate increased to about
900 mol/h.g,14 due to utilizing core/shell co-catalyst structure, and
later to 3.09 mmol/h.g (A.Q.Y = 5.9%,19 ) as a result of photocatalyst
post-calcination. Maeda et al. have also investigated the long-term activity and regeneration of solid solution photocatalysts and reported
that the rate of visible-light water splitting remained unchanged for
3 months (2160 h), producing hydrogen and oxygen continuously at
a stoichiometric rate.41

Medium Temperature Solid State Reaction


Zinc oxide inside the crystalline structure of GaN:ZnO solid
solution is responsible for the visible-light harvesting of the
photocatalyst.15,17,42,43 Although controlling the size distribution of
ZnO before solid-state reaction with pre-heat treatment at 823 K resulted in higher Zn content in the synthesized solid solution using
the traditional route,42 the Zn/Ga ratio in the resulting solid solution
was still very low (<0.28). A small Zn content indicates that most
of the ZnO in the starting material is reduced and volatilized because
of exposure to the reductive atmosphere during nitridation at high
temperatures.15 Various experiments, including using excess amounts
of ZnO in starting materials and shorter synthesis times, have been
performed to attempt to increase the Zn content of the photocata-

Q121

lyst; however, the Zn/Ga ratio obtained from the nitridation reaction
remained low. Considering that a low Zn/Ga ratio is attributed to
the high synthesis temperature and reductive atmosphere, developing
new synthesis routes with lower nitridation temperatures and times or
a different nitrogen source other than ammonia is crucial.
Recently, Yan et al.44,45 proposed a two-step medium-temperature
solid-state synthesis method via nitridation of mesoporous nanocrystalline zinc gallate (ZnGa2 O4 ) under a flow of ammonia (250 mL/min)
for 6 h at 680 C. Using this method, the ion-exchange reaction of
the mesoporous NaGaO2 colloid precursor and Zn acetate resulted
in hydrothermally treated mesoporous ZnGa2 O4 .45 Reaction 4 and
Figure 6 show the formation of a mesoporous colloidal zinc gallate
through the ion-exchange reaction of mesoporous NaGaO2 .
2NaGaO2 + Zn (CH3 COO) ZnGa2 O4 +2CH3 COONa

[4]

The hydrothermal treatment caused improvement in the stability


of the mesoporous structure of zinc gallate and was effective in the
formation of the GaN:ZnO solid solution with the desired mesoporous
structure, higher surface area and Zn content.
Mesoporous Zn gallate as a starting material has a higher surface area for solid-state reactions. As a result, less nitridation time is
needed for the solid-solid diffusion step in solid-state reactions. Moreover, more uniformity, higher Zn/Ga ratios and greater visible-light
harvesting are expected from the solid solution prepared through the
medium-temperature solid-state technique as compared with the traditional method. Hisatomi et al.42 claimed that small ZnO particle-size
distribution and higher surface area enhanced the volatilization of ZnO
in the solid-state reaction. Therefore, by controlling the ZnO particlesize distribution in the starting materials mixture, a high crystalline
GaN:ZnO solid solution might be achievable at lower nitridation times
and temperatures.
Decreasing the rate of recombination of photoexcited electrons/holes is the key factor for developing an efficient photocatalyst
because these charges are the basis of the photocatalytic activity. This
can be achieved by reducing the density of crystal defects and the
synthesis of a highly crystalline photocatalyst.8,4648 Post-calcination
or longer synthesis times at mild temperatures would modify the crystallinity of the resulting photocatalyst. Medium-temperature solidstate reactions offer the benefit of lower nitridation times and
temperatures compared with the traditional method; however, the
crystallinity of the resulting photocatalyst might be lower than the
one obtained from the traditional method as shown by the powder
XRD results.17
There is no report in the literature regarding the application of
the solid solution photocatalyst synthesized through the mediumtemperature solid-state reaction technique for overall water splitting.
Yan et al.44 claimed that the photocatalysts synthesized by this method
had a VB edge potential of 1.511.73 eV, with band-gap energy of

Figure 6. Formation of a mesoporous colloidal Zn gallate and ion-exchange reaction of mesoporous.45 Reprinted with permission from Angew. Chem. Int. Ed.
2010, 49, 64006404. Copyright 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Q122

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)

2.342.57 eV. Thermodynamically, the photocatalyst should be able


to drive the overall water-splitting reaction and produce hydrogen
and oxygen. However, making comments about the photocatalytic activity of the photocatalyst for hydrogen generation solely based on
its band-gap energy and band-edge potential is not easy or reliable.
Because overall water-splitting reaction occurs at the surface of the
photocatalyst, the higher surface area of the resulting mesoporous
GaN:ZnO solid solution photocatalyst may explain its higher activity
than that seen with the traditional method; although, the low crystalline structure of the prepared photocatalyst may affect its activity
dramatically. According to Maeda et al.,19 the post-calcination of the
GaN:ZnO solid solution at 823 K for 1 h improved its photocatalytic
activity for overall water splitting to twice that seen before calcination. Low crystalline structure, caused by crystal defects, is referred
to any type of deviation from ideal crystal structures where the atoms
are arranged in a periodic, regularly repeated three-dimensional pattern. The presence of impurities in starting materials can also impact
the crystalline structure during each preparation technique. Semiconductors, as well as other solid-state materials, may contain a variety
of defects, which are introduced in the material during preparation
process (e.g. pre-treatment, synthesis, and post-treatment). Structural
defects can be classified as (i) point defects, such as substitutional
and interstitial impurity atoms and vacancies, (ii) one-dimensional or
line defects, such as dislocations, (iii) two-dimensional or planar defects, such as surfaces, grain boundaries and stacking faults, and (iv)
three-dimensional or volume defects, such as voids and inclusions.49

Layered Double Hydroxides Precursor


Layered double hydroxides (LDHs) were discovered in the mid19th century50 and have been extensively studied for many applications. The basic features of their structure, including positively
charged brucite-like layers, charge-balancing anions and water in interlayer galleries, are well documented.51 The interlayer bonding of
LHDs is not strong; therefore, LHDs demonstrate unique properties,
such as excellent expandability. Because of their ease of synthesis,
LHD application as a cost-effective, durable, versatile and potentially recyclable source of a variety of host materials, catalyst supports and catalyst precursors with the desired chemical and physical
features have been investigated over the past few years.52 In particular, mixed metal oxides prepared by precipitation in the form
of LDHs with large specific surface areas (100300 m2 /g;)52 have
been widely studied. LDHs are synthetic anionic materials represented as [(M2+ )1x (M3+ )x (OH)2 ]x+ (An1 )x/n .yH2 O. Their structure
contains brucite-like layers in which divalent cations (M2+ ) are substituted by trivalent ions (M3+ ), resulting in a positively charged layered
structure. The positive charges of the layers are compensated by interlayer anions. The characteristic of positively charged layers, including
the divalent and trivalent cations (M2+ and M3+ , respectively) and the
interlayer anion (An ), as well as their relative composition (x), can
be varied over a wide range.16 Figure 7 is a schematic illustration of
LDHs structure.
Studies have provided examples of LDHs using Zn as a divalent cation with other metals such as Al, Fe, Cr and Co as trivalent cations, with different interlayer anions and compositions,5355
but not Ga. However, the structural properties of LDHs suggest the
possibility of applying LDHs
 as a host to prepare the zinc gallium
mix oxide Zn2+ /Ga3+ /Co3 precursors for GaN:ZnO solid solutions
photocatalyst.16,50
Wang et al.16 proposed a new synthesis method for GaN:ZnO photocatalyst using LDHs as precursors. Typically, there are four general
techniques for LDHs preparation: (I) anion-exchange of a precursor LDHs; (II) direct synthesis by co-precipitation; (III) rehydration
of a calcined LDH precursor and (IV) thermal reaction.56 Wang et
al.16 used the co-precipitation technique to prepare zinc gallium mix
oxide LDHs precursor and used the LDHs precursor in further nitridation process to synthesize the GaN:ZnO solid solution photocatalyst.
The proposed synthesis method is more efficient than the high- and

Figure 7. Schematic illustration of LDHs structure. M2+ , M3+ and An represent divalent cation, trivalent cation, and interlayer anion, respectively.

medium-temperature solid-state reactions because of its lower nitridation time (30 min). Similar to the traditional method, Ga2 O3 and
ZnO have been used as precursors in this two-step synthesis method.
Zn/Ga/CO3 LDHs were obtained by 12- to 24-h aging of Zn (NO3 )2
and Ga (NO3 )3 solutions. Zn (NO3 )2 and Ga (NO3 )3 mother liquor was
prepared by dissolving a stoichiometric amount of Ga2 O3 and ZnO in
a HNO3 , NaOH and Na2 CO3 solution at pH 8. Further nitridation of
Zn/Ga/CO3 LDHs at 800 C under an ammonia flow (300 mL/min)
for 30 min resulted in the GaN:ZnO solid solution photocatalyst.16
Decreasing the nitridation time to 30 min (from the original 15 h for
the traditional method) might be mainly attributed to the penetration
of NH3 into the interlayer regions of LDHs and increasing the surface
area of reaction or the formation of an homogeneous mixture of Zn2+
and Ga3+ at an atomic scale in LDHs structure (Figure 8).
According to the energy-dispersive X-ray analysis reported by
Maeda et al.,32 one of the drawbacks of the GaN:ZnO solid solution prepared via a traditional solid-state reaction was the nonhomogeneous bulk composition. Apparently, the solid solution prepared
through the LDHs precursor is uniformly synthesized because the
Zn2+ and Ga3+ ions are homogeneously mixed in the precursor. Higher
Zn/Ga ratios (up to 2.6) reported by Wang et al.16 might be attributed
to the stabilizing of Zn inside the interlayer region of LDHs precursor;
this resulted in decreased Zn volatilization, leading to the synthesis of
a Zn-rich photocatalyst. The stability of Zn inside the resulted solid

Figure 8. Schematic illustration of GaN:ZnO solid solution using LDHs precursor.

Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)

Q123

solution structure was confirmed through elemental analysis presented


by Wang et al. They reported that the composition of Zn remained
almost unchanged after long nitridation times (0.55 h).16
The visible-light activity of the solid solution synthesized via
LDHs precursor for overall water splitting has not been reported in the
literature. The bandgap of the resulted solid solution was calculated
as 2.60 eV at x = 0.46 and was gradually reduced to 2.37 eV by
increasing the Zn content (x) to 0.81.16 Considering the p-d repulsion
in the GaN:ZnO solid solution, lifting the top of the VB maximum
upward without affecting the CB minimum, the solid solution synthesized using LDHs precursor should be able to split water into oxygen
and hydrogen. The narrower bandgap is attributed to the higher Zn
content of the photocatalyst. Although increasing the Zn content of the
solid solution would increase the visible-light activity of the photocatalyst, it would also increase the density of the crystal defect, which
is responsible for photoexcited charges recombination.
Microwave Heating
Microwave heating, as a new energy source, seems to be a popular
and useful technology for endothermic reactions in organic chemistry.
Usually, for endothermic reactions, the reactant is heated up using the
heat supplied from an external energy source through a conduction
or convection heat-transfer mechanism. Conduction and convection
heating are arguably slow and inefficient strategies for transferring
energy into the organic reactants because the role of heat penetration
through the reaction vessel and environment is undeniable. However,
microwave irradiation provides efficient heating by direct coupling of
microwave energy with the polar molecules (e.g., solvents, reagents
and catalysts) that are present in the reaction mixture.57
Over the last few years, it has been demonstrated that the synthesis of nanoporous metal oxides was significantly improved using
microwave irradiation as the source of heating because the synthesis
time was decreased by more than an order of magnitude, the product
could be formed uniformly in structure and composition, and the composition of the synthesized material could be controlled efficiently.58
Detailed studies on microwave effects were conducted by Perreux et
al.58 and Stuerga et al.59
Recently, a fast and cost-effective GaN:ZnO solid solution photocatalyst synthesis method was proposed by Yang et al.60 The stoichiometric amount of Ga2 O3 , ZnO and Zn powder and the excess
amount of urea, CH4 N2 O (as a nitrogen source), were uniformly mixed
and microwave-treated in a domestic microwave oven at a maximum
power of 800 W and operating frequency of 2.45 GHz for 10 min.
Copper (II) oxide (CuO) could also be used as the heated medium
material, which has been reported previously as being effective.61,62
Figure 9 demonstrates the synthesis of GaN:ZnO solid solution photocatalyst through the microwave heating technique.
The key difference of this synthesis method is the use of a different nitrogen source (urea) in the precursor mixture instead of am-

Figure 9. Schematic illustration of GaN:ZnO solid solution through microwave technique.

Figure 10. XRD pattern of GaN:ZnO solid solution obtained through the
microwave route (ZGNO-M) and one synthesized via the traditional method
(ZGNO-S).60

monia. Urea has been used as precursor for the synthesis of nitride
materials;6365 however, this is the first time that the synthesis of
GaN:ZnO solid solution has been reported through the facile method,
which did not involve the nitridation of starting materials with ammonia. The decomposition of urea and foaming at high temperatures
resulted in solid solution photocatalyst with a nanoporous surface
structure. Nanoporosity also might be attributed to the formation of
gaseous side products during the synthesis.
The bandgap of GaN:ZnO solid solution prepared through the microwave route was estimated, based on the light absorption spectra, to
be 2.47 eV, which was lower than the photocatalyst obtained through
the traditional method. Although there is no report in the literature on
the application of the GaN:ZnO photocatalyst prepared through the
microwave heating for overall water splitting, predictions can be made
based on band-gap energy and band-edge potential. The presence of
N2p and Zn3d electrons in the upper VB provides p-d repulsion for
the VB maximum and results in a narrower bandgap without shifting
the CB. Therefore, the photocatalyst synthesis through the facile microwave technique is expected to have the capability for overall water
splitting to produce hydrogen and oxygen under visible light. However, the short preparation time may not be long enough for solid-solid
interface diffusion. This issue may lead to the formation of different
species as side products, such as ZnCN2 and ZnGa2 O4 , due to the partial decomposition of urea and a photocatalyst with a low degree of
crystallinity. The low crystallinity of the resulting photocatalyst can be
easily distinguished from the powder XRD pattern provided by Yang
et al.,60 as seen in Figure 10. Moreover, the poor crystalline structure of the resulting solid solution might be attributed to the sudden
temperature increase and decrease in the bulk of the reaction mixture
during the synthesis process, which is the nature of the microwave
effect. Table II summarizes the key features of the GaN:ZnO solid
solution photocatalyst synthesized through the techniques reviewed
in this article.
As can be seen, the solid solution photocatalysts synthesized
through nontraditional techniques (medium-temperature solid-state,
layered double hydroxides precursor and microwave heating), have
higher surface areas and Zn/Ga ratios. However, the crystallinity of
the prepared photocatalysts is an issue that should be improved for
each technique through pre- and post-treatment techniques or both.

Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Q124

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)

Table II. Properties of the GaN:ZnO solid solution photocatalyst prepared at the high-temperature solid state (HTSS), medium-temperature solid
state (MTSS), layered double hydroxide precursor (LDHP) and microwave heating (MW) techniques.

HTSS
MTSS
LDHP
MW

Cost of synthesis

Crystallinity

BET surface area (m2 /g)

Zn/Ga (%)

Band-gap energy (eV)

Ref.

Very high
High
Medium
Low

High
Medium
High
Low

3.58.1
55.456.9

31.1

<28
<38
<85
<35

2.62.8
2.32.5
2.42.6
2.47

32, 37
44, 45
16
60

One-Dimension Nanostructure GaN:ZnO Solid Solution


Photocatalyst
Despite significant achievements in quantum yield over the
past few years for visible-light overall water splitting using the
GaN:ZnO solid solution photocatalyst, most of the photogenerated
electrons/holes recombine within the bulk of the photocatalyst. Semiconductor nanowires (NWs) and nanotubes (NTs) have been proposed to be beneficial for overall water splitting applications because
of their unique properties.6668 Many benefits are offered when using NWs and NTs for overall water splitting, such as short distances
for carrier collection/transport photocatalyst formation in the form
of hetero-junction interfaces, enhanced light trapping in high-density
one-dimensional (1-D) nanostructure arrays, and controlling the size
and composition over a wide range through flexible synthesis techniques. Photogenerated electrons/holes in bulk crystalline structures
and close-packed crystallites tend to recombine at structural crystal defects and grain boundaries, respectively. However, NWs array
grown on a conducting substrate avoids this disadvantage because
the charge carriers can be collected and transported to the conductive support along individual NWs, which will decrease the rate of

Figure 11. Schematic illustration of charge carrier transport in a single NW


grown on a conductive substrate and its electrical potential: (1) semiconductor
NW, (2) photoexcited electrons path and (3) conductive substrate.

electrons/holes recombination. As a result, photocurrent efficiency is


enhanced.69 This process is schematically illustrated in Figure 11.
Many researchers have focused on the synthesis and application of
semiconductor whiskers over the past few years. Synthesis semiconductors in the form of 1-D nanostructures instead of bulk materials
may change the key features of the material. For example, at room
temperature, bulk silicon is a good thermal conductor and electron
conductor, whereas silicon NWs are inherently thermal insulators and
good electron conductors, which is very useful for thermoelectric
generation, waste heat recovery, and power generation application at
a relevant temperature range.70
Recently, Hwang et al.71 reported the synthesis of threedimensional nanostructures of InGaN on Si NWs through the halide
chemical vapor deposition. Single-phase InGaN NWs have been vertically grown on the sidewalls of Si wires and used as high surface area photoanodes for water-splitting applications. Although GaN
has a large bandgap (3.4 eV), forming a solid solution of InGaN
tunes the bandgap from the ultraviolet to the near-infrared region,
covering the entire solar spectrum.71 The effectiveness of InGaN in
tuning the bandgap has been previously reported.72,73 The growth
of single-crystalline Inx Ga1x N NWs over the entire compositional
range (from x = 0 to 1) has been reported by Kuykendall et al.72 The
low-temperature chemical vapor deposition synthesized NWs were
reported to have tunable emission over the entire solar spectrum from
the ultraviolet to the near-infrared region.
Indium nitride (InN) has the same wurtzite structure as that of
GaN:ZnO.13 Since InN has a narrower bandgap (<1 eV)74 than
GaN:ZnO (2.42.8 eV), it is speculated that the formation of a
GaN:ZnO-InN solid solution caused the shift of the absorption edge of
the solid solution to the longer wavelengths. Kamata et al.75 reported
the synthesis of GaN:ZnO-InN solid solution through the traditional
nitridation of the mixed oxide of gallium, zinc, and indium.75 Although the prepared solid solution exhibited an absorption band near
600 nm, the unmodified solid solution reported was unstable for the
water oxidation half-reaction.75
Comprehensive review articles have been published on the
synthesis, characterization and application of 1-D nanostructure
materials.7678 The synthesis methods of ZnO7982 and GaN8388 NWs
and NTs are well understood. However, only a few research articles
have been published on the synthesis and application of GaN:ZnO
solid solution NWs or NTs. Han et al.89 reported the synthesis and
electronic properties of single-crystal GaN:ZnO solid solution NWs
with relatively low Zn content (x = 0.12), using nanostructured zinc
gallate as a precursor, prepared by a sol-gel method.89 Hahn et al. also
recently reported an epitaxial casting method to synthesize singlecrystalline, high surface area GaN:ZnO solid solution NTs with ZnO
compositions of up to x = 0.10. They further reported that the overall water-splitting efficiency improved 1.52 times by using the NTs,
comparing them to the similar composition powders for the ratelimiting H+ reduction half-reaction.18 Table III compares 1-D NWs
with the bulk photocatalyst prepared through the synthesis techniques
reviewed in this article.
Although the synthesis of uniform GaN:ZnO solid solution nanostructures and surface modification is not well understood and still
involves difficulties and challenges, the use of 1-D nanostructure
GaN:ZnO solid solution offers many opportunities to make solar hydrogen production feasible.

Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)

Q125

Table III. Comparison between 1-D nanostructured and bulk photocatalysts prepared through the high-temperature solid state (HTSS), mediumtemperature solid state (MTSS), layered double hydroxide precursor (LDHP), microwave heating (MW), One-Dimension Nanostructure (1-D)
and nanowires (NWs) techniques.
Advantages
HTSS

MTSS

Disadvantages

Comments

Highly crystalline structure

Inefficient synthesis technique

Applied for overall water splitting

Stable in overall water-splitting reaction

Low Zn content

Highest activity for Rh2y Cry O3


core/shell-loaded solid solution

High efficiency for solar hydrogen


generation

Low surface area

One-step technique

LDHP

Mild synthesis temperature

Average crystallinity

Applied for photodegradation of IPA

High surface area

Two-step technique

Crystallinity improvement by calcination


of precursor

Efficient synthesis

Two-step technique

Applied for photoreduction of Cr6+ ions

Stable crystalline structure

Co-precipitation highly sensitive to


supersaturation level

Pt applied as co-catalyst

Precise control on Zn content


Highly efficient synthesis

Formation of side products

Applied for photodegradation of IPA

Nanoporous surface structure

Low crystallinity

High Zn content
Strong gas absorption

MW

High Zn content

High Zn content
Strong gas absorption
1-D
NWs

Highly crystalline structure

Complex and difficult synthesis technique

High surface area


Enhanced charge carriers collect/transport
Light trapping

Conclusions
Clean and endless sources of energy would be the ultimate desire
of humanity in the upcoming decades. Overall water splitting using
visible-light-activated semiconductor photocatalysts can provide this
source of energy. Numerous combinations of semiconductor materials, electrocatalysts and cell configurations have been studied for photoelectrolysis research; however, the magic component that is capable
of producing hydrogen under visible-light irradiation has not been
found. GaN:ZnO solid solution photocatalyst, as a new visible-light
active photocatalyst, has attracted the attention of many researchers.
The Rh-Cr-oxide-loaded GaN:ZnO solid solution photocatalyst responds to approximately 500 nm for overall water splitting with the
highest quantum yield reported so far in the literature;90 still much
lower than the target in this field (30% at 600 nm).8,37 Considerable
improvement in photocatalytic activity of GaN:ZnO for overall water
splitting requires meeting the feasibility limit for solar hydrogen production on the TW-scale. Extending research to different precursors,
pre-treatment, synthesis techniques and post-treatment, as well as using different co-catalysts and photoreaction conditions, is required for
such an achievement.
This article is the first to review the different GaN:ZnO solid solution photocatalysts synthesis techniques and address the advantages
and disadvantages of each method. The key features of each synthesis method and their effects on final synthesized photocatalysts have
been elaborated. In addition, the effectiveness of the starting material
structures in the form of mesoporous particles has been discussed.
Using different precursors (e.g., Zn2+ /Ga3+ /Co3 and LDHs), nitrogen sources (e.g., urea) and heating strategies of the reaction mixture
(e.g., microwave irradiation) have been presented. Although the band-

gap energy and band-edge potential of the photocatalysts synthesized


through various techniques are suitable for visible-light overall water splitting, to our knowledge, no report has addressed overall water
splitting by applying the photocatalyst synthesized through any technique other than the traditional synthesis method. The development
of an efficient synthesis technique for preparing visible-light activated
photocatalysts such as GaN:ZnO solid solution could be a key factor
in the generation of future solar fuels.

References
1. N. S. Lewis and D. G. Nocera, Proc. Natl. Acad. Sci. U.S.A., 103, 15729 (2006).
2. Basic research needs for solar energy utilization, Report on the Basic Energy Sciences Workshop on Solar Energy Utilization., Available on the web at
http://www.sc.doe.gov/bes/reports/files/SEU_rpt.pdf.
3. D. G. Nocera, Chem. Soc. Rev., 38, 13 (2009).
4. G. W. Crabtree and N. S. Lewis, Phys. Today 60, 37 (2007).
5. M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. Mi, E. A. Santori,
and N. S. Lewis, Chem. Rev., 110, 6446 (2010).
6. R. M. Navarro, M. C. Alvarez-Galvan, J. A. Villoria de la Mano, S. M. Al-Zahrani,
and J. L. G. Fierro, Energy Environ. Sci., 3, 1865 (2010).
7. G. W. Crabtree, M. S. Dresselhaus, and M. V. Buchanan, Phys. Today, 57, 39 (2004).
8. A. Kudo and Y. Miseki, Chem. Soc. Rev., 38, 253 (2009).
9. M. Pelaez, N. T. Nolan, S. C. Pillai, M. K. Seery, P. Falaras, A. G. Kontos,
P. S. M. Dunlop, J. W. J. Hamilton, J. A. Byrne, K. OShea, M. H. Entezari, and
D. D. Dionysiou, Appl. Catal. B, 125, 331 (2012).
10. A. Fujishima and K. Honda, Nature, 238, 37 (1972).
11. K. Maeda, T. Takata, M. Hara, N. Saito, Y. Inoue, H. Kobayashi, and K. Domen, J.
Am. Chem. Soc., 127, 8286 (2005).
12. K. Maeda, K. Teramura, D. Lu, T. Takata, N. Saito, Y. Inoue, and K. Domen, Nature,
440, 295 (2006).
13. K. Maeda, K. Teramura, T. Takata, M. Hara, N. Saito, K. Toda, Y. Inoue, H. Kobayashi,
and K. Domen, J. Phys. Chem. B, 109, 20504 (2005).

Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Q126

ECS Journal of Solid State Science and Technology, 2 (7) Q118-Q126 (2013)

14. K. Maeda, K. Teramura, N. Saito, Y. Inoue, H. Kobayashi, and K. Domen, Pure Appl.
Chem., 78, 2267 (2006).
15. H. Chen, W. Wen, Q. Wang, J. C. Hanson, J. T. Muckerman, E. Fujita, A. I. Frenkel,
and A. Rodriguez, J. Phys. Chem. C, 113, 3650 (2009).
16. J. Wang, B. Huang, Z. Wang, P. Wang, H. Cheng, Z. Zheng, X. Qin, X. Zhang, Y. Dai,
and M. H. Whangbo, J. Mater. Chem., 21, 4562 (2011).
17. A. Kudo, Int. J. Hydrogen Energy, 32, 2673 (2007).
18. C. Hahn, M. A. Fardy, C. Nguyen, M. Natera-Comte, S. C. Andrews, and P. Yang,
Isr. J. Chem., 52, 1111 (2012).
19. K. Maeda, K. Teramura, and K. Domen, J. Catal., 254, 198 (2008).
20. A. Kudo, H. Kato, and I. Tsuji, Chemistry Letters, 33, 1534 (2004).
21. H. Kato and A. Kudo, Catalysis Today, 78, 561 (2003).
22. Q. Z. Liu and S. S. Lau, Solid-State Electron., 42, 677 (1998).
23. F. A. Ponce and D. P. Bour, Nature, 386, 351 (1997).
24. S. S. Kocha, M. W. Peterson, D. J. Arent, J. M. Redwing, M. A. Tischler, and
J. A. Turner, J. Electrochem. Soc., 142, L238 (1995).
25. I. M. Huygens, K. Strubbe, and W. P. Gomes, J. Electrochem. Soc., 147, 1797 (2000).
26. D. K. Hwang, M. S. Oh, J. H. Lim, and S. J. Park, J. Phys. D: Appl. Phys., 40, R387
(2007).
27. M. J. Chen, J. R. Yang, and M. Shiojiri, Semicond. Sci. Technol., 27, 074005 (2012).
28. Y.-S. Choi, J.-W. Kang, D.-K. Hwang, and S.-J. Park, IEEE Trans. on Electron
Devices, 57, 26 (2010).
29. J. Bao, M. a. Zimmler, F. Capasso, X. Wang, and Z. F. Ren, Nano Lett., 6, 1719
(2006).
30. X. Duan and C. M. Lieber J. Am. Chem. Soc., 122, 188 (2000).
31. S. Maensiri, P. Laokul, and V. Promarak, J. Cryst. Growth, 289, 102 (2006).
32. K. Maeda and K. Domen, Chem. Mater., 22, 612 (2010).
33. C. Randorn, J. T. S. Irvine, and P. Robertson, Int. J. Photoenergy, 2008, 1 (2008).
34. Y. Lee, H. Terashima, Y. Shimodaira, K. Teramura, M. Hara, H. Kobayashi,
K. Domen, and M. Yashima, J. Phys. Chem. C, 111, 1042 (2007).
35. S. H. Wei and A. Zunger, Phys. Rev. B, 37, 8958 (1988).
36. H. Shi and Y. Duan, Eur. Phys. J. B, 66, 439 (2008).
37. K. Maeda and K. Domen, J. Phys. Chem. C, 111, 7851 (2007).
38. W. Zhao, W. Ma, C. Chen, J. Zhao, and Z. Shuai, J. AM. CHEM. SOC., 126, 4782
(2004).
39. K. Maeda, K. Teramura, D. Lu, N. Saito, Y. Inoue, and K. Domen, J. Phys. Chem. C,
111, 7554 (2007).
40. T. Hisatomi, K. Maeda, K. Takanabe, J. Kubota, and K. Domen, J. Phys. Chem. C,
113, 21458 (2009).
41. T. Ohno, L. Bai, T. Hisatomi, K. Maeda, and K. Domen, J. Am. Chem. Soc., 134,
8254 (2012).
42. T. Hisatomi, K. Maeda, D. Lu, and K. Domen, ChemSusChem, 2, 336 (2009).
43. W. Wei, Y. Dai, K. Yang, M. Guo, and B. Huang, J. Phys. Chem. C, 112, 15915
(2008).
44. S. Yan, Z. Wang, Z. Li, and Z. Zou, J. Mater. Chem., 5682 (2011).
45. S. C. Yan, S. X. Ouyang, J. Gao, M. Yang, J. Y. Feng, X. X. Fan, L. J. Wan, Z. S. Li,
J. H. Ye, Y. Zhou, and Z. G. Zou, Angew Chem. Int. ed., 49, 6400 (2010).
46. A. G. Agrios and P. Pichat, J. Photochem. Photobiol. A, 180, 130 (2006).
47. G. Tian, H. Fu, L. Jing, and C. Tian, J. Hazard. Mater., 161, 1122 (2009).
48. K. B. Jaimy, S. Ghosh, and K. Gopakumar, J. Solid State Chem., 196, 465 (2012).
49. B. G. Yacob, Semiconductor Materials : An Introduction to Basic Principles, Kluwer
Academic/Plenum, 2003.
50. A. I. Khan and D. OHare, J. Mater. Chem., 12, 3191 (2002).
51. D. G. Evans and C. T. Slade, Struct. Bond., 119, 1 (2006).
52. F. Li and X. Duan, Struct. Bond., 119, 193 (2006).
53. F. Leroux and J. P. Besse, Chem. Mater., 13, 3507 (2001).
54. V. Prevot, C. Forano, and J. P. Besse, Appl. Clay Sci., 18, 3 (2001).

55. M. Lakraimi, A. Legrouri, A. Barroug, A. De Roy, and J. Pierre Besse, J. Mater.


Chem., 10, 1007 (2000).
56. S. P. Newman and W. Jones, New J. Chem., 22, 105 (1998).
57. C. O. Kappe and D. Dallinger, Nat. rev. Drug discovery, 5, 51 (2006).
58. L. Perreux and A. Loupy, Tetrahedron, 57, 9199 (2001).
59. D. A. C. Stuerga and P. Gillard, J. Microwave Power and Electromagn. Energy, 31,
87 (1996).
60. M. Yang, Q. Huang, and X. Jin, Solid State Sci., 14, 465 (2012).
61. D. Kim, K. H. Ko, and K. S. Hong, J. Am. Ceram. Soc., 84, 1286 (2001).
62. Y. Zhao, J.-J. Zhu, J.-M. Hong, N. Bian, and H.-Y. Chen, Eur. J. Inorg. Chem., 2004,
4072 (2004).
63. S. C. Lee, H. O. Lintang, and L. Yuliati, Asian J. of Chem., 7, 2139 (2012).
64. K. Uheda, M. Takahashi, H. Takizawa, T. Endo, and M. Shimada, Key Eng. Mater.,
159160, 53 (1999).
65. N. G. Moustakas, A. G. Kontos, V. Likodimos, F. Katsaros, N. Boukos, D. Tsoutsou,
A. Dimoulas, G. E. Romanos, D. D. Dionysiou, and P. Falaras, Appl. Catal. B,
130131, 14 (2013).
66. H. M. Chen, C. K. Chen, Y. Chang, C. Tsai, R. Liu, S. Hu, W. Chang, and K. Chen,
Angew. Chem., 122, 6102 (2010).
67. Z. Zhang, M. F. Hossain, and T. Takahashi, Int. J. Hydrogen Energy, 35, 8528
(2010).
68. G. Wang, H. Wang, Y. Ling, Y. Tang, X. Yang, R. C. Fitzmorris, C. Wang, J. Z. Zhang,
and Y. Li, Nano lett., 11, 3026 (2011).
69. Xuebo Cao, Peng Chen, and Yang Guo, J. Phys. Chem. C, 112, 20560 (2008).
70. P. Yang, R. Yan, and M. Fardy, Nano lett., 10, 1529 (2010).
71. Y. J. Hwang, C. H. Wu, C. Hahn, H. E. Jeong, and P. Yang, Nano lett., 12, 1678
(2012).
72. T. Kuykendall, P. Ulrich, S. Aloni, and P. Yang, Nat. mater., 6, 951 (2007).
73. C. Pendyala, J. B. Jasinski, J. H. Kim, V. K. Vendra, S. Lisenkov, M. Menon, and
M. K. Sunkara, Nanoscale, 4, 6269 (2012).
74. F. Bechstedt and J. Furthm, J. Cryst. Growth, 246, 315 (2002).
75. K. Kamata, K. Maeda, D. Lu, Y. Kako, and K. Domen, Chem. Phys. Lett., 470, 90
(2009).
76. B. Y. Xia, P. Yang, Y. Sun, Y. Wu, B. Mayers, B. Gates, Y. Yin, F. Kim, and H. Yan,
Adv. Mater., 15, 353 (2003).
77. J. Hu, T. W. Odom, and C. M. Lieber, Acc. Chem. Res., 32, 435 (1999).
78. A. I. Hochbaum and P. Yang, Chem. Rev., 110, 527 (2010).
79. Y. Zhang, M. K. Ram, E. K. Stefanakos, and D. Y. Goswami, J. Nanomaterials, 2012,
1 (2012).
80. Z. L. Wang, Mater. Sci. Eng. R, 64, 33 (2009).
81. T. Ngo-Duc, K. Singh, M. Meyyappan, and M. M. Oye, Nanotechnol., 23, 194015
(2012).
82. Y. W. Heo, D. P. Norton, L. C. Tien, Y. Kwon, B. S. Kang, F. Ren, S. J. Pearton, and
J. R. LaRoche, Mater. Sci. Eng. R, 47, 1 (2004).
83. H. Wu, H. Cha, M. Chandrashekhar, M. G. Spencer, and G. Koley, J. Electron. Mater.,
35, 670 (2006).
84. F. Qian, Y. Li, S. Gradec, and C. M. Lieber, Nano Lett., 4, 1975 (2004).
85. S. C. Lyu, O. H. Cha, E. Suh, H. Ruh, H. J. Lee, and C. J. Lee, Chem. Phys. Lett.,
367, 136 (2003).
86. J. Y. Li, X. L. Chen, Z. Y. Qiao, Y. G. Cao, M. He, and T. Xu, Appl. Phys. A, 71, 349
(2000).
87. J. C. Wang, C. Z. Zhan, and F. G. Li, Appl. Phys. A, 76, 609 (2003).
88. J. Goldberger, R. He, Y. Zhang, S. Lee, H. Yan, H. J. Choi, and P. Yang, Nature, 422,
599 (2003).
89. W. Q. Han, Y. Zhang, C. Y. Nam, C. T. Black, and E. E. Mendez, Appl. Phys. Lett.,
97, 083108 (2010).
90. N. Saito, Y. Inoue, and K. Domen, Angew. Chem. Int. Ed., 45, 7806 (2006).

Downloaded on 2016-02-29 to IP 132.248.156.253 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

You might also like