You are on page 1of 12

DOI: 10.1002/cssc.

201500383

Full Papers

Time- and Energy-Efficient Solution Combustion Synthesis


of Binary Metal Tungstate Nanoparticles with Enhanced
Photocatalytic Activity
Abegayl Thomas,[a] Csaba Janky,*[b, c] Gergely F. Samu,[b, c] Muhammad N. Huda,[d]
Pranab Sarker,[d] J. Ping Liu,[d] Vuong van Nguyen,[d] Evelyn H. Wang,[e] Kevin A. Schug,[e] and
Krishnan Rajeshwar*[a, f]
In the search for stable and efficient photocatalysts beyond
TiO2, the tungsten-based oxide semiconductors silver tungstate
(Ag2WO4), copper tungstate (CuWO4), and zinc tungstate
(ZnWO4) were prepared using solution combustion synthesis
(SCS). The tungsten precursors influence on the product was
of particular relevance to this study, and the most significant
effects are highlighted. Each samples photocatalytic activity
towards methyl orange degradation was studied and benchmarked against their respective commercial oxide sample ob-

tained by solid-state ceramic synthesis. Based on the results


herein, we conclude that SCS is a time- and energy-efficient
method to synthesize crystalline binary tungstate nanomaterials even without additional excessive heat treatment. As many
of these photocatalysts possess excellent photocatalytic activity, the discussed synthetic strategy may open sustainable materials chemistry avenues to solar energy conversion and environmental remediation.

Introduction
The global need to focus on renewable and earth-abundant resources has resparked much attention on inorganic oxide semiconductor materials for solar energy conversion[1] and photo[a] A. Thomas, Prof. Dr. K. Rajeshwar
Department of Chemistry & Biochemistry
University of Texas at Arlington
Arlington, TX 76019 (USA)
E-mail: rajeshwar@uta.edu
[b] Prof. Dr. C. Janky, G. F. Samu
Department of Physical Chemistry and Materials Science
University of Szeged
Szeged, Rerrich Sq. 1, 6720 (Hungary)
E-mail: janaky@chem.u-szeged.hu
[c] Prof. Dr. C. Janky, G. F. Samu
MTA-SZTE Lendlet Photoelectrochemistry Research Group
Szeged, Rerrich Sq. 1, 6720 (Hungary)
[d] Prof. Dr. M. N. Huda, P. Sarker, Prof. Dr. J. P. Liu, Dr. V. van Nguyen
Department of Physics
University of Texas at Arlington
Arlington, TX 76019 (USA)
[e] E. H. Wang, Prof. K. A. Schug
Department of Chemistry & Biochemistry
University of Texas at Arlington
Arlington, TX 76019 (USA)
[f] Prof. Dr. K. Rajeshwar
Center for Renewable Energy Science & Technology
University of Texas at Arlington
Arlington, TX 76019 (USA)
Supporting Information for this article is available on the WWW under
http://dx.doi.org/10.1002/cssc.201500383.
This publication is part of a Special Issue on Green Chemistry and the
Environment. To view the complete issue, visit:
http://onlinelibrary.wiley.com/doi/10.1002/cssc.v8.10/issuetoc

ChemSusChem 2015, 8, 1652 1663

catalytic environmental remediation.[2, 3] Titanium dioxide (TiO2)


has been the most extensively studied oxide due to its excellent photoelectrochemical and photocatalytic behavior.[4] However, TiO2 has a relatively wide band gap (3.03.2 eV), thus restricting light absorption to the UV region of the solar spectrum that comprises only ~ 4 % of the solar spectrum. On the
other hand, this material exhibits little or no absorption in the
visible region (comprising ~ 50 % of the solar spectrum). Aside
from TiO2, tungsten trioxide (WO3) has been widely studied,
serving as an attractive material for both water oxidation and
environmental remediation.[5] Tungsten trioxide shares characteristics similar to TiO2 in terms of chemical inertness and
photo- and chemical stability in aqueous solutions in a relatively broad pH range. Importantly, WO3 has a band gap energy
(Eg) of 2.7 eV (vs. 3.2 eV for anatase TiO2), with its absorption
edge located just at the cusp of the visible light spectrum. The
high valence band edge energy (VB) of + 3.1 eV makes WO3
a suitable photocatalyst for water photo-oxidation; however,
its conduction band (CB) is more positive than that necessary
for spontaneous solar hydrogen evolution.[6]
Driven by the above considerations and a recognition of the
critical need to discover families of new materials, the search
for oxide semiconductors besides TiO2 and WO3 has gained
considerable momentum in recent years. Appreciable efforts
have been devoted to the shifting of the VB and CB edge positions of different compound semiconductors (bandgap engineering) to tailor their interfacial energetics to targeted photooxidation or photoreduction processes, respectively.[1, 6] This
can be done by doping/alloying or synthesizing composites, in
many cases using TiO2 or WO3 as one component (Scheme 1).

1652

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers

Scheme 1. Approximate CB and VB edge positions of the synthesized binary


tungstates. The interfacial energetics for TiO2 and WO3 are also shown for
comparison.

ic.[21] See Figure 1 A and B for model structure representations


of each of the discussed binary tungstates.
On the other hand, introducing a monovalent cation such as
Ag favors a complicated network structure of silver tungstate
(Ag2WO4), which can exhibit three different structural phases;
a-, b-, and g-Ag2WO4.[22] Among those polymorphs a-Ag2WO4
is thermodynamically most stable[22] and belongs to orthorhombic symmetry Pn2n.[22, 23] All tungsten atoms are six-coordinate and form WO6 octahedra. These WO6, W2O6, and W3O6
octahedra are connected by sharing edges and grouped together at a particular position (see Figure 1 C). On the other
hand, the number of different sites occupied by the silver
atoms in Ag2WO4 is six, which form different polyhedra.[23]
Literature survey indicates that synthesis of ZnWO4, CuWO4,
and Ag2WO4 has been dominated by conventional methods
such as hydrothermal synthesis or solid-state reaction.[13, 18, 2430]
These methods suffer from similar drawbacks, most importantly the long reaction time (usually several hours) and the need

A shift of the band edge positions can alter the energy band
gap of these materials, which in
turn may enhance photocatalytic
activity. Such a shift may also
shrink the band gap, thus increasing solar light absorption
without changing the reduction/
oxidation potential of the photogenerated charge carriers. There Figure 1. Ball-and-stick model of the crystal structure for ZnWO (A), CuWO (B), and Ag WO (C).
4
4
2
4
are many examples where the
photocatalytic or photoelectrochemical performance of binary
for high energy input to maintain the elevated temperature
(or even ternary) tungstates outperformed WO3,[6] notably,
during the synthesis. We note here that minimization of both
Bi2WO6,[79] SnWO4,[10, 11] ZnWO4,[1214] AgBiW2O8,[15] and
these factors is crucially important from a sustainability perAgInW2O8.[16, 17]
spective, especially if we plan to utilize the materials in appliWe present below the strategy of incorporating Zn, Cu, and
cations such as solar fuel generation or environmental remeAg cations into WO3 to form ZnWO4, CuWO4, and Ag2WO4, rediation.
spectively, thereby opening up an avenue for tuning interfacial
The process known as solution combustion synthesis (SCS)
energetics. Introduction of a heteroatom in the WO3 structure
addresses the above challenges admirably well; however,
can result in the formation of one of two major crystal strucsomewhat surprisingly, its use has been quite limited in the
tures: wolframite (smaller divalent cations), scheelite (larger diphotocatalysis community. In this method, reaction times are
valent cations), or other possible crystal structures (monovalent
short, no special equipment is needed, and the process is cost
cations). Accordingly, small divalent cations (ionic radius
effective.[3133] The above-listed benefits are rooted in the exo< 0.77 ) such as Zn and Cu tend to form wolframite structures
[18]
thermicity of the reaction, along with the expulsion of noxious
Zinc tungstate (ZnWO4) crystallizes to
(WO6 octahedra).
gases, which together result in small-sized solid particles of the
a monoclinic wolframite structure (P2/c), with both Zn and W
synthesized products. SCS has been deployed to prepare oxide
forming ZnO6 and WO6 octahedra connected by edge sharing.
semiconductors, such as TiO2,[34, 35] ZnO,[36] Fe2O3,[37] and WO3.[5]
Copper tungstate (CuWO4) crystallizes in a triclinic structure

with symmetry P1. In the structure of CuWO4, both Cu and W


Furthermore, the role of fuel was carefully studied, even as
a synthetic strategy tool to obtain doped oxide as a result of
are surrounded by six oxygen atoms to form CuO6 and WO6
the SCS process.[33, 37] On the other hand, much less is known
octahedra.[19] These two different octahedra connected by
about the role of precursors and their effect on the resultant
edge sharing oxygen atoms, form infinite zigzag chains. Almaterials. Finally, to the best of our knowledge, there is no
though CuWO4 is triclinic, its structure is topologically related
precedence in the literature on the SCS of binary tungstates.
to that of monoclinic wolframite (P2/c).[20] The CuO6 octahedra
In this study, SCS was employed to prepare three members
demonstrate JahnTeller distortion to remove the degeneracy
of the binary tungstate family, namely ZnWO4, CuWO4, and
of Cu2 + 3d orbitals. This distortion elongates the octahedron,
Ag2WO4. The precursor influence on the morphological and
causing reduction in the symmetry from monoclinic to triclinstructural attributes of the resultant materials was examined
ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

1653

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers
systematically by using two different tungsten precursors in
the synthesis process. Moreover, the photocatalytic activity of
each sample was evaluated for its ability to degrade an organic
pollutant, namely, methyl orange dye. Apparent rate constants
were compared and contrasted with those obtained for their
commercial counterparts prepared by solid-state reaction. The
major innovation of this study was the fact that the SCS-synthesized new families of binary tungstates, obtained within
five minutes with very little energy input, outperformed their
commercial counterparts (synthesized using time- and energyinefficient methods) in photocatalytic tests.

Results and Discussion


Thermogravimetric analysis
The use of thermal analysis can provide a better understanding
of the SCS process. Hence, for the simulation of the SCS procedure, thermogravimetric analysis/differential scanning calorimetry (TGA/DSC) measurements were performed on each of the
precursor mixtures. As an example, TGA/DSC profiles for the
formation of CuWO4 from two different precursor mixtures are
given in Figure 2 A and B. The tungstate samples prepared
using both precursors showed multiple mass loss regimes. The
initial mass loss between 100150 8C corresponds to the loss
of water from the precursor mixture in both cases. Using ammonium tungstate as precursor, all subsequent mass loss
occurs in one single step above 300 8C, which originates from
the combinations of fuel and metal-salt-precursor decomposition (and combustion). Note, however, that the above-mentioned temperature is only the temperature where combustion
initiates; the actual temperatures after the combustion can be
much higher than those indicated on the x-axis (given by the
programmed-temperature ramp, see the Experimental Section).
On the other hand, samples prepared using the sodium tungstate precursor had an expanded temperature range of 120
210 8C, which was assigned to the decomposition of the fuel
and the metal salt precursor. At temperatures over 480 8C an-

other mass loss regime was observed, which corresponds to


the loss of residual carbon in the matrix.
Quantitative assessment of the TGA curves was also performed. As the first step, the initial mass loss (related to water
evaporation) was subtracted. Subsequently, the relative mass
losses were determined for both mixtures and these values
were compared with the theoretically calculated values. The
good agreement between the actual and stoichiometric mass
losses for both the sodium (34 % vs. 31 %) and the ammonium
precursor (39 % vs. 41 %) confirmed that the proposed combustion reaction occurred in both cases, albeit with different
intensity.
DSC analysis yielded both endothermic and exothermic
peaks. A fairly large and broad endothermic peak was witnessed at the early stage of analysis, which is assignable to the
initial loss of moisture within the precursor material and correlates with the first mass loss on the corresponding TGA profile.
More intense and sharp exothermic peaks were observed for
the precursor mixtures containing (NH4)2WO4, compared to
those with Na2WO4 (Figure 2 B), perhaps not surprisingly in
light of its combustible nitrogen content. This exothermic peak
suggests that a high combustion temperature is reached
during the reaction, which also coincides with the oxidative
decomposition of the fuel.
Crystal structure and morphology
Powder X-ray diffraction (XRD) was employed to characterize
the crystalline phases of the synthesized and respective commercial oxide samples. The as-synthesized materials were annealed at various temperatures up to 500 8C (except for
CuWO4, where the maximum temperature was 400 8C because
of the limited thermal stability of this particular oxide) to enhance the degree of crystallinity and to remove any carbon
residue and impurities that may be present. Figure 3 AC illustrate the XRD patterns mapping the changes in the sample
crystallinity from as-synthesized to the sample annealed at the
highest temperature. The most vital message of these data is
that most SCS samples [except ZnWO4 (Na) and Ag2WO4 (NH4)]

Figure 2. TGA (A) and DSC (B) profiles of the SCS precursor mixtures for CuWO4.

ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

1654

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers
precursors show distinctly different purity and degree of crystallinity. Therefore, careful Rietveld refinement analysis was performed on the experimental data obtained at the highest annealing temperature [see the fitted XRD patterns (Figure S1)
and more details in the Supporting Information].
Briefly, Rietveld analysis of the samples synthesized with the
sodium tungstate precursor yielded single-phase structures in
agreement with their respective JCPDS file. The materials prepared using the ammonium tungstate precursor, however,
showed the presence of an additional minority WOx phase (Figure S1 AF). Additionally, what is indeed surprising, the commercial samples (except silver tungstate) were not of a single
phase and exhibited low crystallinity. The presence of additional peaks in the XRD pattern for the ammonium-derived samples may be a result of the use of a low pH value caused by
the addition of HNO3. Recent literature supports this notion; in
the synthesis of ZnWO4 using the hydrothermal process, WO3
formation was noted in the XRD pattern when the solution
mixture was at a low pH range (pH < 4).[13, 29] Additionally, the
more exothermic environment (and consequently a faster reaction) in the case of the ammonium precursor can be another
rationale for the appearance of the minority WO3 phase.
The average crystallite sizes were calculated for the synthesized samples using the most intense peaks by using the
Scherrer equation (Table 1; more details in the Supporting in-

Table 1. Calculated average crystallite sizes of the various tungstate samples as a function of annealing and tungstate precursor.
SCS sample

Crystallite size [nm]


as-synthesized

annealed

ZnWO4 (Na)
CuWO4 (Na)
Ag2WO4 (Na)

17
16 1
22 1
19 1
10 1
21 1
11 2
26
82

32 1
22 1
33 2
23 1
16 2
20 1
21 10
14 11
15 3

ZnWO4 (NH4)
CuWO4 (NH4)
Ag2WO4 (NH4)

Figure 3. XRD patterns for A) ZnWO4, B) CuWO4, and C) Ag2WO4 samples.


* represents the presence of WO3 in the synthesized sample.

are at least partially crystalline in nature, even without any additional heat treatment. However, the various oxides and more
interestingly even the same oxide synthesized using different
ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

ZnWO4
WO3x
CuWO4
WO3x
Ag2WO4
WO3x

formation). The first clear trend is that the synthesized particles


are nanocrystalline in nature, and the results show an increase
in crystallite size with increase in annealing temperature, as expected. Interestingly, the calculated crystallite size for the assynthesized ZnWO4 (Na) sample was as low as 1 nm. After annealing, all tungstate samples synthesized using the ammonium precursor showed a smaller particle size compared to their
counterparts using the sodium-containing precursor. The smaller particle size appears to be associated with less single-phase
material as a result of the higher reaction temperature and the
larger amount of gas released during the synthesis. Note, however, that the Scherrer equation can only be applied to wellcrystallized samples; otherwise XRD peak broadening can reflect low crystallinity instead of nanosized entities. Therefore,
the obtained crystallite sizes are considered as only estimates
in these cases.

1655

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers
TEM images of the tungstate
samples were additionally obtained to compare any contrasting morphologies that may arise
from the use of different precursors (Figure 4). In general, the
particle sizes obtained from TEM
were larger than the calculated
crystallite sizes from XRD data
(or in other words the size of the
crystalline domains). As for
CuWO4 (similarly for both precursors), the as-synthesized samples
yielded particle sizes ranging
from 1550 nm with the size increasing with increasing annealing
temperature.
Figure 4 A
shows samples with rounded
edges signaling that they are at
least partly in an amorphous
phase. Closer inspection, however, revealed crystalline domains,
where the lattice fringes were
visible (corresponding to the
[011] lattice plane),[38] in agreement with the previously presented XRD data (Figure 4 I).
In contrast, the as-synthesized
ZnWO4 (Na) was completely
amorphous with agglomerated
small-sized nanoparticles. For
the annealed ZnWO4 samples,
however, the observable sharp
edges of much larger particles
were seen, indicating improved
crystallinity consistent with the
XRD patterns. The most interesting morphological feature was
noted for the ZnWO4 (NH4) samples (Figure 4 E and F), where
the sample showed partial crystallinity and well-defined hexagonal nanorod morphology even
without heat treatment, again
corroborating the previously
shown XRD patterns. Note, that
this type of morphology was observed for ZnWO4 before but as
a result of a lengthy hydrothermal synthetic procedure.[13, 14, 26]
High-resolution (HR)TEM images
confirmed that ZnWO4 nanorods
preferentially grow along the
[100] direction. Interplanar spacings were determined to be
0.575 and 0.468 nm, corresponding to the [010] and [100] lattice
ChemSusChem 2015, 8, 1652 1663

Figure 4. TEM images of A, B) CuWO4 (Na) as-synthesized and annealed at 400 8C, respectively; C, D) ZnWO4 (Na)
as-synthesized and annealed at 500 8C, respectively; E, F) ZnWO4 (NH4) as-synthesized and annealed at 500 8C, respectively; G, H) Ag2WO4 (NH4) as-synthesized and annealed at 500 8C. I, J) HRTEM images of the nanoparticles in
panels A and E.

www.chemsuschem.org

1656

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers
planes, respectively, which is entirely consistent with earlier reports.[13, 39] Besides the rod-shaped crystals, some sphericalshaped nanoparticles can be also seen in the images for the
as-synthesized samples, whereas strictly rod-shaped crystals
(lengths 100150 nm) are present after annealing.
Ag2WO4 (Na) appeared to form much smaller particles compared to the other cases with sizes in the 410 nm range, with
many of the particles covered by a thin carbon film (as residue
from the SCS precursor). Notably, such carbon coating can be
indeed useful for certain applications, such as in Li-ion batteries, an alternative application area of these binary tungstates.[39, 40] Upon subsequent heat treatment, this carbon
matrix is removed and large aggregates of the small nanoparticles can be observed. Overall, we can conclude that a variety
of morphologies can be achieved upon using different precursors and subsequent heat treatment, tailoring the morphology
of these oxides towards a targeted application.
Further structural analysis of the synthesized tungstates was
performed using Raman spectroscopy. Using group theory, different studies have reported the existence of as many as 18
Raman-active vibrational modes for ZnWO4 and CuWO4.[4143]
Experimental results show that 1011 of the Raman active vibrations were observed for both oxides within the frequency
range of the experiment (see Table S1). Of the ten Ramanactive bands observed, 56 internal stretching modes were
present due to the WO bonds in the WO6 octahedra of the
metal tungstate structure. Importantly, assignment of the
Raman bands (see Figure 5 and Figures S2 and 3, and Table S1)
confirmed the formation of the respective compound structures for all samples. In addition, trends very similar to the
sample crystallinity are seen here and similar to those concluded from XRD (see the Supporting Information). The positions
of the most intense band in each sample corresponding to the
WO stretching mode were 907, 899, and 883 cm1 for ZnWO4,
CuWO4, and Ag2WO4, respectively. The 805 cm1 vibrational
mode characteristic of WO3 appeared in all samples prepared
using ammonium tungstate, thus confirming the multiphase
structure.[44]
Optical properties

Figure 5. Raman spectra of ZnWO4 (Na), CuWO4 (Na), and Ag2WO4 (Na) samples.

Figure 6. Tauc plots for the different ZnWO4 samples.

optical measurements yielded direct bandgap values in the


range 3.243.89 eV (depending on the synthesis and heat
treatment), all were within the range reported in previous
studies (see Table 2). CuWO4 and Ag2WO4 samples showed indirect bandgap values in the 1.802.04 and 2.443.10 eV
ranges, again in reasonable accord with the literature-reported
values. It is worth noting the very sizeable spread in literature

UV/Vis diffuse reflectance spectroscopy was employed to estimate the band gap values of the synthesized oxides. This was
achieved by generating Tauc plots, namely, a plot of the KubelkaMunk function versus photon
energy (ahn0.5 vs. hn).[45] Figure 6
contains an example for ZnWO4.
Table 2. Optical properties of the synthesized samples prepared using different tungstate precursors.
Table 2 displays the experimental
Sample
Band gap [eV]
Calculated band gap Reported values
band gap values for the syntheas-synthesized annealed at highest temperature values [eV]
[eV]
sized powders along with the
direct
band
gap
calculated optical properties (see
3.89 0.02
3.24 0.03
ZnWO4 (Na)
below). The resulting tungstate
2.92
3.025.85[25, 26, 46]
ZnWO4 (NH4)
3.76 0.02
3.45 0.03
products were observed to be
indirect band gap
white, yellow-brown, and gray in
2.04 0.01
2.03 0.01
CuWO4 (Na)
2.10
1.782.79[21, 4648]
CuWO4 (NH4)
1.91 0.01
1.80 0.01
color for ZnWO4, CuWO4, and
3.10 0.01

Ag2WO4 (Na)
Ag2WO4, respectively (see Fig1.22
3.063.55[23, 27, 28]
Ag2WO4 (NH4)
2.44 0.01
2.83 0.02
ure S4). For ZnWO4, experimental
ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

1657

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers

Figure 7. Electronic band structure for Ag2WO4 (A), CuWO4 (B), and ZnWO4 (C).

values, presumably reflecting the extreme sensitivity of the optical characteristics to synthesis and post-synthesis treatment
details.

Electronic band structure calculations


Computational studies were performed to rationalize the
above optical properties, and to correlate them with the photocatalytic activity of these materials (discussed below). The
electronic band structure for each of the tungstate materials
were calculated along the special symmetry points in the Brillouin zone. For Ag2WO4, an indirect band gap value of 1.17 eV
was found between X and G (Figure 7 A). Along the YSX
region, the mid-gap bands are degenerate. Moreover, the less
dispersive nature of the bands in both VB and CB along this
region suggests higher effective masses and, hence, lower conductivity for both electrons and holes. The computed band
gap corresponds to optical transition due to electron transfer
from O 2p at the top of the VB to W 5d/Ag 4d at the bottom of
the mid-gap CB or from Ag 4d at the top of the VB to O 2p at
the bottom of the mid CB. Both these values are smaller than
the experimentally measured band gaps for as-synthesized
Ag2WO4 (Na) and Ag2WO4 (NH4) samples.
In Figure 7 B, an indirect band gap with a minimal value of
0.707 eV occurs along the Y!G region for CuWO4. The VB
along the G!X regions is very flat, suggesting higher effective
mass of holes. The CB is divided into two regions in which the
first region contains only two bands that are mid-gap levels
0.707 eV higher than the VB and dispersive throughout all
symmetry points. The bottom of the CB of the second region
is about 3.087 eV higher from the VB maximum. They are flat
along G!X and R!T but dispersive along the other symmetry
points. The minimal optical band gap was calculated to be
2.1 eV. By comparing optical absorption and band structure, it
can be concluded that the electron transition does not occur
between the top or just below the top of the VB and bottom
or just above the bottom of the CB, which are dominated by
Cu 3d and W 5d, respectively, since the dd transition is forbidden. Hence, the origin of the optical gap is attributed to elecChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

tron transfer from occupied O 2p states within the VB to the


unoccupied Cu 3d mid bands (first part of CB).
The DFT + U electronic band gap structure for ZnWO4 represents a direct band gap of 2.94 eV (very close to the calculated
optical absorption) occurring at the Z point. In Figure 7 C, the
VB is dispersive throughout all regions, indicating lower effective masses of holes. Similar dispersive features are also found
for the CB along the regions except for D to B, which suggests
higher effective masses of electrons, while other regions indicate lower effective masses of electrons. In addition, the presence of these lower effective masses of electron regions suggests an ease of transfer of electrons from the VB to the CB.
The calculated optical band gap (2.92 eV) corresponds to the
electronic band gap. Since the dd electron transition is forbidden and the top of VB is largely contributed by O 2p, the first
peak at 3.43 eV corresponds to electron transfer from O 2p at
the top of VB to W 5d at the bottom of CB at the Z point.
Photocatalytic activity
Figure 8 AC compares the photocatalytic activity profiles for
methyl orange (MO) dye degradation for the blank case (no
photocatalyst present), the commercial sample (obtained by
high temperature ceramic synthesis), and the as-prepared
nanoparticles (for both precursors). Clearly, the freshly synthesized nanoparticles showed good photocatalytic activity compared to the commercial samples under UV/Vis light irradiation. In fact, ZnWO4 (Na) exhibited the best photocatalytic activity, most likely owing to its pure, monoclinic single-phase
structure and relatively small particle size as confirmed by Rietveld analysis of the XRD data and TEM images. After 20 min irradiation, both SCS ZnWO4 samples had degraded ~ 90 % of
the MO dye whereas the commercial sample required 50 min.
Additionally, the results show that both samples of CuWO4 and
Ag2WO4 outperformed their respective commercial counterparts, with a striking difference for CuWO4.
To confirm the reliability of UV/Vis spectrophotometry for
monitoring the MO degradation, LCMS was also employed to
follow the photocatalytic reaction. As shown in the example of
ZnWO4 (Na) in Figure 9, the ion peak area for MO (m/z = 304.07

1658

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers

Figure 8. MO photodegradation using the as-synthesized samples under UV/Vis light irradiation.

Figure 10. Comparison of the apparent rate constant values obtained from
linear regression of MO photodegradation plots (Figure 8).

Figure 9. LCMS monitoring of methyl orange photodegradation using the


as-synthesized ZnWO4 (Na) sample under UV/Vis light irradiation

in negative ion mode) decreased significantly with increasing


reaction time. More importantly, the degradation pattern is
similar to that derived from UV/Vis spectrophotometry: the
MO concentration reached zero within 40 min. Detailed mechanistic analysis of the degradation process is beyond the scope
of this study. Nonetheless, major intermediates were detected
using MS2, a tandem mass spectrometry technique for structural determination. An oxidative degradation pathway was identified similar to those presented for experiments using TiO2
earlier, involving demethylation and hydroxylation as intermediate steps.[49, 50]
Linear regression was performed for the first 40 min of each
photodegradation profile data (Table S2). The apparent rate
constants compared in Figure 10 suggest that the SCS samples
exhibited faster kinetics for MO dye degradation than their respective commercial samples, with the kapp values ordered
thus: ZnWO4 > CuWO4 ~ Ag2WO4. The enhanced degradation
rate (especially for ZnWO4) can be rooted in several factors
such as optical, electronic, structural/morphological, and surface chemical properties. The impact of the above factors,
however, is complex and convoluted. In the following sections
an attempt to deconvolute these factors is presented by elaborating on the previously shown XRD and TEM data, as well as
on photoelectrochemical and surface area measurements.
ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

Any discussion of the physicochemical attributes of a given


photocatalyst must start with its surface area [see Brunauer
EmmettTeller (BET) surface area data in Table S3] as this parameter manifests most directly in photocatalysis by dictating
the amount of substrate species that can be initially bound by
adsorption. In evaluating surface area contributions, however,
surface chemical properties also have to be taken into consideration because these factors strongly influence the adsorption
process. As the first step, the kinetics of the adsorption was
studied, and we found that the equilibrium was reached approximately within 4050 min (see Figure S5). Therefore,
a 60 min incubation period was employed before each photocatalysis (and adsorption) experiment. The amount of the adsorbed dye was determined for all tungstate samples (Figure S6). Considering the absolute values of the adsorbed dye
amounts, most values fall in the same range (1525 mmol g1),
except for CuWO4 (commercial) and Ag2WO4 (NH4) for which
much higher values were obtained (~ 65 mmol g1). The most
important message from the adsorption studies is that dye adsorption is not the rate-limiting step in our experiments. If that
were the case, the CuWO4 (commercial) and Ag2WO4 (NH4)
samples should have outperformed their counterparts because
of the vastly larger adsorbed dye amounts. But this was clearly
not the case; in fact, CuWO4 (commercial) showed the worst
photocatalytic activity among all studied samples (Figure 8).

1659

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers

Figure 11. A) Photovoltammogram of a CuWO4 (NH4) sample recorded between 0.1 and 1.1 V in 0.1 m Na2SO4, at a sweep rate of 2 mV s1 using a 300 W Hg
Xe arc lamp. B) Comparison of the onset potential for the three tungstate samples in 0.1 m Na2SO3.

These observations suggest that various factors other than surface area (and dye adsorption) play a role in the increased
photocatalytic activity of the SCS-synthesized tungstate nanoparticles.
To reveal possibly existing differences in the surface chemistry, surface area-normalized values of the adsorbed dye
amounts are also given in Figure S6. By using this approach
we can filter out the effect of higher surface area, and the adsorbed amounts reflect the affinity of MO to the studied tungstate surfaces only. Importantly, this representation reveals that
the two anomalously high dye loadings in the case of CuWO4
(commercial) and Ag2WO4 (NH4) stemmed solely from a surface-area effect because the normalized values are in good
agreement with those obtained for their counterparts. It is also
important that the normalized values are similar for each oxide
sample, independently from the precursor. If we compare the
different tungstates, a similar trend can be seen to that shown
in Figure 10 (ZnWO4 > CuWO4 ~ Ag2WO4), indicating that surface chemistry may play a role in the observed PC performance.
In solid-state science, crystallinity is well accepted to improve charge transport dynamics and, therefore, this parameter should be of importance in dictating the photocatalytic activity. However, this factor only seems to be a minor contributor in our case. Note, that samples with vastly different crystallinity [e.g., ZnWO4 (Na) and ZnWO4 (NH4)] show very similar
performance. We may assume that the previously mentioned
give-and-take relationship between phase purity versus crystallinity plays a key role even here. Similarly, the multiphasic composition (together with differences in the surface chemistry,
see Figure S6) of the commercial sample may well be the
reason for the inferior performance of these materials.
Differences among the oxidative photocatalytic activity of
the different metal tungstates can be understood by comparing their electronic properties, specifically their VB edge posiChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

tions. In this vein, linear sweep photovoltammograms were recorded for the various tungstate samples. This voltammetry
technique has been described elsewhere;[44] briefly, it employs
a slow potential scan while irradiation of the electrode is periodically interrupted. In this manner, both the dark and the
light-induced photoresponse of the samples can be assessed
in a single experiment. Interestingly, the best photoelectrochemical performance (i.e., highest photocurrents) was observed for CuWO4 (Figure 11 A). ZnWO4 generated stable but
small photocurrents whereas Ag2WO4 was unstable under
these conditions. Photovoltammograms, employing a narrower
potential window spanning the open-circuit potential under illumination (Figure 11 B), were also recorded in sulfite ion-containing electrolyte in which sulfite acted as the hole scavenger
(electron donor). The onset potential of these curves can be
associated with the Fermi level of the various semiconductor
samples, and thus their relative position can be correlated to
the position of the CB edge.[51]
The obtained values increase in the Zn, Ag, Cu-series (see
the arrows in Figure 11 pointing to 0.2, 0.1, and + 0.05 V respectively). Using the above data and taking into account the
determined band gap energies (also depicted in Scheme 1),
we may assume that the high photocatalytic activity for
ZnWO4 (compared to its Ag- and Cu-containing counterparts)
can be attributed to the more positive VB edge position and
thus the higher oxidation power of the photogenerated holes.
In contrast, despite the excellent photoelectrochemical properties (Figure 11 A) of CuWO4, it exhibits rather inferior photocatalytic behavior for our test reaction because of the less positive VB edge position.

Conclusions
In this study, binary tungstate nanoparticles of ZnWO4, CuWO4,
and Ag2WO4 were synthesized by solution combustion synthe-

1660

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers
sis (SCS) using two different precursors. Structural analysis
showed a pure monophasic structure forming for the samples
synthesized using the Na2WO4 precursor whereas a biphasic
structure was observed for samples when using (NH4)2WO4.
Most as-synthesized oxide samples were crystalline in nature;
however, the degree of crystallinity improved upon subsequent heat treatment. TEM images confirmed the formation of
various nanocrystalline structures for which the observed morphology strongly depended on the employed precursor molecule. The optical band gap energies (as determined by diffuse
reflectance UV/Vis spectroscopy) were in good agreement with
previous literature reports and our own computational data.
The as-prepared tungstate materials showed good photocatalytic activity (relatively high apparent rate constants) for the
photodegradation of methyl orange (MO) dye. Taking into account the specific surface area as well as the adsorbed dye
amount, it was revealed that factors other than surface area
[e.g., surface chemical properties, phase purity, and valence
band (VB) edge position] contributed to the increased photocatalytic activity of the materials. These conclusions were further supported by DFT (+ U) calculations and photoelectrochemical measurements. This latter technique (together with
the energy band gap values) allowed comparison of both the
conduction band (CB) and VB band edge positions of the
tungstate samples. Overall, we found that surface chemical
properties and the VB edge position were the two dominant
contributors to superior photocatalytic activity of SCS ZnWO4
compared to the other studied materials.
This study, in a broader vein, also illustrates the utility of inserting different transition metal ions into various structural
frameworks (e.g., WO3) and thus tuning their optical and optoelectronic attributes for targeted applications.[6] As demonstrated in this study, SCS is an extremely simple and versatile approach for this purpose. In this direction, further studies on the
synthesis of various narrow band gap binary oxide semiconductors (e.g., delafossites) for solar energy conversion and environmental remediation applications may be profitable; such
studies are in progress.

Experimental Section
Materials
Zinc nitrate hexahydrateZn(NO3)26 H2O (Alfa Aesar), copper(II) nitrate hemihydrate - Cu(NO3)22.5 H2O (Alfa Aesar), and silver nitrateAgNO3 (Fisher) were used as the zinc, copper, and silver precursors, respectively. Sodium tungsten oxide dihydrate
Na2WO42 H2O (Alfa Aesar) and ammonium tungsten oxide
(NH4)2WO4 (Alfa Aesar) were used as the tungsten precursors and
urea as the fuel. All chemicals were used without further purification. Double-distilled water (Corning Megapure) was used to prepare all solutions. Commercial samples of ZnWO4, CuWO4, and
Ag2WO4 (all Alfa Aesar) were used as reference materials for benchmarking the characteristics of the combustion-synthesized nanopowders. The synthesized samples are summarized in Table 3.
ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

Table 3. Synthesized samples prepared from different tungsten precursors.


Sample

Metal precursor

ZnWO4 (Na)
CuWO4 (Na)
Ag2WO4 (Na)
ZnWO4 (NH4)
CuWO4 (NH4)
Ag2WO4 (NH4)

Zn(NO3)26 H2O
Cu(NO3)22.5 H2O
AgNO3
Zn(NO3)26 H2O
Cu(NO3)22.5 H2O
AgNO3

Tungsten precursor

Fuel

Na2WO42 H2O
urea
(NH4)2WO4

Synthesis of tungstates
Stoichiometric amounts of the respective metal, the sodium tungstate precursor, and fuel were placed in a crucible and homogenized in double distilled water. The solution mixture was then transferred to a preheated furnace at 350 8C and left for 5 min. This
allows for dehydration of the precursor mixture and the promotion
of spontaneous combustion. As for ammonium tungstate, it was
first homogenized in 1.5 m HNO3 medium and then placed in a crucible with the metal precursor and fuel solution. The solution was
placed in a preheated furnace at 350 8C and left for 5 min.
Once the SCS process was completed, the samples were removed,
finely grinded in a mortar and pestle, and thermally annealed at
temperatures 400 8C, 450 8C, and 500 8C for 30 min. Thereafter,
each sample (both the as-synthesized and the heat treated) was
washed with double-distilled water to remove any soluble residue
from the precursor species, filtered, and dried in an oven at 100 8C.

Physical characterization
To model the SCS procedure, thermogravimetric analysis/differential scanning calorimetry (TGA/DSC) analyses on the precursor mixtures were carried out using a TA Instruments model Q600 instrument. The precursor mixtures were placed in an alumina crucible
in air atmosphere with a flow rate of 100 mL min1 and at a heating
rate of 10 8C min1 up to 1000 8C. Powder X-ray diffraction (XRD)
measurements were performed within the angle range 2q = 108
808 using a Rigaku Ultima IV instrument with a CuKa radiation
source (l = 1.5406). Rietveld refinements were carried out on MDI
Jade 8 software, using initial model structures from reported literature. Refinements were also performed on each of the samples
with reduced scale and background parameters and individual
FWHM (full width half maximum) curves to account for all the
peaks in the XRD pattern. The peak shapes were fitted using the
Gaussian profile function with a displacement selection used allowing for any deviations from the model structure. UV-vis diffuse reflectance spectra were collected on a PerkinElmer Lambda 35 spectrophotometer equipped with an integrating sphere. High-resolution transmission electron microscopy (HRTEM) was performed on
a Hitachi H-9500 instrument at various magnifications. Raman
spectra were obtained from a Horiba Jobin Yvon Labram Aramis
spectrometer at an excitation wavelength of 633 nm, using a He
Ne laser by averaging 64 spectra. For BET surface area, a Quantachrome NOVA 3000 adsorption test instrument was utilized for the
measurements.

Computational details
The present calculations were performed within the framework of
the standard frozen-core projector augmented-wave (PAW)
method[52, 53] using DFT as implemented in Vienna ab initio simula-

1661

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers
tion package (VASP 4.6) code.[54, 55] In the PAW method, a nonlinear
core-correction is not necessary because it is an all-electron-like
method. Exchange and correlation potentials were treated in the
generalized gradient approximation (GGA) as parameterized by
Perdew-Burke-Ernzerhof (PBE).[56, 57] The PBE functional does not
contain any empirically optimized parameters, and hence works
better on a wide range of elements. It is well known that underestimation of electron localization is a major failure of standard DFT
calculations, in particular, for systems with localized d and f electrons.[5860] This failure manifests in the general trend of DFT to underestimate the band gap and to produce incorrect solutions for
some 3d based metal oxides. In order to correct this shortcoming,
in our calculations we employed an on-site Coulomb correlation
through the Hubbard-based U correction parameter.[6062] In the
present work, we used U = 7 eV externally providing Coulomb correlation to only Cu 3d and Zn 3d orbitals. Hence, all results presented here are based on DFT + U calculations for both ZnWO4
and CuWO4 and DFT calculation for Ag2WO4. The basis sets were
expanded with plane-waves with a kinetic energy cut-off of 400 eV,
and the BZ integrations were performed using the second-order
Methfessel-Paxton method.[63] The optimization of ZnWO4, CuWO4,
and Ag2WO4, was done using 11 9 11, 5 9 9, and 5 5 9 Monkhorst-Pack k-point sampling.[64] However, more refined 9 13 13,
15 13 15, and 7 7 11, k-point samplings were used for ZnWO4,
CuWO4, and Ag2WO4 respectively for optical property calculations.
For visualization of the crystal structures, VESTA (Visualization for
Electronic and Structural Analysis) was used.[65, 66] Calculations were
performed at the High Performance Computing (HPC) Center at
the University of Texas at Arlington.

Photocatalysis experiments
The oxidative photocatalytic activity of the tungstate nanoparticles
was evaluated by their ability to degrade methyl orange (MO) as
a probe dye in aqueous solution. A photocatalytic reactor consisting of an outer cylindrical glass vessel (400700 nm transmittance)
and an interchangeable quartz/glass inner vessel was used for the
photocatalysis process. The light source, a 400 W medium pressure
Hg arc (incident photon flux estimated by potassium ferrioxalate
actinometry was 5.21 104 Einstein per minute), was placed in the
inner vessel, which was also equipped for water circulation aimed
at maintaining a constant reactor temperature, as well as filtering
infrared radiation. The outer reaction vessel was equipped with
a sample collection port, in addition to a gas purge connection
and a stirrer to ensure dispersion of the mixture and to prevent
the catalyst from sedimentation.
Testing of the photocatalytic activity was carried out using
a 250 mL MO (50 mm) reaction solution along with the photocatalyst (dose: 2 g L1) while being purged with O2 gas. The reaction
suspension was initially stirred in the dark for 60 min to attain adsorption equilibrium of MO on the catalyst surface. After 60 min,
the solution was irradiated with UV-visible light. Sample aliquots of
5 mL were periodically removed from the outer vessel every
10 min with the suspension being centrifuged and filtered
(0.45 mm PVDF membrane filter) to remove the photocatalyst
powder.

Detection of degradation
The extent of MO degradation was assessed by measuring the absorbance of the solution using an Agilent 8453 UV-visible spectrophotometer. Blank runs were performed under identical experimental conditions without the photocatalyst.
ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

The photocatalytic process was also monitored by LC-MS analyses,


performed on a Shimadzu LCMS-ITTOF (ion traptime-of-flight)
mass spectrometer (Shimadzu Scientific Instruments, Inc., Columbia MD).[67] Samples were taken from the reaction vessel at different times (t = 0, 10, 20, 30, 40, 50, 60, 70, and 80 min) during the
reaction and were analyzed to determine the degradation of
methyl orange and the formation of intermediates and products.
30 mL of sample was injected onto a reverse-phase column (Raptor
C18, 2.1 100 mm; 2.7 mm, Restek Corporation, Bellefonte, PA, USA)
and binary pumps carrying mobile phase of 10 mm ammonium
acetate or 0.1 % formic acid and acetonitrile with a flow rate of
0.3 mL min1 were used to obtain separation for scouting purposes.
A 10 min gradient was developed with 599 % acetonitrile followed by the washing and equilibrating steps. The eluent entered
an electrospray ionization (ESI) interface, which was operated in
either positive or negative ionization mode. Interface voltages for
positive or negative ionization modes were + 4.50 kV or 4.00 kV,
respectively. Mass range was set to record from 100500 m/z. The
curved desolvation transfer line and heat block temperatures were
both set at 200 8C. Nebulizing gas (nitrogen) was flowed at
1.5 L min1, and the detector voltage was set at 1.64 kV.
MS2 was performed to identify the major intermediates formed
during the photocatalytic reaction. Precursor ion isolation width
was set at 3.0000 Da with 10 ms ion accumulation time. Collision
energy was at 27 % and collision gas was at 50 % with a frequency
of 45.0 kHz.

Photoelectrochemical measurements
All photoelectrochemical measurements were performed on an Autolab PGSTAT302 instrument, in a classical one-compartment,
three-electrode electrochemical cell. Various tungstate nanoparticles were spray coated on ITO glass electrodes (ca. 0.1 mg cm2)
and were used as working electrodes. A large Pt foil counter-electrode and an Ag/AgCl/3 m KCl reference electrode completed the
cell setup. The light source was a 300 W Hg-Xe arc lamp (Hamamatsu L8251). The radiation source was placed 2 cm away from
the working electrode surface. Photovoltammetry profiles were recorded in both 0.1 m Na2SO3 and 0.1 m Na2SO4 electrolyte, using
a slow potential sweep (2 mV s1) in conjunction with interrupted
irradiation (0.1 Hz) on the semiconductor coated electrodes. All experiments were performed at the laboratory ambient temperature
(20 2 8C).

Acknowledgements
The authors thank Dr. Jiechao Jiang, facility manager of the UTA
Characterization Center for Materials and Biology, for his technical assistance with the HR-TEM images. KR thanks the National
Science Foundation (CHE-1303803) for partial funding support. CJ
thanks the Hungarian Academy of Sciences for financial support
through the Momentum Excellence Program. The work of MNH
was supported partially by National Science Foundation (CBET1133672). E.H.W. and K.A.S. thank Restek Corporation for research
support. Finally, we thank the two anonymous reviewers for constructive criticisms of an earlier manuscript version.
Keywords: bandgap engineering combustion synthesis
photocatalysis semiconductors solar energy

1662

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Full Papers
[1] R. M. Navarro Yerga, M. C. Alvarez Galvn, F. del Valle, J. a Villoria de La
Mano, J. L. G. Fierro, ChemSusChem 2009, 2, 471 485.
[2] K. Rajeshwar, J. Phys. Chem. Lett. 2011, 2, 1301 1309.
[3] K. Rajeshwar, A. Thomas, C. Janky, J. Phys. Chem. Lett. 2015, 6, 139
147.
[4] A. Linsebigler, G. Lu, J. Yates, Chem. Rev. 1995, 95, 735 758.
[5] W. Morales, M. Cason, O. Aina, N. R. de Tacconi, K. Rajeshwar, J. Am.
Chem. Soc. 2008, 130, 6318 6319.
[6] C. Janky, K. Rajeshwar, N. R. de Tacconi, W. Chanmanee, M. N. Huda,
Catal. Today 2013, 199, 53 64.
[7] J. Tian, Y. Sang, G. Yu, H. Jiang, X. Mu, H. Liu, Adv. Mater. 2013, 25,
5075 5080.
[8] S. Sun, W. Wang, L. Zhang, E. Gao, D. Jiang, Y. Sun, Y. Xie, ChemSusChem
2013, 6, 1873 1877.
[9] L. Zhang, D. Bahnemann, ChemSusChem 2013, 6, 283 290.
[10] I.-S. Cho, C. H. Kwak, D. W. Kim, S. Lee, K. S. Hong, J. Phys. Chem. C
2009, 113, 10647 10653.
[11] J. Ungelenk, C. Feldmann, Chem. Commun. 2012, 48, 7838 7840.
[12] H. Fu, J. Lin, L. Zhang, Y. Zhu, Appl. Catal. A 2006, 306, 58 67.
[13] R. Shi, Y. Wang, D. Li, J. Xu, Y. Zhu, Appl. Catal. B 2010, 100, 173 178.
[14] C. Zhang, H. Zhang, K. Zhang, X. Li, Q. Leng, C. Hu, ACS Appl. Mater. Interfaces 2014, 6, 14423 14432.
[15] N. R. de Tacconi, H. K. Timmaji, W. Chanmanee, M. N. Huda, P. Sarker, C.
Janky, K. Rajeshwar, ChemPhysChem 2012, 13, 2945 2955.
[16] J. Tang, Z. Zou, J. Ye, J. Phys. Chem. B 2003, 107, 14265 14269.
[17] S. Song, Y. Zhang, Y. Xing, C. Wang, J. Feng, W. Shi, G. Zheng, H. Zhang,
Adv. Funct. Mater. 2008, 18, 2328 2334.
[18] D. W. Kim, I.-S. Cho, S. S. Shin, S. Lee, T. H. Noh, D. H. Kim, H. S. Jung,
K. S. Hong, J. Solid State Chem. 2011, 184, 2103 2107.
[19] L. Kihlborg, E. Gebert, Acta Crystallogr. Sect. B 1970, 26, 1020 1026.
[20] J. Forsyth, C. Wilkinson, A. Zvyagin, J. Phys. Condens. Matter 1991, 3,
8433 8440.
[21] J. E. Yourey, B. M. Bartlett, J. Mater. Chem. 2011, 21, 7651 7660.
[22] A. Van den Berg, C. Juffermans, J. Appl. Crystallogr. 1982, 15, 114 116.
[23] L. Cavalcante, M. Almeida, W. Avansi, R. Tranquilin, E. Longo, N. Batista,
V. Mastelaro, S. Li, Inorg. Chem. 2012, 51, 10675 10687.
[24] G. B. Kumar, K. Sivaiah, S. Buddhudu, Ceram. Int. 2010, 36, 199 202.
[25] G. Huang, C. Zhang, Y. Zhu, J. Alloys Compd. 2007, 432, 269 276.
[26] P. Siriwong, T. Thongtem, A. Phuruangrat, S. Thongtem, CrystEngComm
2011, 13, 1564 1569.
[27] J. Tang, J. Ye, J. Mater. Chem. 2005, 15, 4246 4251.
[28] E. Longo, D. Volanti, V. Longo, L. Garcia, I. Nogueira, M. A. P. Almeida, A.
Pinheiro, M. Ferrer, L. S. Cavalcante, J. Andres, J. Phys. Chem. C 2014,
118, 1229 1239.
[29] M. Hojamberdiev, G. Zhu, Y. Xu, Mater. Res. Bull. 2010, 45, 1934 1940.
[30] R. Zhang, H. Cui, X. Yang, H. Liu, H. Tang, Y. Li, Micro Nano Lett. 2012, 7,
1285 1288.
[31] A. G. Merzhanov, J. Mater. Chem. 2004, 14, 1779 1786.
[32] K. C. Patil, M. S. Hegde, T. Rattan, S. T. Aruna in Chemistry of Nanocrystalline Oxide Materials: Combustion Synthesis Properties and Applications,
World Scientific Publishing, Singapore, 2008.
[33] K. Rajeshwar, N. R. de Tacconi, Chem. Soc. Rev. 2009, 38, 1984 1998.
[34] K. Nagaveni, M. S. Hegde, N. Ravishankar, G. N. Subbanna, G. Madras,
Langmuir 2004, 20, 2900 2907.

ChemSusChem 2015, 8, 1652 1663

www.chemsuschem.org

[35] G. Sivalingam, K. Nagaveni, M. S. Hegde, G. Madras, Appl. Catal. B 2003,


45, 23 38.
[36] S. Kikkawa, S. Hosokawa, H. Ogawa, J. Am. Ceram. Soc. 2005, 88, 308
311.
[37] K. Deshpande, A. Mukasyan, A. Varma, Chem. Mater. 2004, 16, 4896
4904.
[38] C.-L. Li, Z.-W. Fu, Electrochim. Acta 2008, 53, 4293 4301.
[39] H. Shim, I. Cho, K. Hong, A.-H. Lim, D.-W. Kim, J. Phys. Chem. C 2011,
115, 16228 16233.
[40] H.-W. Shim, A.-H. Lim, G.-H. Lee, H.-C. Jung, D.-W. Kim, Nanoscale Res.
Lett. 2012, 7, 9 16.
[41] Y. Liu, H. Wang, G. Chen, Y. D. Zhou, B. Y. Gu, B. Q. Hu, J. Appl. Phys.
1988, 64, 4651 4653.
[42] E. Ross-Medgaarden, I. Wachs, J. Phys. Chem. C 2007, 111, 15089 15099.
[43] T. T. Basiev, A. Y. Karasik, A. A. Sobol, D. S. Chunaev, V. E. Shukshin, Quantum Electron. 2011, 41, 370 372.
[44] C. Janky, N. R. de Tacconi, W. Chanmanee, K. Rajeshwar, J. Phys. Chem.
C 2012, 116, 4234 4242.
[45] A. B. Murphy, Sol. Energy Mater. Sol. Cells 2007, 91, 1326 1337.
[46] T. Montini, V. Gombac, A. Hameed, L. Felisari, G. Adami, P. Fornsaiero,
Chem. Phys. Lett. 2010, 498, 113 119.
[47] J. Ruiz-Fuertes, D. Errandonea, A. Segura, F. J. Manjon, Z. Zhu, C. Y. Tu,
High Pressure Res. 2008, 28, 565 570.
[48] S. J. Naik, A. V. Salker, Solid State Sci. 2010, 12, 2065 2072.
[49] C. Baiocchi, M. C. Brussino, E. Pramauro, A. B. Prevot, L. Palmisano, G.
Marc, Int. J. Mass Spectrom. 2002, 214, 247 256.
[50] K. Dai, H. Chen, T. Peng, D. Ke, H. Yi, Chemosphere 2007, 69, 1361 1367.
[51] K. Rajeshwar in Electron Transf. Chem. (Ed.: V. Balzani), Wiley- VCH, Weinheim, 2001.
[52] G. Kresse, D. Joubert, Phys. Rev. B 1999, 59, 11 19.
[53] P. Blchl, Phys. Rev. B 1994, 50, 17953 17979.
[54] G. Kresse, J. Furthmller, Phys. Rev. B 1996, 54, 11169 11186.
[55] G. Kresse, J. Furthmller, Comput. Mater. Sci. 1996, 6, 15 50.
[56] J. Perdew, J. Chevary, S. Vosko, Phys. Rev. B 1992, 46, 6671 6687.
[57] J. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 1996, 77, 3865 3868.
[58] F. Zhou, C. Marianetti, M. Cococcioni, D. Morgan, G. Ceder, Phys. Rev. B
2004, 69, 201101.
[59] M. V. Ganduglia-Pirovano, A. Hofmann, J. Sauer, Surf. Sci. Rep. 2007, 62,
219 270.
[60] V. Anisimov, J. Zaanen, O. Andersen, Phys. Rev. B 1991, 44, 943 954.
[61] V. Anisimov, I. Solovyev, M. Korotin, M. Czyzyk, G. Sawatzky, Phys. Rev. B
1993, 48, 16929 16934.
[62] I. Solovyev, P. Dederichs, V. Anisimov, Phys. Rev. B 1994, 50, 16861
16871.
[63] M. Methfessel, A. T. Paxton, Phys. Rev. B 1989, 40, 3616 3621.
[64] H. Monkhorst, J. Pack, Phys. Rev. B 1976, 13, 5188 5192.
[65] K. Momma, F. Izumi, J. Appl. Crystallogr. 2008, 41, 653 658.
[66] F. Izumi, K. Momma, Solid State Phenom. 2007, 130, 15 20.
[67] J. S. Barnes, K. A. Schug, J. Agric. Food Chem. 2014, 62, 4322 4331.

Received: March 17, 2015

1663

2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like