Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Accepted Manuscript

Electrochemically-synthesized tungstate nanocomposites γ-WO3/CuWO4 and γ-


WO3/NiWO4 thin films with improved band gap and photoactivity for solar-driven
photoelectrochemical water oxidation

Tao Zhu, Meng Nan Chong, Eng Seng Chan, Joey D. Ocon
PII: S0925-8388(18)31863-2
DOI: 10.1016/j.jallcom.2018.05.147
Reference: JALCOM 46123

To appear in: Journal of Alloys and Compounds

Received Date: 1 February 2018


Revised Date: 30 April 2018
Accepted Date: 13 May 2018

Please cite this article as: T. Zhu, M.N. Chong, E.S. Chan, J.D. Ocon, Electrochemically-synthesized
tungstate nanocomposites γ-WO3/CuWO4 and γ-WO3/NiWO4 thin films with improved band gap and
photoactivity for solar-driven photoelectrochemical water oxidation, Journal of Alloys and Compounds
(2018), doi: 10.1016/j.jallcom.2018.05.147.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Graphical abstract

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Electrochemically-synthesized Tungstate Nanocomposites γ-WO3/CuWO4 and

γ-WO3/NiWO4 Thin Films with Improved Band Gap and Photoactivity for

Solar-driven Photoelectrochemical Water Oxidation

PT
RI
Tao Zhu1, Meng Nan Chong1,2*, Eng Seng Chan1, Joey D. Ocon3

SC
1
School of Engineering, Chemical Engineering Discipline, Monash University Malaysia, Jalan Lagoon Selatan,

U
Bandar Sunway, Selangor Darul Ehsan 47500, Malaysia
AN
2
Sustainable Water Alliance, Advanced Engineering Platform, Monash University Malaysia, Jalan Lagoon Selatan,

Bandar Sunway, Selangor Darul Ehsan 47500, Malaysia


M

3
Laboratory of Electrochemical Engineering (LEE), Department of Chemical Engineering, University of the

Philippines Diliman, Quezon City 1101, Philippines


D

*Corresponding author: Associate Professor Dr. Meng Nan Chong


TE

School of Engineering, Chemical Engineering Discipline, Monash University Malaysia,


EP

Jalan Lagoon Selatan, Bandar Sunway, Selangor Darul Ehsan 47500, Malaysia

Tel: +603 5514 5680, Fax: +603 5514 6207, Email: Chong.Meng.Nan@monash.edu
C
AC
ACCEPTED MANUSCRIPT

Abstract

The main aim of this study was to synthesize and characterise tungstate (WO3) nanocomposites

with its metal-based nanostructures, such as copper (II) tungstate (CuWO4) and nickel tungsten

PT
oxide (NiWO4), as visible-light active thin film photoanodes for solar-driven

photoelectrochemical (PEC) water oxidation. FE-SEM and AFM results showed that the bare as-

RI
deposited WO3 films were transformed into polycrystalline WO3 structure with highly

agglomerated surfaces and roughness during the annealing-induced crystallisation process. XRD

SC
results suggested that the bare as-deposited WO3 films undergone phase transformation process

U
from amorphous to the photoactive monoclinic-I (γ-WO3) at 550 oC. XPS results indicated the
AN
existence of WO42-, Ni2+ and Cu2+ ions at 35.58 eV, 856 eV and 932.4 eV, respectively. Through

the formation of WO3 nanocomposites, the energy band gap was effectively lowered from 2.7 eV
M

(γ-WO3)  2.3 eV (γ-WO3/CuWO4)  2.1 eV (γ-WO3/NiWO4) as estimated from the UV-Vis

spectra. Finally, the corresponding photoactivity of WO3 nanocomposites was estimated by


D

measuring the photocurrent density and γ-WO3/NiWO4 nanocomposite structure was found to
TE

give the highest photocurrent density of 400 µA/cm2 at 1.5 V vs Ag/AgCl (4M KCl).
EP

Keywords: Photoelectrochemical water splitting; Cathodic electrodeposition; Photoelectrocatalyst; Tungsten


C

trioxide; Solar hydrogen fuel.


AC
ACCEPTED MANUSCRIPT

1. Introduction

With the rapid development of modern society, the demands on traditional energy fuels such as

coal, petroleum and natural gas are increasingly needed. However, these energy fuels are non-

PT
renewable and thus, it is important to explore and develop alternative energy resources that are

renewable, sustainable and economical to meet these ever increasing energy demands. In recent

RI
years, the utilisation and harnessing of solar energy are gaining momentum owing to its abundant

and clean energy nature [1]. Among the various methods used for solar energy capture and

SC
conversion, the solar-driven photoelectrochemical (PEC) water splitting process is a promising

U
method that can convert solar into storable form of chemical energy (i.e. hydrogen fuel). To date,
AN
many researches have been carried out on solar energy conversion using PEC process but it is

still hindered by issues associated with low energy conversion efficiency [2-4].
M

In a conventional PEC cell, it has three main components of a working electrode, a counter
D

electrode and a reference electrode. The working electrode (i.e. photoanode) is usually made of

nanostructured semiconductor-based photocatalytic materials. This semiconductor-based


TE

photoanode plays an important role in the conversion of solar energy, through the three
EP

fundamental mechanistic steps of: (1) absorption of light photon energy that is greater than or

equal to its band gap in generating electron-hole pairs; (2) separation of photogenerated electron-
C

hole pairs and; (3) oxidation of water molecules at photoanode and reduction of H+ ions to
AC

produce molecular H2. Thus, the development of new photoanodes with high efficiency and

stability is extremely important for resolving the issues associated with low energy conversion

efficiency in PEC water splitting process. Among all, semiconductor tungstate (WO3) is a

suitable photoanode candidate owing to its suitable band gap of 2.7-2.8 eV that can absorb up to

460 nm in the visible light spectrum [5]. Additionally, WO3 exhibits high resistance and
ACCEPTED MANUSCRIPT

chemical stability in acidic solution. However, the applicability of WO3 photoanode is still

limited by its high recombination rate of photogenerated excitons.

In recent years, many researches have been attempted to address the mechanistic issue of high

PT
recombination rate of photogenerated excitons by promoting their efficient and prolonged

separation but with limited success. Several tungstate metal-based nanostructures, such as

RI
CuWO4, NiWO4 and, Bi2WO6 have been investigated for their physicochemical properties and

water splitting potential [5-7]. For example, CuWO4 exhibited high stability in photochemical

SC
cell under illumination and has a very well positioned valence band for water oxidation with a

U
band gap of 2.3 eV. CuWO4 has a flat band potential and carrier density of +0.4 V (vs NHE) and
AN
2.7×1021 cm-3 respectively, which is suitable to be used as the photoanode material for water

oxidation [8]. Another potential candidate for water oxidation is NiWO4, which has a lower band
M

gap of 2 eV. Previously, it was reported that the NiWO4/WO3 heterojunction structure offers

enhanced photoconversion efficiency and increased the density of carriers that resulted in
D

enhancing PEC performance [9-10].


TE

The main aim of this study was to synthesize and characterise tungstate (WO3) nanocomposites
EP

with its metal-based nanostructures, such as copper (II) tungstate (CuWO4) and nickel tungsten

oxide (NiWO4), as visible-light active thin film photoanodes for efficient solar-driven PEC water
C

oxidation. Initially, the bare as-deposited WO3 thin films were electrochemically-synthesized on
AC

fluorine-doped tin oxide (FTO) at a constant applied potential of -0.45 V. This was followed by

forming nanocomposites by covering the surface of bare as-deposited WO3 films with Cu(NO3)2

and NiNO3 solution and annealing treatment at 550 oC for 6 h. The WO3 nanocomposites were

characterised by using: field emission-scanning electron microscopy (FE-SEM), atomic force

microscopy (AFM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and
ACCEPTED MANUSCRIPT

ultraviolet-visible (UV-vis) spectrophotometry. It is hopeful that the finding from this study

could provide a good understanding on tungstate nanocomposites, and achieving good

photoactivity through rational design and effective band gap engineering of tungstate

nanocomposites for solar-driven PEC water oxidation.

PT
RI
2. Experimental sections

SC
2.1. Materials

All the chemicals were used as received without further purification. Hydrogen peroxide (H2O2)

U
(30 %) was obtained from HmbG Chemicals, USA. Tungsten (W) powder with particle size of
AN
325 meshes was purchased from Chem Soln, USA. Platinum (Pt) black (≥ 99.97 %) with particle
M

size ≤ 20 µm was supplied by ChemSoln, USA. All other chemicals were purchased from Merck,

USA. The electrodeposition synthesis electrolyte was prepared according to our previously
D

published study [4]. Initially, the electrolyte used for electrodeposition was prepared by
TE

dissolving 1.8 g of W powder in 50 mL of H2O2 and keep dissolving up to 24 h. After that, the

excess H2O2 was decomposed by adding a small amount of Pt black. The solution was further
EP

heated at 60 ℃ until no gas bubble was evident. Then the electrolyte solution was diluted to 50

mM via the addition of 150 mL of 50/50 (v/v) water/2-propanol. In this instance, the function of
C

propanol-2-ol was used to extend the stability of electrolyte solution by preventing the
AC

precipitation of an amorphous WO3-based hydrated phase.


ACCEPTED MANUSCRIPT

2.2. Preparation of WO3 nanocomposites

The electrodeposition synthesis of nanostructured WO3 thin films was performed at room

temperature using a conventional three-electrode electrochemical cell system of PGSTAT204

PT
Applied Potentiostat (Metrohm, Netherlands). FTO glass slide (ChemSoln, USA; 14 Ω/sq; 2.5

cm×1.5 cm) was used as the working electrode after being cleaned with acetone and extra pure

RI
water, while Pt was used as the counter electrode and Ag/AgCl (4M KCl) as the reference

electrode. All the measured potentials were made reference to the Ag/AgCl (4M KCl) electrode.

SC
During the electrodeposition synthesis process, the FTO area immersed in the electrolyte was

U
fixed constant at 2 cm×1.5 cm. The applied potential between the working and reference
AN
electrode was fixed at -0.45 V. After an electrodeposition synthesis cycle of 15 min, the as-

deposited amorphous WO3 film on FTO was rinsed by using distilled water and subsequently
M

dried by using a hot-air gun with the controlled temperature of 100 oC for 1 min. The dried WO3

film was further subjected to another electrodeposition synthesis cycle of 15 min. This procedure
D

was repeated for three times before the inclusion of metal-based tungstate nanostructures as the
TE

top-layer. Subsequently, the as-deposited WO3 films were annealed at 550 oC for 6 h (i.e. at

heating and cooling rates of 10 oC/min and 2.5 oC/min, respectively).


EP

As for the synthesis of WO3/NiWO4 film, 100 µL of 0.2 M Ni(NO3)2 aqueous solution was
C

applied over the surface of the as-deposited WO3 films. This was followed by annealing the
AC

nanocomposite films at 550 oC for 6 h. After the cooling down of WO3/NiWO4 films, they were

soaked in 0.5 M HCl bath for approximately 15 min in order to dissolve the excess NiO formed

on the film surfaces. Finally, the WO3/NiWO4 films were washed by using deionised water and

dried at 120 oC for 2 h. Similar procedures were adopted for the synthesis of WO3/CuWO4 films,

except that 100 µL of 0.2 M Cu(NO3)2•3H2O aqueous solution was used.


ACCEPTED MANUSCRIPT

2.3. Characterisation of WO3 nanostructures

The characterisation of surface morphology of the nanostructured WO3 thin films and the

nanocomposite WO3/CuWO4 and WO3/NiWO4 films was investigated by FESEM (FEI Nova

PT
NanoSEM). AFM was used to study the surface characteristics, such as topography and grain

size of the various WO3 nanostructures formed (AFM, Digital Instrument, Nanoscope III). XRD

RI
measurements were carried out at room temperature using Cu Kα radiation (λ = 1.54 Å) with

potential of 40 kV and a current of 30 mA (Philips PW1830, Netherlands). The surface elemental

SC
analysis was carried out by using X-ray photoelectron spectroscopy (XPS, K-Alpha 1063,

U
ThermoFisher Scientific). UV-vis absorption
AN spectra were recorded by using a

spectrophotometer (DR-UVS, Shimadzu 2450 spectrophotometer). The PEC properties were

measured at room temperature in a dark box using the same Autolab potentiostat/galvanostat
M

system with nanostructured WO3 films on FTO used as the working electrode. For the

measurements of PEC properties, the peroxy-tungstic acid (PTA) electrolyte solution was
D

replaced by 0.1 M sodium acetate (CH3COONa) aqueous solution. A 100W halogen lamp
TE

restricted at a frequency of 0.05 Hz was used as the light source. The linear potentiodynamic

voltammetry was applied at a scan rate of 5 mV/s with a step size of 1 mV.
C EP

3. Results and discussion


AC

3.1. FE-SEM

The surface morphology of the nanostructured WO3, WO3/CuWO4 and WO3/NiWO4 films was

characterised by using FE-SEM, as shown in Figure 1. Previously, we have discussed on the

influence of annealing-induced crystallisation process on nanostructured WO3 thin films,


ACCEPTED MANUSCRIPT

including the surface morphologies, polycrystalline structures and transformation between active

WO3 phases [3]. Figure 1(a) and (b) show the FE-SEM images of WO3 films before and after the

annealing-induced crystallisation process at 550 oC for 6 h, respectively.

PT
(a) (b)

RI
U SC
AN
(c) (d)
M
D
TE
C EP

Figure 1. FE-SEM images of WO3 films with or without nanocompositing with CuWO4 and
AC

NiWO4 films: (a) as-deposited WO3 film; (b) nanostructured WO3 thin film; (c) nanocomposite
WO3/CuWO4 thin film; and, (d) nanocomposite WO3/NiWO4 thin film. All the thin films (b, c,
d) were annealed at 550 oC for 6 h.

From Figure 1, it can be observed that the as-deposited WO3 film was transformed into

polycrystalline WO3 structure with highly agglomerated surface during the annealing-induced
ACCEPTED MANUSCRIPT

crystallisation process. This was predominantly due to the water loss phenomenon during the

annealing process, where this could lead to surface defects (i.e. cracks) as previously

communicated [3]. Another potential reason that resulted in the surface defects (i.e. cracks)

maybe due to the prolonged annealing treatment of 6 h than 20 min previously used [3]. From

PT
this study, it was measured that the average particle size for the as-deposited WO3 was

RI
approximately 40 nm. Owing to the highly agglomerated surface after annealing treatment,

however, it was quite impossible to determine the average WO3 particle size. In comparison,

SC
both the nanocomposite films of WO3/CuWO4 (Figure 1c) and WO3/NiWO4 (Figure 1d)

exhibited more orderly, particle-like and interconnected porous structures that can enhance the

U
absorption ability of ions in the electrolyte. For these nanocomposite structures, the average
AN
particle size for WO3/CuWO4 and WO3/NiWO4 was determined to be 80 nm and 70 – 100 nm,

respectively.
M
D

3.2. AFM
TE

Both the two- (2D) and three-dimensional (3D, insets) surface characteristics of nanostructured
EP

WO3 thin films were probed by using AFM as shown in Figure 2. In this instance, the surface

roughness was expressed in terms of the root-mean-square (Rrms) value. Figure 2(a) and (b) show
C

that the as-deposited WO3 film had a columnar structure and undergone the annealing-induced
AC

crystallisation process at 550 oC to form the polycrystalline WO3 structure. During the

crystallisation process, the Rrms values were seen to increase diminutively from the as-deposited

WO3 film of 7.42 nm to the polycrystalline WO3 thin film (i.e. annealed at 550 oC) of 7.61 nm.

The small increase in Rrms values suggested that electrodeposition followed by annealing-induced

crystallisation process is a good synthesis method for the preparation of nanostructured WO3 thin
ACCEPTED MANUSCRIPT

films with high surface uniformity. The Rrms values were increased after the annealing process

that may be attributed to the grain growth and agglomeration phenomena. Similar results were

previously reported by Saleem et al. [10], where they related the findings to surface species

having attained sufficient thermal energy and migrated towards the energy favorable sites for

PT
WO3 grains growth. However, the surface roughness of nanostructured WO3 thin films was

RI
further reduced at higher annealing temperature that may be owing to the decrease in WO3 grain

boundaries [10].

U SC
AN
M
D
TE

(d)
C EP
AC
ACCEPTED MANUSCRIPT

Figure 2. AFM images of 2D and 3D (insets) surface characteristics of nanostructured WO3 thin
films: (a) as-deposited WO3 film; (b) nanostructured WO3 thin film; (c) nanocomposite
WO3/CuWO4 thin film; and (d) nanocomposite WO3/NiWO4 thin film. All the thin films (b, c, d)
were annealed at 550 oC for 6 h.
As a result of the water loss phenomenon during the annealing process, a larger average

PT
nanoparticles size was measured on Figure 2(b) of nanostructured WO3 thin film that was in

agreement with the FE-SEM analysis. Meanwhile, the surface roughness for nanocomposite

RI
WO3/CuWO4 film was measured to be 13.8 nm owing to the presence of interconnected porous

SC
structure. In comparison, however, the surface roughness for the nanocomposite WO3/NiWO4

(i.e. 16.1 nm) was greater than the WO3/CuWO4 structure as shown in Figure 2(d) indicating a

U
more extensive and highly interconnected porous structure.
AN
M

3.3. XRD

Figure 3 shows the XRD spectra for nanostructured WO3, nanocomposite WO3/CuWO4 and
D

WO3/NiWO4 thin films. From Figure 3, the characteristic peaks of FTO are indicated by the
TE

black square spots. It can be seen that the as-deposited WO3 film showed no distinctive and

sharp diffraction peaks, indicating the amorphous nature of the WO3 film [11]. In comparison,
EP

the annealed WO3 thin film at 550 oC showed distinctive and sharp diffraction peaks at 23.3o,

23.8o and 24.6o that are indexed to the {002}, {020} and {200} WO3 planes. These suggested
C
AC

that the WO3 films undergone phase transformation process from amorphous to the photoactive

monoclinic-I (γ-WO3) (JCPDS 43-1035) at 550 oC [12].

As for the nanocomposite γ-WO3/CuWO4 thin film annealed at 550 oC, three distinctive and

sharp diffraction peaks at 23.3o, 23.8o and 24.6o were evidenced. Previous studies have reported

that no diffraction peaks were evidenced for CuWO4 below 500 oC, and these were known to be
ACCEPTED MANUSCRIPT

highly dependent on the combination of synthesis method and annealing temperature used [13-

14]. Since the diffraction peaks for both WO3 and CuWO4 appeared to be identical owing to their

similar crystal structures [15], the presence of CuWO4 in the nanocomposite structure was

proven using XPS analysis in the next section. On the contrary, the diffraction pattern for NiWO4

PT
was exclusive whereby characteristic XRD peaks were evidenced at 30.80o, 35.1o and 54.51o that

RI
corresponded to the {110}, {111} and {202} of the monoclinic NiWO4 crystalline planes

(JCPDS 15-0755) [16-17].

U SC
AN
M
D
TE
EP

Figure 3. XRD spectra for nanostructured WO3, nanocomposite WO3/CuWO4 and WO3/NiWO4
thin films: as-deposited WO3 film (black line), nanostructure WO3 thin film (blue line),
C

nanocomposite WO3/CuWO4 thin film (red line), and nanocomposite WO3/NiWO4 thin film
(green line). All the thin films (except the as-deposited WO3 film) were annealed at 550 oC for 6
AC

h.

3.4. XPS
ACCEPTED MANUSCRIPT

XPS analysis was used to verify the existence of key chemical elements and investigate their

corresponding valence states in the nanostructured WO3 thin film, and nanocomposites γ-

WO3/CuWO4 and γ-WO3/NiWO4 thin films annealed at 550 oC and as shown in Figure 4: (a)

wide XPS scan on nanostructured WO3 thin film; (b) W 4f nanostructured WO3 thin film; (c) W

PT
4f nanocomposite

RI
γ-WO3/NiWO4 thin film; (d)

W 4f nanocomposite

SC
γ-WO3/CuWO4 thin film; and

(e) Ni 2p nanocomposite

γ-WO3/NiWO4
U thin film; (f)
AN
Cu 2p nanocomposite

γ-WO3/CuWO4 thin film. All


M

the thin films were annealed at 550 oC for 6 h.


D
TE
C EP
AC

Figure 4. XPS analysis for


the key chemical elements and
corresponding valence
states in
ACCEPTED MANUSCRIPT

nanostructured WO3 thin film, and nanocomposite γ-WO3/CuWO4 and γ-WO3/NiWO4 thin
films: (a) wide XPS scan on nanostructured WO3 thin film; (b) W 4f nanostructured WO3 thin
film; (c) (c) W 4f nanocomposite γ-WO3/NiWO4 thin film; (d) W 4f nanocomposite γ-
WO3/CuWO4 thin film; and (e) Ni 2p nanocomposite γ-WO3/NiWO4 thin film; (f) Cu 2p
nanocomposite γ-WO3/CuWO4 thin film. All the thin films were annealed at 550 oC for 6 h.

PT
Figure 4(a) shows the wide XPS scan for nanostructured WO3 thin film. From Figure 4(a), it can

RI
be observed that the basic chemical elements such as C, F, Si and Sn that are related to the FTO

showing peaks of 285.10, 689.60, 99.60 and 488.10 eV, respectively. Meanwhile, Figure 4(b),

SC
(c) and (d) show the high resolution XPS spectra of tungsten at different oxidation states.

U
Previously, the binding energies of W4f7/2 and W4f5/2 were widely investigated [18]. In this
AN
instance, tungsten elements with different binding energies and oxidation states actually

suggested the presence of both stoichiometric and sub-stoichiometric WO3. In our study, we
M

found that the W4f spectrum of WO3 was deconvoluted into two major peaks of W4f7/2 and W4f5/2

that are centered at the binding energies of 35.58 and 37.68 eV, respectively. According to
D

Saleem et al. [10], these are the typical binding energies that correspond to the W+6 oxidation
TE

state. Meanwhile, the other major peak located at 530.6 eV was related to O2-, and the peaks at

247.42 eV and 260 eV were corresponded to the W4d5 and W4d3 (see Supplementary
EP

Information).
C

From Figure 4(e), the high resolution Ni 2p spectrum shows the Ni2p3/3 peak at 856 eV that
AC

indicates the presence of Ni2+ ions. Further analysis on the Ni 2p spectrum evidenced that the

W4f7/2 peak was combined with the Ni peak at 856 eV in forming the γ-WO3/NiWO4

nanocomposite structure [19-20]. The corresponding binding energy for the Ni2p3/2 peak at 856.1

eV was in good agreement with the previously published data [20]. This indicates the presence of

NiO peak, whose binding energy of the Ni2p3/2 peak is at 854 eV. Both the O 1s spectra and peak
ACCEPTED MANUSCRIPT

fitting indicated the presence of γ-WO3 and NiWO4 compounds. The second peak at 530.6 eV

was corresponded to the binding energy of O2- in the polycrystalline NiWO4 structure [14]. From

Figure 4(f), the high resolution Cu spectrum was found the peak at 932.4 eV, corresponding the

Cu2+ in the nanocomposite γ-WO3/CuWO4 films. Similar to NiWO4, the peak at 530.6 eV was

PT
also found which indicating the existence of CuWO4 structure.

RI
SC
3.5. UV-Vis spectra

Figure 5 shows the UV-Vis absorption spectra of the nanostructured WO3 thin film, and

U
nanocomposite γ-WO3/CuWO4 and γ-WO3/NiWO4 thin films. From Figure 5, it can be observed
AN
that the nanostructured WO3 thin film had an obvious absorption edge at 450 nm. In this

instance, the estimated band gap for the nanostructured WO3 thin film was 2.70 eV from the
M

Kubelka-Monk function on the UV-Vis data, which is in accordance with previous published
D

studies [21-22]. Similarly, there were also small absorption edges from of 550 nm to 700 nm for
TE

nanostructured WO3 thin film that corresponded to the existence of a few W5+ species [14].
C EP
AC
ACCEPTED MANUSCRIPT

PT
(a) (b)

RI
Figure 5. (a) UV-vis absorption spectra for nanostructured WO3 thin film, and nanocomposite γ-

SC
WO3/CuWO4 and γ-WO3/NiWO4 thin films. (b) Tauc plot for the WO3 thin film, and
nanocompositeγ-WO3/CuWO4 and γ-WO3/NiWO4 thin films. All the thin films were annealed at
550 oC for 6 h.

U
AN
However, the UV-Vis spectrum for nanocomposite γ-WO3/CuWO4 thin film observed two

visible absorption edges between 300 nm and 550 nm, where an initial delay peak around 390-
M

420 nm and a shoulder at 450 nm were evidenced. These transitions were well assigned to the
D

W(5d)-O(2p) band and W(5d)-Cu(2p) band, respectively [6]. In this instance, the absorption tail
TE

was seen to be extended to a much longer wavelength of 650 nm, and this could be attributed to

the presence of defect levels within the band gap of nanocomposite γ-WO3/CuWO4 thin film.
EP

Additionally, the observed absorption tail characteristic for nanocomposite γ-WO3/CuWO4 thin

film at wavelengths longer than 650 nm could be due to the localized d-d transitions in Cu2+ ions.
C

The estimated band gap for nanocomposite γ-WO3/CuWO4 thin film was 2.30 eV. Previously, it
AC

was reported that the lower band gap for CuWO4 was due to the lowering of its valence band and

thus, resulted in photo-generated holes with sufficient overpotential for PEC water oxidation

reaction [6]. As for the nanocomposite γ-WO3/NiWO4 thin film, it exhibited three absorption

edges at 420 – 430 nm, 440-460 nm and 675 – 700 nm owing to the oxidation state of the

cations. In this instance, the first absorption edge may be attributed to the charge transfer
ACCEPTED MANUSCRIPT

transition in the WO6 matrix. Whilst the other two absorption edges were corroborated to the

presence of Ni2+ ions. The estimated band gap for the nanocomposite γ-WO3/NiWO4 was 2.10

eV.

PT
3.6. Photocurrent density measurements

RI
Figure 6 shows the photocurrent density profiles (versus applied potential), which are indirect

SC
photoactivity indicators for nanostructured WO3 thin film, and nanocomposite γ-WO3/CuWO4

and γ-WO3/NiWO4 thin films. From Figure 6, it can be seen that the photocurrent density

U
increased with increasing applied potential. Among the synthesized thin films, the bare
AN
nanostructured WO3 thin film measured the lowest photocurrent density of 100 µA/cm2 at 1.5 V

vs Ag/AgCl (4M KCl). This was caused by the surface agglomeration phenomenon observed in
M

the nanostructured WO3 thin film that leads to reduced surface contact area with the electrolyte
D

used. However, the photocurrent density of nanocomposite γ-WO3/CuWO4 was seen to improve
TE

after the incorporation of CuWO4 film where a two-fold enhancement to 250 µA/cm2 at 1.5 V vs

Ag/AgCl (4M KCl) was achievable as per Figure 6. This was predominantly due to that CuWO4
EP

was having a relatively lower band gap than WO3 and thus, contributed to the enhancement in its

photoactivity.
C
AC
ACCEPTED MANUSCRIPT

PT
RI
SC
Figure 6. Photocurrent density profiles for nanostructured WO3 thin film, and nanocomposite γ-
WO3/CuWO4 and γ-WO3/NiWO4 thin films in the solution of HCl containing SDS at pH 2. All

U
the thin films were annealed at 550 oC for 6 h.
AN
In this study, the nanocomposite γ-WO3/NiWO4 thin film resulted in the highest photocurrent
M

density of 400 µA/cm2 at 1.5 V vs Ag/AgCl (4M KCl) that was more than three-fold

enhancement than the bare nanostructured WO3 thin film. Through this study, it was shown that
D

NiWO4 is a good material candidate for the application in PEC water oxidation reaction. The
TE

heterojunction structure offers enhanced photoconversion efficiency and increased the density of
EP

carriers. The NiWO4/WO3 film has advantages of effective separation and restraining the

recombination of the photogenerated electron-hole pairs and better electron transport properties,
C

which resulted in enhancing PEC performance.


AC

The mechanism of these two heterojunctions (i.e. WO3/NiWO4 and WO3/CuWO4) for enhancing

the PEC performance can explained from the valence band (VB) and conduction band (CB)

position. Photogenerated electrons and holes will be formed by the surface absorption of

CuWO4/WO3 under the visible light irradiation. In the CuWO4/WO3 heterojunction films, the

generated electron of CuWO4 can move toward to the conduction band (CB) of WO3 due to the
ACCEPTED MANUSCRIPT

more negative CB edge potential compare with WO3. On the other hand, the valence band (VB)

edge potential of WO3 is lower than that of CuWO4, which lead to the generated holes in the VB

of WO3 flowing to that of CuWO4 and then reaction of oxygen evolution in the water. Through

the photogenerated carriers’ transfer at the interface of the heterojunction, the electron-hole

PT
separation will be accelerated and recombination of photo-generated carriers will be restrained,

RI
resulting in enhancing PEC performance. The mechanism for the NiWO4 is the same but with

more negative CB and positive VB comparing to CuWO4. Thus, the NiWO4/WO3 heterojunction

SC
showed higher photocurrent than CuWO4/WO3 composition.

U
AN
4. Conclusion

In this study, the monoclinic-I γ-WO3 crystal structure was electrochemically synthesized
M

followed by annealing treatment at 550 oC for 6 h. Due to the low photocurrent density of 100
D

µA/cm2 for nanostructured WO3 thin film, a modified synthesis approach was adopted where
TE

nanocomposite thin films of γ-WO3/CuWO4 and γ-WO3/NiWO4 aiming to improve the overall

photoactivity under visible light irradiation. Both the nanocomposite thin films of γ-
EP

WO3/CuWO4 and γ-WO3/NiWO4 exhibited more orderly, particle-like and porous structures that

can improve the surface contact with the electrolyte used in PEC water oxidation reaction. AFM
C

analysis revealed that the surface roughness was increased diminutively after the annealing
AC

treatment, indicating the superiority of electrodeposition synthesis followed by annealing-

induced crystallisation process. XRD results suggested that the WO3 films undergone phase

transformation process from amorphous to the photoactive monoclinic-I (γ-WO3) at 550 oC.

Similarly, XRD characteristics peaks for both the CuWO4 and NiWO4 were observed indicating

their successful incorporation into the thin film structures. This has been further confirmed by
ACCEPTED MANUSCRIPT

the XPS results that indicated the existences of WO42-, Ni2+ and Cu2+ ions at 35.58 eV, 856 eV

and 932.4 eV, respectively. Through the formation of WO3 nanocomposite structures, it can be

observed that the energy band gap was effectively lowered from 2.7 eV (γ-WO3)  2.3 eV (γ-

WO3/CuWO4)  2.1 eV (γ-WO3/NiWO4). From this study, it was found that the γ-WO3/NiWO4

PT
nanocomposite structure showed the highest photocurrent density of 400 µA/cm2 at 1.5 V vs

RI
Ag/AgCl (4M KCl).

SC
Acknowledgement

U
The authors are grateful to the financial support provided by the eScience fund (Project No: 03-
AN
02-10-SF0121) from Ministry of Science, Technology and Innovation (MOSTI), Malaysia.

Similar gratitude also goes to the Advanced Engineering Platform and School of Engineering of
M

Monash University Malaysia.


D
TE

References
EP

[1] Ahmad, H.; Kamarudin, S. K.; Minggu, L. J.; Kassim, M., Hydrogen from photo-catalytic

water splitting process: A review. Renewable and Sustainable Energy Reviews 2015, 43,
C

599-610.
AC

[2] Zhu, T.; Chong, M. N.; Chan, E. S., Nanostructured Tungsten Trioxide Thin Films

Synthesized for Photoelectrocatalytic Water Oxidation: A review. ChemSusChem 2014, 7

(11), 2974-2997.

[3] Zhu, T.; Chong, M. N.; Phuan, Y. W.; Chan, E.-S., Electrochemically synthesized tungsten

trioxide nanostructures for photoelectrochemical water splitting: Influence of heat treatment


ACCEPTED MANUSCRIPT

on physicochemical properties, photocurrent densities and electron shuttling. Colloids and

Surfaces A: Physicochemical and Engineering Aspects 2015, 484, 297-303.

[4] Niu, M.; Huang, F.; Cui, L.; Huang, P.; Yu, Y.; Wang, Y., Hydrothermal Synthesis,

Structural Characteristics, and Enhanced Photocatalysis of SnO2/α-Fe2O3 Semiconductor

PT
Nanoheterostructures. ACS Nano 2010, 4 (2), 681-688.

RI
[5] Amano, F.; Ishinaga, E.; Yamakata, A., Effect of Particle Size on the Photocatalytic Activity

of WO3 Particles for Water Oxidation. The Journal of Physical Chemistry C 2013, 117 (44),

SC
22584-22590.

[6] Yourey, J. E.; Kurtz, J. B.; Bartlett, B. M., Water Oxidation on a CuWO4–WO3 Composite

U
Electrode in the Presence of [Fe(CN)6]3–: Toward Solar Z-Scheme Water Splitting at Zero
AN
Bias. The Journal of Physical Chemistry C 2012, 116 (4), 3200-3205.

[7] Pyper, K. J.; Yourey, J. E.; Bartlett, B. M., Reactivity of CuWO4 in Photoelectrochemical
M

Water Oxidation Is Dictated by a Midgap Electronic State. The Journal of Physical


D

Chemistry C 2013, 117 (47), 24726-24732.


TE

[8] Yourey, J. E.; Pyper, K. J.; Kurtz, J. B.; Bartlett, B. M., Chemical Stability of CuWO4 for

Photoelectrochemical Water Oxidation. The Journal of Physical Chemistry C 2013, 117 (17),
EP

8708-8718.

[9] Pourmortazavi, S. M.; Rahimi-Nasrabadi, M.; Khalilian-Shalamzari, M.; Zahedi, M. M.;


C

Hajimirsadeghi, S. S.; Omrani, I., Synthesis, structure characterization and catalytic activity
AC

of nickel tungstate nanoparticles. Applied Surface Science 2012, 263 (Supplement C), 745-

752.
ACCEPTED MANUSCRIPT

[10] Saleem, M.; Al-Kuhaili, M. F.; Durrani, S. M. A.; Hendi, A. H. Y.; Bakhtiari, I. A.; Ali,

S., Influence of hydrogen annealing on the optoelectronic properties of WO3 thin films.

International Journal of Hydrogen Energy 2015, 40 (36), 12343-12351.

[11] Caramori, S.; Cristino, V.; Meda, L.; Tacca, A.; Argazzi, R.; Bignozzi, C. A., Efficient

PT
Anodically Grown WO3 for Photoelectrochemical Water Splitting. Energy Procedia 2012, 22

RI
(0), 127-136.

[12] Ng, C.; Ng, Y. H.; Iwase, A.; Amal, R., Influence of Annealing Temperature of WO3 in

SC
Photoelectrochemical Conversion and Energy Storage for Water Splitting. ACS Applied

Materials & Interfaces 2013, 5 (11), 5269-5275.

[13]
U
Pilli, S. K.; Deutsch, T. G.; Furtak, T. E.; Brown, L. D.; Turner, J. A.; Herring, A. M.,
AN
BiVO4/CuWO4 heterojunction photoanodes for efficient solar driven water oxidation.

Physical Chemistry Chemical Physics 2013, 15 (9), 3273-3278.


M

[14] Zhu, J.; Li, W.; Li, J.; Li, Y.; Hu, H.; Yang, Y., Photoelectrochemical activity of
D

NiWO4/WO3 heterojunction photoanode under visible light irradiation. Electrochimica Acta


TE

2013, 112 (Supplement C), 191-198.

[15] Gaillard, N.; Chang, Y.; DeAngelis, A.; Higgins, S.; Braun, A., A nanocomposite
EP

photoelectrode made of 2.2 eV band gap copper tungstate (CuWO4) and multi-wall carbon

nanotubes for solar-assisted water splitting. International Journal of Hydrogen Energy 2013,
C

38 (8), 3166-3176.
AC

[16] Yourey, J. E.; Bartlett, B. M., Electrochemical deposition and photoelectrochemistry of

CuWO4, a promising photoanode for water oxidation. Journal of Materials Chemistry 2011,

21 (21), 7651-7660.
ACCEPTED MANUSCRIPT

[17] Karthiga, R.; Kavitha, B.; Rajarajan, M.; Suganthi, A., Photocatalytic and antimicrobial

activity of NiWO4 nanoparticles stabilized by the plant extract. Materials Science in

Semiconductor Processing 2015, 40 (Supplement C), 123-129.

[18] Xu, X.; Gao, J.; Huang, G.; Qiu, H.; Wang, Z.; Wu, J.; Pan, Z.; Xing, F., Fabrication of

PT
CoWO4@NiWO4 nanocomposites with good supercapacitve performances. Electrochimica

RI
Acta 2015, 174 (Supplement C), 837-845.

[19] Bignozzi, C. A.; Caramori, S.; Cristino, V.; Argazzi, R.; Meda, L.; Tacca, A.,

SC
Nanostructured photoelectrodes based on WO3: applications to photooxidation of aqueous

electrolytes. Chemical Society Reviews 2013, 42 (6), 2228-2246.

[20]
U
Cimino, A.; Lo Jacono, M.; Schiavello, M., Structural, magnetic, and optical properties
AN
of nickel oxide supported on .eta.- and .gamma.-aluminas. The Journal of Physical Chemistry

1971, 75 (8), 1044-1050.


M

[21] Ma, S. S. K.; Maeda, K.; Abe, R.; Domen, K., Visible-light-driven nonsacrificial water
D

oxidation over tungsten trioxide powder modified with two different cocatalysts. Energy &
TE

Environmental Science 2012, 5 (8), 8390-8397.

[22] Ni, M.; Leung, M. K. H.; Leung, D. Y. C.; Sumathy, K., A review and recent
EP

developments in photocatalytic water-splitting using for hydrogen production. Renewable

and Sustainable Energy Reviews 2007, 11 (3), 401-425.


C
AC
ACCEPTED MANUSCRIPT

Highlights

• Tungstate metal nanocomposites were synthesized through facile electrodeposition.

• Resultant monoclinic-I WO3 resulted in the highest photoactivity.

PT
• Bandgap of WO3 was effectively lowered from 2.7 to 2.1 eV through γ-WO3/NiWO4.

• γ-WO3/NiWO4 resulted in the highest photocurrent density of 400 µA/cm2 at 1.5 V.

RI
U SC
AN
M
D
TE
C EP
AC

You might also like