You are on page 1of 10

Journal of Alloys and Compounds 947 (2023) 169480

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Boosting photoelectrochemical chlorine and hydrogen production with


oxygen vacancy rich TiO2 photoanodes ]]
]]]]]]
]]

Yuanchang Ye a, Bin Liao a, Miao Li a, Manfang Mai c, Li Zhang a, Lianke Ma a, Donghai Lin e,
⁎ ⁎⁎
Jishi Zhao d, , Dongchu Chen a,b, Xinzhou Ma a,b,
a
School of Materials Science and Hydrogen Energy, Foshan University, Foshan 528000, PR China
b
Guangdong Key Laboratory for Hydrogen Energy Technologies, Foshan 528000, PR China
c
School of Physics and Optoelectronic Engineering, Foshan University, Foshan 528000, PR China
d
Foshan Institute of Environment and Energy Technology, Foshan 528000, PR China
e
Shanghai Engineering Research Center of Advanced Thermal Functional Materials and School of Energy and Materials, Shanghai Polytechnic University, Shanghai
201209, PR China

a r t i cl e i nfo a bstr ac t

Article history: Photoelectrochemical chlorine (Cl2) evolution as the anodic reaction of solar hydrogen production is helpful
Received 5 December 2022 for increasing the additional value of hydrogen and reducing production cost. In this work, chemically stable
Received in revised form 16 February 2023 TiO2 nanotubes are used as the photoanodes for photoelectrochemical chlorine and hydrogen production.
Accepted 27 February 2023
Oxygen vacancies (VO) are introduced by post-annealing in H2/N2 atmosphere. By controlling the post-
Available online 28 February 2023
annealing temperature, concentration of oxygen vacancies is tuned efficiently. Systematic photoelec­
trochemical measurements reveal that the oxygen vacancies significantly enhance the PEC performance for
Keywords:
Solar hydrogen production Cl2 production with respect to the pristine TiO2. The sample post-annealed at 450 °C in H2/N2 achieves the
Oxygen vacancy highest Cl2 production rate of 37.12 μmol h−1 cm−2 with a Faradaic efficiency of 73.2% that resulted in fast
Chlorine evolution degradation of methyl orange at a speed of 5.73 mg h−1 cm−2. Furthermore, photocurrent of the two-
Methyl orange degradation electrode cell constructed with the VO-rich TiO2 is strongly increased by 250%, demonstrating that in­
Wastewater treatment troducing oxygen vacancies is a promising way for promoting the photoelectrochemical performance for Cl2
and H2 production. By careful measuring the Faradaic efficiency of various photoanodes, it is revealed that
the strongly boosted up Cl2 production is owing to enhancing utilization of photogenerated holes via the
oxygen vacancies.
© 2023 Elsevier B.V. All rights reserved.

1. Introduction evolution reaction (OER) at the (photo)anode, respectively. Photo­


electrochemical hydrogen conversion efficiency is mainly hindered
Converting the abundant but discontinuous sunlight to chemical by the kinetically slow OER [4]. On the other hand, producing low
fuels is considered as a promising way to deal with the issues of commercially valuable oxygen (O2) is not helpful for increasing the
global energy shortage and environmental pollution [1]. Hydrogen additional value of the photoelectrochemical produced hydrogen [5].
with high energy density is one of ideal solar fuel candidates [2]. Recently, some efforts have been conducted to break these limita­
Photoelectrochemical (PEC) water splitting as an attractive route for tions by substituting OER with other fast photoanodic reactions that
producing hydrogen has been widely studied to increase the con­ can produce high-value-added chemicals [6–8], such as H2O2 [9],
version efficiency and reduce the cost of hydrogen [3]. Generally, PEC S2O8- [10], Dihydroxyacetone [11], Chlorocyclohexane [12] and so on.
water splitting includes two separated half reactions of hydrogen Chlorine (Cl2) as an important chemical is widely used for was­
evolution reaction (HER) at the (photo)cathode and oxygen tewater treatment, polymer production and other industries, which
is more commercially valuable than O2 with a price of about $240
per ton [5]. 75 million tons Cl2 are produced annually via electrolysis
of NaCl that require huge amount of electric energy [13]. Therefore,

Corresponding author. producing H2 and Cl2 with abundant solar energy is very attractive.
⁎⁎
Corresponding author at: School of Materials Science and Hydrogen Energy,
Foshan University, Foshan 528000, PR China.
On the other hand, the 2-electron chlorine evolution reaction (CER)
E-mail addresses: zhaojishi0219@163.com (J. Zhao), is kinetically more favorable than the 4-electron OER [14]. Thus, CER
xinzhou.ma@hotmail.com (X. Ma).

https://doi.org/10.1016/j.jallcom.2023.169480
0925-8388/© 2023 Elsevier B.V. All rights reserved.
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

is a promising alternative counter reaction of the PEC hydrogen calcination temperature determined by studying the effect of calci­
evolution that would also be helpful for adding the extra value of nation temperature on PEC performance of the TiO2 photoanodes
solar hydrogen. with linear sweep voltammetry (Fig. S1). To obtain VO-rich TiO2
The standard potential of CER of 1.36 VNHE [15] requires the photoanodes, the pristine TiO2 photoanodes were then post-an­
photoanodes having very positive valance band position, which can nealed respectively in pure N2 and H2/N2 (10%/90%) atmosphere
be satisfied by some oxide semiconductors such as BiVO4 [16,17], (flowing 40 sccm H2 and 400 sccm N2) at different temperature of
WO3 [8,18,19] and TiO2 [20,21]. Since Zou’s work on using the RhO2 300, 350, 400, 450, 500, 550 °C for 2 h at a heating rate of 5 °C min−1.
modified Mo-doped BiVO4 photoanodes for producing Cl2 from Herein, the post-annealed TiO2 photoanodes are denoted as TiO2-N-
seawater [22], BiVO4-based photoanodes have attracted some at­ 450 (in pure N2 atmosphere), TiO2-H-300, TiO2-H-350, TiO2-H-400,
tentions for photoelectrochemical chloride oxidation [17,23,24]. TiO2-H-450, TiO2-H-500 and TiO2-H-550 (in H2/N2 atmosphere),
However, inhibiting photo-corrosion of the BiVO4 is still needed to respectively.
deal with for long-term application [25]. In comparison with BiVO4,
WO3 and TiO2 are relatively stable against photo-corrosion. Recently, 2.2. Characterization of TiO2 photoanodes
Augustrynski et. al. achieved efficient photoelectrochemical CER
with porous WO3 [8]. However, the high surface recombination of Morphology of the TiO2 photoanodes was imaged with a Hitachi
the photogenerated charge carriers at WO3 resulted in a very posi­ S4800 scanning electron microscope (SEM). Crystallization and
tive onset potential of photocurrent that seriously limited the energy phase composition of the TiO2 were characterized by a X′Pert PRO X-
efficiency [26]. Alternatively, TiO2 as a representative semiconductor ray diffractometer using Cu Kα (λ = 0.154 nm) radiation. X-ray pho­
is frequently used as photoanode, but most of efforts are con­ toelectron spectra of the TiO2 were recorded with a Thermo Fisher
centrated on PEC water oxidation [27,28]. Up to now, only a few Scientific ESCALAB 250 XI, irradiated with Al Kα (hν=1486.60 eV) at a
studies have focused on PEC chloride oxidation that requires further 500 µm diameter spot. The XPS results were calibrated with the
explored [19–21]. binding energy of C 1 s photoelectrons of 284.8 eV. Raman spectra of
Chemically and optically stable TiO2 nanotube arrays have very the TiO2 were recorded by a Horiba LabRam HR Evolution Raman
high specific surface area and open channel structure that are ben­ microscope with a 532 nm laser as the excitation light. The relative
eficial for decreasing transport length of charge carriers and in­ diffuse reflectance spectra of the TiO2 photoanodes in the range from
creasing light absorption efficiency [29,30]. In this work, TiO2 200 nm to 800 nm were measured with a Shimadzu UV-3600 Plus
nanotube arrays are used as the photoanodes to construct photo­ UV–VIS–NIR spectrophotometer.
electrochemical cells for chlorine and hydrogen production. Never­
theless, the wide band gap and high recombination of the charge 2.3. Photoelectrochemical characterizations
carriers still limit the PEC performance of the TiO2-based photo­
anodes. Various methods have been developed to improve the PEC Photoelectrochemical performance of the TiO2 photoanodes was
performance of the TiO2-based photoanodes [30,31]. For example, investigated with an electrochemical workstation of CH Instruments
introducing oxygen vacancies is one of efficient methods for im­ 660D. As sketched in Fig. 1, all the photoelectrochemical measure­
proving the PEC performance of the TiO2 [32–36] and other oxide ments were conducted with a two-compartment H-cell (Foshan Aoshi
photoanodes [37,38]. In this work, oxygen vacancy (VO) rich TiO2 Operanalysis, Fig. S2) with its two chambers separated by a cation
photoanodes were obtained by post-annealing of the nanotube ar­ exchange membrane (Fumasep FKS-50). The TiO2 photoanode with
rays in N2 and H2/N2 atmosphere at different temperatures. Sys­ exposed area of 0.45 cm2 as the working electrode and an Ag/AgCl
tematic photoelectrochemical measurements illustrated that all the (sat. KCl) as the reference electrode were sealed to the anodic
VO-rich TiO2 photoanodes had boosted up performance for Cl2 and chamber, where 60 mL 1 M NaCl electrolyte with the pH adjusted to 1
H2 production with respect to the pristine TiO2 photoanode. The by HCl was filled. A Pt wire counter electrode (ca. 2 cm2) was
highest Cl2 production rate of 37.12 μmol h−1 cm−2 with a Faradaic mounted to the cathodic chamber which was filled with 35 mL 0.5 M
efficiency of 73.2% was achieved by the VO-rich TiO2 post-annealed Na2SO4 with the pH adjusted to 1 using H2SO4. Before measurement,
at 450 °C in H2/N2. Potential application of the produced Cl2 for the Na2SO4 was bubbled with Ar (99.99%) for 30 min to eliminate the
wastewater treatment has also been demonstrated by the rapid dissolved oxygen and air inside the sealed cathodic chamber. A 300 W
degradation of methyl orange at a rate of 5.73 mg h−1 cm−2. By Xenon lamp (Perfect light PLS-SXE300D) was used as the light source
careful study on the Faradaic efficiency of Cl2 production of various for all the PEC measurements. Periodic illumination was generated by
TiO2, the dominate mechanism of the boosted performance for Cl2 an electronic shutter. The optical power of illumination was measured
production is attributed to the enhanced utilization of the photo­ with an optical power meter (FieldBest) to be 500 mW cm−2. For
generated charge carriers rather than upon improving selectivity of linear sweep voltammetry measurements, the potential of the TiO2
Cl- oxidation. This work would be very constructive for developing photoanodes was scanned positively from − 0.215–1.25 VAg/AgCl at a
VO-rich TiO2 photoanodes for high-performance solar Cl2 and H2 rate of 5 mV s−1 under periodic illumination. The Mott-Schottky plots
production as well as for degradation of organic pollution. under dark were obtained by sweeping the potential from
− 0.1–1.2 VAg/AgCl with a step of 0.01 V while superimposing a sinu­
2. Experimental soidal wave of 1 kHz, 10 mV. In the figures, the presented potential is
either referred to reversible hydrogen electrode (RHE) converted with
2.1. Preparation of VO-rich TiO2 photoanodes the relation of VRHE = (VAg/AgCl+0.197 +0.059 pH) V, or to the potential
of the standard chloride electrode (SClE) converted with the relation
Similar to the method reported by Yu [39], the TiO2 nanotube of VSClE= (VAg/AgCl-1.163) V.
photoanodes were fabricated with electrochemical anodization fol­ Photoelectrochemical Cl2 evolution performance of various TiO2
lowed by calcination. Firstly, amorphous TiO2 nanotubes were pre­ photoanodes at different potentials was measured by cycling the gas
pared by electrochemical anodization of a mechanically polished Ti inside the sealed anodic chamber to 200 mL indicative solution of
foil using a DC voltage of 30 V for 3 h in the mixture solution of 0.36 M Kl, 0.025 M NaOH and 0.049 M C8H5KO4, which would gra­
glycerol and H2O (50:50 vol%) containing 1 wt% NH4F electrolyte. dually turn yellow upon the formation of I3- by Cl2 (see Fig. S8), si­
After electrochemical anodization, the as-prepared amorphous TiO2 milar to the method described by Boyle [20]. The amount of Cl2 was
nanotubes were then calcined at 600 °C for 2 h in air at a heating rate quantitatively obtained by measuring the absorbance of the I3-
of 5 °C min−1 to improve their crystallization. 600 °C was the optimal peaking at 351 nm with a UV–vis spectrometer (UV-5200, Shanghai

2
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

Fig. 1. Schematic of the two-compartment H-cell for photoelectrochemical measurements of the TiO2 photoanodes for Cl2 and H2 production and methyl orange degradation.

Metash). The standard curve between absorbance and the amount of confirm that the TiO2 photoanodes contained the mixed phases of
Cl2 is shown in Fig. S3. To further demonstrate the application for anatase and rutile and the post-annealing treatment did not change
degradation of organic pollution of methyl orange (MO), the pro­ the crystalline and phase composition of the TiO2 photoanodes.
ducing Cl2 was cycling into 250 mL H2O containing 2.5 mg MO The light harvest properties of the TiO2 photoanodes were eval­
(10 mg L−1). The degradation rate of MO was obtained by measuring uated by the diffuse reflectance UV–vis spectra in Fig. 2e. With re­
the absorbance of the sampling solution each 0.5 h. The above used spect to the pristine TiO2, the characteristic absorption edge of the
three-electrode PEC cell can only study the PEC performance of the post-annealed TiO2 photoanodes just slightly red-shifted to around
TiO2 photoanode individually. Two-electrode configured PEC cells 390 nm and the adsorption intensity slightly increased as well. Ac­
are more practical for actual application. Therefore, the PEC perfor­ cording to the Kubelka-Munk function:
mance of the two-electrode cells constructed with the TiO2 photo­
anodes and a Pt wire dark cathode at different cell voltages was also ( h )n (hv Eg)
investigated. The produced H2 was measured with the thermal
conductivity detector of a gas chromatography (GC-7900, Zhongjiao where α is absorption efficiency, hν is energy of light, Eg is the band
Jinyuan). gap of the semiconductor, n is 0.5 for indirect bandgap semiconductor
and is 2 for direct bandgap semiconductor, the band gap of the TiO2
photoanodes can be determined [47]. Fig. 2f shows the Tauc plots
3. Results and discussion derived from the diffuse reflectance UV–vis spectra via the Kubelka-
Munk function. The Eg of the TiO2 photoanodes was obtained by the
3.1. Morphology, structure, optical properties of the TiO2 photoanodes intercept the linear portion of the Tauc plots with the hv-axis, as
shown in Fig. S6. It can be seen from the inset in Fig. 2f that the Eg of
Fig. 2a shows the SEM image of the top-view morphology of the the pristine TiO2 is 3.13 eV and the Eg of the VO-rich TiO2 was just
pristine TiO2 nanotube array. The TiO2 nanotubes are highly ordered. slightly reduced with the biggest variation of 0.07 eV. The VO-rich TiO2
The average pore diameter of the nanotube is around 180–200 nm photoanodes show considerably enhanced visible light absorption
and the thickness of the wall is about 20 nm. As revealed by Fig. S4, capability upon the light trapping effect of the nanotube structure and
the post-annealing treatment did not change the morphology of the the transitions between the VO level and the energy bands of the TiO2
TiO2 photoanodes. The crystalline and phase composition of the [48]. With respect to the pristine TiO2, light harvesting capability of
pristine TiO2 and the TiO2-N-450 and TiO2-H-450 were character­ the VO-rich TiO2 was slightly improved. However, it needs to point out
ized with X-ray diffraction that are shown in Fig. 2b (XRD patterns of that the small enhancement on light harvesting should not be the
other photoanodes can be seen in Fig. S5). In addition to the in­ dominated reason for the strongly boosted up PEC performance
tensive diffraction peaks from the Ti substrate, diffraction peaks of achieved by the VO-rich TiO2 photoanodes.
the anatase phase and rutile phase are also observed, demonstrating Impact of the post-annealing temperature and atmosphere on
that the TiO2 nanotubes mixed with these two phases [40,41]. It is the surface chemical composition and oxygen vacancy content was
believed that the heterojunctions between the two phases are explored with high resolution XPS spectra of Ti 2p and O 1 s Fig. 3a
helpful for minimizing the recombination rate of the photo­ shows the Ti 2p XPS spectra of the pristine TiO2 and VO-rich TiO2
generated charge carriers and the ratio of the two phases on PEC photoanodes. There are four binding energy (BE) peaks of Ti 2p
performance has been comprehensively studied [42–44]. The posi­ observed for all the photoanodes. The BE peaks of 458.7 eV and
tion and width of the diffraction peaks are almost the same, in­ 464.4 eV of the pristine TiO2 are corresponding to the Ti 2p3/2 and Ti
dicating that the post-annealing treatment did not significantly 2p1/2 of Ti4+, respectively [33]. Separation of the two BE peaks is
change the crystalline and phase composition of the TiO2. 5.7 eV, which is typical for the Ti4+-O bonds in TiO2 [48]. There are
Fig. 2c and d show the Raman spectra of the TiO2 photoanodes. two small peaks centered at low binding energy of 458.4 eV and
The most intense peak at 143 cm−1 and the less intense peak at 463.6 eV are assigned to 2p3/2 and 2p1/2 of Ti3+, demonstrating the
196 cm−1 can be attributed to the Eg modes of the anatase phase [45]. existence of Ti3+ on the surface of the pristine TiO2 [48]. The surface
The peaks at 395 cm−1, 514 cm−1 and 636 cm−1 correspond to the B1 g, VO content of the pristine TiO2 is calculated to be 6.34% with the XPS
A1 g+B1 g and Eg modes of the anatase phase, respectively [46]. The peak areas of Ti 2p (see Table S1) [48]. The BE peak positions of Ti 2p
presence of the rutile phase is confirmed by its Eg mode at 445 cm−1 of the TiO2-N-450 did not shift with respect to the pristine TiO2,
and the A1 g mode at 612 cm−1 [45]. XRD results and Raman spectra while the surface VO content increased to 8.10%. Post-annealing with

3
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

Fig. 2. (a) SEM image of the pristine TiO2 nanotubes; (b) XRD patterns of the pristine TiO2, the TiO2-N-450 and the TiO2-H-450 photoanodes; (c) Raman spectra and (d) Zoom-in
Raman spectra of the TiO2 photoanodes; (e) diffuse reflectance UV–vis spectra of the TiO2 photoanodes; (f) the derived Tauc Plots and the Eg of the TiO2 photoanodes.

the present of 10% H2 increased the VO content to 9.72% at 450 °C. potential region (< 0.3 VRHE) was produced, the obvious cathodic
Surprisingly, the surface VO content was not further increased by spikes soon after illumination off indicate serious surface combina­
post-annealing at 500 °C and 550 °C. It might because that lattice tion loss, since the measured cathodic spikes related to the electron
oxygen in bulk had sufficient energy to diffuse to surface and re­ flux moving to combine with the holes on surface. Photocurrent of the
sulted in the formation of oxygen vacancies in depth that could not pristine TiO2 rapidly increased with the positively increasing potential
be analyzed with XPS. This suggestion is supported by the Mott- to a plateau of about 1.88 mA cm−2. In comparison with the pristine
Schottky results in Fig. 4(b), which demonstrate that the donor TiO2, all the VO-rich TiO2 photoanodes generated significantly boosted
concentration related to oxygen vacancy content in bulk indeed in­ up photocurrent, demonstrating the strong enhance effect of the
creased with the increasing post-annealing temperature. oxygen vacancies on PEC performance. The photocurrent of the TiO2-
Since the missed oxygen atoms do not generate any X-ray pho­ N-450 reached to 2.2 mA cm−2 at 1.23 VRHE, enhanced by a factor of
toelectron, oxygen vacancies cannot be directly measured with XPS 1.17. The TiO2-H-300, TiO2-H-350 and TiO2-H-400 had comparable
of oxygen. However, upon exposure to ambient for certain time, PEC performance and their photocurrent was increased by a factor of
surface hydroxyls (absorbed -OH group) would be formed by dis­ 1.42 to around 2.68 mA cm−2 at 1.23 VRHE. The TiO2-H-450 photoanode
sociation of water at oxygen vacancies [49]. Therefore, the existence achieved the highest photocurrent of 3.21 mA cm−2 with an enhance
of the oxygen vacancies were evidenced by the XPS signal of the factor of 1.71. After that, the PEC performance was not further im­
absorbed -OH group [32]. Fig. 3b shows the high resolution XPS proved with the increasing post-annealing temperature. The photo­
spectra of O 1 s that were deconvoluted into three peaks of 529.8 eV, current of the TiO2-H-500 and TiO2-H-550 was slightly lower with
330.7 eV and 532.5 eV, corresponding to the lattice oxygen (OL), the respect to that of the TiO2-H-450.
surface absorbed oxygen (OA) and the oxygen of the absorbed -OH To elucidate the enhance effect of the oxygen vacancies on the
group (OOH) [50], respectively. Obviously, the surface of all the TiO2 photocurrent, interfacial properties of the TiO2 photoanodes were
photoanodes contained Ti3+ and VO. With respect to the pristine studied with Mott-Schottky plots, as shown in Fig. 4b. Obviously, the
TiO2, the content of Ti3+ and VO was increased by post-annealing slope of the linear portion of the Mott-Schottky plots is strongly
treatment. Generally, the BE peak positions of Ti 2p and O 1 s were dependent on the post-annealing conditions. Concentration of donor
just slightly shifted. The largest shifts were observed for the TiO2-H- ND and the flat band potential φFB of the TiO2 photoanodes can be
350 that the Ti4+ 2p1/2 shifted positively by 0.2 eV and the Ti4+ 2p3/2 determined from the Mott-Schottky plots [51] with Eq. (1):
shifted positively by 0.1 eV. Meanwhile the O 1 s peak of OL shifted
1 2 kT
positively by 0.15 eV. The positively shifted BE peak positions in­
2
= FB
dicate the shorten of the Ti4+-O bond length. CSC e o r ND e (1)

in which CSC is capacitance of semiconductor electrode, φ is the


3.2. Photoelectrochemical performance of the TiO2 photoanodes applied electrode potential, φFB is flat band potential, k is Boltzmann
constant; T is absolute temperature, e is electron charge, εo is di­
Firstly, the photoelectrochemical performance of the pristine TiO2 electric constant in vacuum, εr is relative dielectric constant which is
and the VO-rich TiO2 photoanodes in 1 M NaCl (pH=1) was studied 55 for TiO2 [52] and ND is donor concentration. The obtained ND of
with linear sweep voltammetry (LSV). Fig. 4a shows the LSV curves of the photoanodes is presented on Fig. 4b. ND of the pristine TiO2 is
the TiO2 photoanodes under periodic light illumination. Considerable 6.24 × 1017 cm−3. ND of the TiO2-H-300, TiO2-H-350 and TiO2-H-400
photocurrent was produced even when the potential was as low as is around 2.20 × 1018 cm−3 that is consistent with their comparable
0.08 VRHE for all the photoanodes, pointing to the low onset potential photocurrent. ND of the TiO2-H-450 rapidly increased to 8.80 × 1018
of PEC reactions. Though relatively high photocurrent at the low cm−3, 3.24 times of that of the TiO2-N-450, revealing that oxygen

4
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

Fig. 3. XPS spectra of the pristine TiO2, the TiO2-N-450, TiO2-H-300, TiO2-H-350, TiO2-H-400, TiO2-H-450, TiO2-H-500 and TiO2-H-550, (a) Ti 2p and (b) O 1s.

vacancies were introduced more easily by the present of hydrogen in function of the applied electrode potential that was calculated with
the post-annealing atmosphere. The TiO2-H-550 has the highest ND the CH of 200 μF cm−2 deduced from the Mott-Schottky plots in the
of 8.58 × 1020 cm−3 which is 3 orders of magnitude higher with re­ section where the interfacial capacitance was dominated by the CH.
spect to the pristine TiO2. Apparently, the Φsc of the pristine TiO2 was largest. However, this
As indicated by Fig. 4b, not only the ND was strongly increased by result seems contradictory with the photocurrent in Fig. 4a.
post-annealing, the φFB of the VO-rich TiO2 shifted positively to Separation efficiency of the photo-generated charge carriers is
around 0.4 VRHE with respect to the φFB of the pristine TiO2 of not only determined by the Φsc but also by the thickness and the
0.136 VRHE. In fact, the positively shifted φFB is not helpful for en­ electric field of the depletion layer. According to Eq. (4),
hancing photocurrent since the potential drop across the depletion
layer would be decreased. The band bending Φsc can be predicted by 20 r kT
w= sc
Eqs. (2) and (3), eND e (4)
sc = FB H (2)
the depletion layer thicknesses w of the TiO2 photoanodes were
CH calculated [53] and shown in Fig. 4d. Although the Φsc of the pristine
sc
= TiO2 was highest, the lowest ND of 6.24 × 1017 cm−3 made the
H Csc (3)
thickness of depletion layer largest that resulted in the weakest
where ΦH is the potential drop across the Helmholtz layer and CH is strength of the electric field (see Fig. S7). This is the main reason for
the capacitance of the Helmholtz layer [51]. Fig. 4c shows the Φsc as a the lowest photocurrent generated by pristine TiO2 photoanode.

5
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

Fig. 4. (a) LSV curves of various TiO2 photoanodes under periodic illumination; (b) Mott-Schottky plots of the TiO2 photoanodes measured under dark; (c) the calculated potential
drop across the depletion layer as a function of the applied potential; (d) the calculated thickness of the depletion layer of the TiO2 photoanodes as a function of the applied
potential.

As shown in Fig. 4d, the depletion layer thicknesses decreased distributed across the Helmholtz layer in this potential region that
with the increasing post-annealing temperature. The depletion layer only resulted in a small upward band bending across the accumu­
thicknesses of the TiO2-H-300, TiO2-H-350 and TiO2-H-400 were lation layer. The very positive valance band position of the TiO2 still
around 40 nm, twice of the nanotube wall thickness. The higher would allow portion of the photo-generated holes transport to sur­
electric field upon the thinner depletion layer enhanced the charge face for taking part in anodic reactions.
separation efficiency and can be attributed to the increased photo­ Fig. 5a-f show the j-t curves of the TiO2 photoanodes at different
current with respect to the pristine TiO2. The highest photocurrent potentials under periodic illumination. Herein, the presented po­
achieved by the TiO2-H-450 can also be explained by this point of tential is referred to the standard potential of chloride electrode
view. The depletion layer thickness of the TiO2-H-450 was around (VSClE). The j-t results are consistent with the LSV results that sig­
20 nm that was close to the wall thickness of the TiO2 nanotubes, nificantly enhanced photocurrent was obtained by the VO-rich TiO2
indicating that the wall of the TiO2-H-450 was completely depleted photoanodes. As shown in Fig. 5a, even at − 1.0 VSClE, photocurrent of
with the highest electric field horizontally for efficient charge carrier 1.5 mA cm−2 was produced by the TiO2-H-450 photoanode, in­
separation. On the other hand, the existed oxygen vacancies in the dicating a large photovoltage was provided for PEC Cl2 evolution,
depletion layer could extract the holes for prolonging their lifetime benefit from the very positive valance band position of the TiO2. As
[34]. Long lifetime of the photogenerated holes inside the com­ expected, the photocurrent of all the photoanodes was increased
pletely depleted nanotubes were very helpful for electrochemical with the increasing potential. At − 0.3 VSClE, the highest photocurrent
reaction in timescale of ms to s [54]. Though the higher ND made the of 3.0 mA cm−2 was generated by the TiO2-H-450, which was in­
TiO2-H-500 and the TiO2-H-550 obtained stronger electric field upon creased by a factor of 1.71 with respect to the pristine TiO2.
the much thinner depletion layer, the wall of the nanotubes could The produced Cl2 of the TiO2 photoanodes at different applied
not be completely depleted in these cases, and the charge carriers potentials was measured by cycling the gas inside the anodic
generated in the middle of the wall needed to diffuse to the deple­ chamber into the indicative solution of KI which was gradually
tion layer for separation. Therefore, photocurrent of the TiO2-H-500 turning to yellow (see Fig. S8). The amount of Cl2 was calculated
and TiO2-H-550 did not further increase with the increasing ND. with the absorbance peak at 351 nm. Fig. 5i shows the corresponding
It is interesting to observed that considerable photocurrent was produced amount and Faradaic efficiency (FE) of Cl2 in 2 h with ef­
generated in the potential region more negative than the φFB in fective illumination time of 1 h. The photocurrent in full timescale of
Fig. 4a, especially for the VO-rich TiO2 photoanodes. The explanation 2 h can be found in Fig. S9. At − 0.4 VSClE, the TiO2-H-450 produced
for this observation is because the applied potential was mainly the highest amount of Cl2 of 37.12 μmol cm−2 with a FECl2 of 73.92%,

6
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

Fig. 5. Photocurrent of the TiO2 photoanodes at different potentials of (a) − 1.0 VSClE, (b) − 0.9 VSClE, (c) − 0.8 VSClE, (d) − 0.7 VSClE, (e) − 0.6 VSClE, (f) − 0.5 VSClE, (g) − 0.4 VSClE, (h)
− 0.3 VSClE; (i) the corresponding amount and Faradaic efficiency of Cl2.

Fig. 6. (a) Long-term photocurrent of the TiO2-H-450 at −0.4 VSClE; (b) degradation of methyl orange, production of H2 and the corresponding amount of the passed charge.

which was 1.56 times of that of the pristine TiO2 of 23.76 μmol cm−2. would be understood by the self-healing of the surface oxygen va­
However, though the Cl2 production was significantly enhanced by cancies in aqueous solutions [49]. On the other hand, the FECl2
the VO-rich TiO2, as indicated by the FECl2, the tendency of FECl2 steeply decreased with the increasing potential. For example, FECl2 of
generally decreased with the increasing post-annealing temperature the TiO2-H-450 decreased from 85.6% to 52.9% as the potential in­
and was not consistent with the tendency of surface vacancy content creased from − 1.0 VSClE to − 0.3 VSClE, revealing that the kinetically
(listed in Table S1). Therefore, this reveals that the selectivity of Cl- slow oxygen evolution was more favorable at relatively positive
oxidation was not improved by the surface oxygen vacancies, the potential. But in general, the FECl2 of the VO-rich TiO2 is higher than
enhanced Cl2 production of the VO-rich TiO2 was mainly because of that reported FECl2 [8].
the more efficient utilization of the photogenerated charge carriers. Fig. 6a shows long-term photocurrent of the TiO2-H-450 at
The non-catalysis effect of the surface vacancies on Cl- oxidation − 0.4 VSClE. The photocurrent steeply decreased about 15% in the first

7
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

Fig. 7. (a) Scheme of the two-electrode cell for PEC production of H2 and Cl2; (b) LSV curves of the two-electrode cells constructed with the pristine TiO2 and the TiO2-H-450 as
the photoanode, respectively; (c) long-term photocurrent of the two-electrode cell with the TiO2-H-450 under different cell voltages; (d) the corresponding production rate of H2,
Cl2 and its Faradaic efficiency.

3 h and then trended to stable at around 2.5 mA cm−2. One possible can be used for long-term Cl2 and H2 production as well as organic
reason for the decreased photocurrent would be the dilution effect pollution degradation.
of Cl- inside the nanotube. Blockage effect of the generated gas
bubbles attached on TiO2 surface was another possible reason for the 3.3. PEC performance of the two-electrode cells for H2 and Cl2
decreased photocurrent. As shown by Fig. 6b, the amount of passed production
charge almost linearly increased with time, demonstrating the high
stability of the TiO2-H-450. In order to demonstrate the application In above studies, the PEC performance of the TiO2 photoanodes
of the produced Cl2 for wastewater treatment, e.g., degradation of was individually studied in the three-electrode configured cell. The
organic pollution, the gas inside the anodic chamber was cycling into effect of the dark cathode on the overall PEC performance could not
250 mL H2O containing 2.5 mg simulated organic dye of methyl or­ be figured out with the three-electrode system [59]. Furthermore,
ange (MO). The degradation of the MO was measured each 0.5 h with the three-electrode PEC cell is relatively complicate for practical
a UV–vis spectrometer. The corresponding absorbance spectra are application. In contrast, two-electrode configured PEC cells are more
shown in Fig. S10. As shown by the point-line curves in Fig. 6b, practical for actual application. Therefore, in this work, two-elec­
2.5 mg MO was almost completely degraded in 2 h. After that, an­ trode PEC cells were also constructed with a Pt wire as the dark
other 2.5 mg MO was injected into the 250 mL H2O for the second- cathode and the pristine TiO2 and the VO-rich TiO2 as the photo­
round measurement. Totally 12.5 mg MO was injected in 10 h. The anode, respectively, as sketched in Fig. 7a. Effect of the cell voltage
degradation rate of the MO by the PEC producing Cl2 was about on the performance of H2 and Cl2 production was studied. Ther­
5.73 mg h−1 cm−2 (calculated with the effective illumination time), modynamically, the equilibrium cell voltage for Cl2 and H2 evolution
which is comparable with the reported results [55–57]. Meanwhile is 1.419 V, which means that H2 and Cl2 would not be produced
the produced H2 at the Pt dark electrode was collected in the sealed under dark if the cell voltage was less than 1.419 V.
cathodic chamber and the evolution rate was determined to be 53.64 Under illumination electrons and holes were generated that could
μmol h−1 cm−2, which is much higher than the production rate of 2.1 make the quasi-Fermi level of the holes much more positive than the
μmol h−1 cm−2 obtained with the AgIn5Se8 sensitized TiO2 nanotubes equilibrium potential of chlorine evolution, and PEC production of H2
[58]. The Faradaic efficiency of the H2 was close to 100%. The above and Cl2 could be driven with a cell voltage less than 1.419 V.
studies demonstrate that the highly stable VO-rich TiO2 photoanodes Considerable photocurrent was observed in the LSV curves in Fig. 7b

8
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

even when the cell voltage was as low as 0 V. Photocurrent of the two- Acknowledgments
electrode cell using the TiO2-H-450 as the photoanode generated
3.5 mA cm−2 at 1.419 V, enhanced by 250% with respect to its coun­ This work was financially supported the National Natural Science
terpart of the pristine TiO2, indicating that the overall performance of Foundation of China (No. 11504242); the Project of Guangdong
the cell was dominated by the photoanode. As mentioned above, the Provincial Education (No. 2020KTSCX131); Guangdong Basic and
strongly boosted up photocurrent was resulted from the introduced Applied Basic Research Foundation (Nos. 2020B1515120006 and
oxygen vacancies in the TiO2-H-450, which increased the strength of 2020B1515120097); the Innovation Team of Universities of
the electric field in the depletion layer and acted as the trapping Guangdong Province (Nos. 2020KCXTD011 and 2022KCXTD030); the
centers for the photogenerated holes to prolong their lifetime for Guangdong Key Laboratory for Hydrogen Energy Technologies (No.
promoting oxidation of Cl-, as sketched in Fig. 7a. 2018B030322005).
Fig. 7b shows the long-term photocurrent of the PEC cell with the
TiO2-H-450 as the photoanode under various cell voltages and the Appendix A. Supplementary material
corresponding production rates of H2 and Cl2 were shown in Fig. 7c.
The photocurrent was almost constant at low cell voltage of 0.619 V, Supplementary data associated with this article can be found in
0.719 V and 0.819 V. Steep decay of the photocurrent was observed the online version at doi:10.1016/j.jallcom.2023.169480.
when the cell voltage was higher than 0.919 V. As expected, the
evolution rate of H2 increased from 26.2 μmol h−1 cm−2 to 52.0 μmol References
h−1 cm−2 with the increasing cell voltage. The FE of Cl2 varied be­
tween 52% and 65% with the cell voltage. The highest FE of Cl2 of 65% [1] Y. Liu, F. Wang, Z. Jiao, S. Bai, H. Qiu, L. Guo, Photochemical systems for solar-to-
was obtained at 1.119 V that resulted in the highest Cl2 production fuel production, Electrochem. Energy Rev. 5 (2022) 5.
[2] N. Pirrone, F. Bella, S. Hernández, Solar H2 production systems: current status
rate of 28.7 μmol h−1 cm−2. and prospective applications, Green Chem. 24 (2022) 5379–5402.
[3] A. Thakur, D. Ghosh, P. Devi, K.H. Kim, P. Kumar, Current progress and challenges
in photoelectrode materials for the production of hydrogen, Chem. Eng. J. 397
4. Conclusions (2020) 125415.
[4] J. Zhang, J. Cui, S. Eslava, Oxygen evolution catalysts at transition metal oxide
In this work, VO-rich TiO2 photoanodes fabricated by post-an­ photoanodes: their differing roles for solar water splitting, Adv. Energy Mater. 11
(2021) 2003111.
nealing treatment in H2/N2 atmosphere at different temperatures
[5] K. Sayama, Production of high-value-added chemicals on oxide semiconductor
were used as photoanodes for photoelectrochemical chlorine and photoanodes under visible light for solar chemical-conversion processes, ACS
hydrogen production. With respect to the pristine TiO2 photoanode, Energy Lett. 3 (2018) 1093–1101.
all the VO-rich TiO2 gained boosted up performance for H2 and Cl2 [6] K. Fuku, N. Wang, Y. Miseki, T. Funaki, K. Sayama, Photoelectrochemical reaction
for the efficient production of hydrogen and high-value-added oxidation re­
production. Among them, the TiO2-H-450 photoanode exhibited the agents, ChemSusChem 8 (2015) 1593–1600.
best performance, which achieved the highest Cl2 production of [7] M. Natali, A. Sartorel, A. Ruggi, Beyond water oxidation: hybrid, molecular-based
37.12 μmol cm−2 h−1 with an FE of 73.9% at − 0.4 VSClE. By system­ photoanodes for the production of value-added organics, Front. Chem. 10 (2022)
907510.
atically studied the FE of Cl2 of the various TiO2 photoanodes, it is [8] M. Jadwiszczak, K. Jakubow-Piotrowska, P. Kedzierzawski, K. Bienkowski,
revealed that the increased Cl2 production was not due to increasing J. Augustynski, Highly efficient sunlight-driven seawater splitting in a photo­
selectivity of Cl- oxidation but upon the enhanced utilization of electrochemical cell with chlorine evolved at nanostructured WO3 photoanode
and hydrogen stored as hydride within metallic cathode, Adv. Energy Mater. 10
photogenerated holes by the oxygen vacancies that minimized the (2020) 1903213.
recombination loss. As demonstrated by the rapid degradation of [9] K. Zhang, J. Liu, L. Wang, B. Jin, X. Yang, S. Zhang, J.H. Park, Near-complete
methyl orange at a rate of 5.73 mg h−1 cm−2, the PEC production of suppression of oxygen evolution for photoelectrochemical H2O oxidative H2O2
synthesis, J. Am. Chem. Soc. 142 (2020) 8641–8648.
Cl2 showed a great potential application for organic dye degradation [10] W. Xie, Z. Huang, R. Wang, C. Wen, Y. Zhou, Metallic Pt and PtOx dual-cocatalyst-
and wastewater treatment. Finally, the two-electrode PEC cell con­ loaded WO3 for photocatalytic production of peroxydisulfate and hydrogen
structed with TiO2-H-450 achieved H2 production rate of 49.2 μmol peroxide, J. Mater. Sci. 55 (2020) 11829–11840.
[11] H. Tateno, S.-Y. Chen, Y. Miseki, T. Nakajima, T. Mochizuki, K. Sayama,
cm−2 h−1 and the highest Cl2 production rate of 28.7 μmol cm−2 h−1
Photoelectrochemical oxidation of glycerol to dihydroxyacetone over an acid-
with the highest EF of 65% at the cell voltage of 1.119 V. This work is resistant Ta:BiVO4 photoanode, ACS Sustain. Chem. Eng. 10 (2022) 7586–7594.
constructive for guiding the fabrication of the high-performance VO- [12] Z. Li, L. Luo, M. Li, W. Chen, Y. Liu, J. Yang, S.M. Xu, H. Zhou, L. Ma, M. Xu, X. Kong,
rich TiO2 photoanodes and provides a deep insight into the me­ H. Duan, Photoelectrocatalytic C-H halogenation over an oxygen vacancy-rich
TiO2 photoanode, Nat. Commun. 12 (2021) 6698.
chanism of oxygen vacancy enhanced PEC Cl2 and H2 production. [13] H. Wang, Z. Han, Y. Zhou, X. Liu, D. Zeng, W. Wang, D. Sarker, L. Zhang, W. Wang,
Efficient photocatalytic chlorine production on bismuth oxychloride in chloride
solution, Appl. Catal. B 297 (2021) 120436.
CRediT authorship contribution statement [14] R.K.B. Karlsson, A. Cornell, Selectivity between oxygen and chlorine evolution in
the chlor-alkali and chlorate processes, Chem. Rev. 116 (2016) 2982–3028.
Yuanchang Ye: Methodology, Investigation, Conceptualization. [15] K.S. Exner, Controlling stability and selectivity in the competing chlorine and
oxygen evolution reaction over transition metal oxide electrodes,
Bin Liao: Methodology, Investigation. Miao ChemElectroChem 6 (2019) 3401–3409.
Li: Investigation. Manfang Mai: Resources. Li Zhang: Data [16] L. Ma, T. Chen, Q. Li, M. Mai, X. Ye, J. Mai, C. Liu, J. Zhang, D. Lin, X. Ma, Yb3+, Er3+
curation. Lianke Ma: Resources. Donghai Lin: Data curation. Jishi co-doped NaGdF4/BiVO4 embedded Cu2O photocathodes for photoelec­
trochemical water reduction with near infrared light, Appl. Surf. Sci. 585 (2022)
Zhao: Writing – original draft. Dongchu Chen: Writing – original 152650.
draft. Xinzhou Ma: Conceptualization, Methodology, Supervision. [17] A.M. Rassoolkhani, W. Cheng, J. Lee, A. McKee, J. Koonce, J. Coffel, A.H. Ghanim,
G.A. Aurand, C. Soo Kim, W. Ik Park, H. Jung, S. Mubeen, Nanostructured bismuth
vanadate/tungsten oxide photoanode for chlorine production with hydrogen
Data availability generation at the dark cathode, Commun. Chem. 2 (2019) 57.
[18] R. Pang, Y. Miseki, S. Okunaka, K. Sayama, Photocatalytic production of hypo­
chlorous acid over Pt/WO3 under simulated solar light, ACS Sustain. Chem. Eng. 8
No data was used for the research described in the article.
(2020) 8629–8637.
[19] X. Huang, Y. Zhang, J. Bai, J. Li, L. Li, T. Zhou, S. Chen, J. Wang, M. Rahim, X. Guan,
Declaration of Competing Interest B. Zhou, Efficient degradation of N-containing organic wastewater via chlorine
oxide radical generated by a photoelectrochemical system, Chem. Eng. J. 392
(2020) 123695.
The authors declare that they have no known competing fi­ [20] C. Boyle, N. Skillen, H.Q.N. Gunaratne, P.K. Sharma, J.A. Byrne, P.K.J. Robertson,
nancial interests or personal relationships that could have appeared The use of titanium dioxide nanotubes as photoanodes for chloride oxidation,
Mater. Sci. Semicond. Process. 109 (2020) 104930.
to influence the work reported in this paper.

9
Y. Ye, B. Liao, M. Li et al. Journal of Alloys and Compounds 947 (2023) 169480

[21] G. Chehade, N. Alrawahi, B. Yuzer, I. Dincer, A photoelectrochemical system for [40] K.M. Chahrour, F.K. Yam, A.M. Eid, A.A. Nazeer, Enhanced photoelectrochemical
hydrogen and chlorine production from industrial waste acids, Sci. Total Environ. properties of hierarchical black TiO2−x nanolaces for Cr (VI) photocatalytic re­
712 (2020) 136358. duction, Int. J. Hydrog. Energy 45 (2020) 22674–22690.
[22] W. Luo, Z. Yang, Z. Li, J. Zhang, J. Liu, Z. Zhao, Z. Wang, S. Yan, T. Yu, Z. Zou, Solar [41] J.M. Macak, H. Tsuchiya, A. Ghicov, K. Yasuda, R. Hahn, S. Bauer, P. Schmuki, TiO2
hydrogen generation from seawater with a modified BiVO4 photoanode, Energy nanotubes: self-organized electrochemical formation, properties and applica­
Environ. Sci. 4 (2011) 4046–4051. tions, Curr. Opin. Solid State Mater. Sci. 11 (2007) 3–18.
[23] S. Iguchi, Y. Miseki, K. Sayama, Efficient hypochlorous acid (HClO) production via [42] X. Yu, B. Kim, Y.K. Kim, Highly enhanced photoactivity of anatase TiO2 nano­
photoelectrochemical solar energy conversion using a BiVO4-based photoanode, crystals by controlled hydrogenation-induced surface defects, ACS Catal. 3
Sustain. Energy Fuels 2 (2018) 155–162. (2013) 2479–2486.
[24] Z. Zheng, Y.H. Ng, Y. Tang, Y. Li, W. Chen, J. Wang, X. Li, L. Li, Visible-light-driven [43] F. Cao, J. Xiong, F. Wu, Q. Liu, Z. Shi, Y. Yu, X. Wang, L. Li, Enhanced photoelec­
photoelectrocatalytic activation of chloride by nanoporous MoS2@BiVO4 pho­ trochemical performance from rationally designed anatase/rutile TiO2 hetero­
toanode for enhanced degradation of bisphenol A, Chemosphere 263 (2021) structures, ACS Appl. Mater. Interfaces 8 (2016) 12239–12245.
128279. [44] K.L. Schulte, P.A. DeSario, K.A. Gray, Effect of crystal phase composition on the
[25] D.K. Lee, K.-S. Choi, Enhancing long-term photostability of BiVO4 photoanodes for reductive and oxidative abilities of TiO2 nanotubes under UV and visible light,
solar water splitting by tuning electrolyte composition, Nat. Energy 3 (2018) 53–60. Appl. Catal. B 97 (2010) 354–360.
[26] J. Juodkazytė, M. Petrulevičienė, M. Parvin, B. Šebeka, I. Savickaja, V. Pakštas, [45] P. Bamola, A. Bhoumik, C. Dwivedi, V. Kaushik, H. Sharma, Enhanced photo­
A. Naujokaitis, J. Virkutis, A. Gegeckas, Activity of sol-gel derived nanocrystalline catalytic activity in TiO2 mixed phase nanostructures, Mater. Today Proc. 28
WO3 films in photoelectrochemical generation of reactive chlorine species, J. (2020) 32–36.
Electroanal. Chem. 871 (2020) 114277. [46] T. Gakhar, A. Hazra, Oxygen vacancy modulation of titania nanotubes by
[27] X. Zhang, S. Zhang, X. Cui, W. Zhou, W. Cao, D. Cheng, Y. Sun, Recent advances in cathodic polarization and chemical reduction routes for efficient detection of
TiO2-based photoanodes for photoelectrochemical water splitting, Chem. Asian volatile organic compounds, Nanoscale 12 (2020) 9082–9093.
J. 17 (2022) e202200668. [47] H. Huang, X. Hou, J. Xiao, L. Zhao, Q. Huang, H. Chen, Y. Li, Effect of annealing
[28] N.R. Reddy, P.M. Reddy, N. Jyothi, A.S. Kumar, J.H. Jung, S.W. Joo, Versatile TiO2 atmosphere on the performance of TiO2 nanorod arrays in photoelectrochemical
bandgap modification with metal, non-metal, noble metal, carbon material, and water splitting, Catal. Today 330 (2019) 189–194.
semiconductor for the photoelectrochemical water splitting and photocatalytic [48] X. Bi, G. Du, A. Kalam, D. Sun, Y. Yu, Q. Su, B. Xu, A.G. Al-Sehemi, Tuning oxygen
dye degradation performance, J. Alloy. Compd. 935 (2023) 167713. vacancy content in TiO2 nanoparticles to enhance the photocatalytic perfor­
[29] J. Lim, Y.U. Shin, A. Son, S.W. Hong, S. Hong, TiO2 nanotube electrode for organic mance, Chem. Eng. Sci. 234 (2021) 116440.
degradation coupled with flow-electrode capacitive deionization for brackish [49] H. Idriss, On the wrong assignment of the XPS O1s signal at 531–532 eV at­
water desalination, npj Clean Water 5 (2022) 7. tributed to oxygen vacancies in photo- and electro-catalysts for water splitting
[30] K. Arifin, R.M. Yunus, L.J. Minggu, M.B. Kassim, Improvement of TiO2 nanotubes and other materials applications, Surf. Sci. 712 (2021) 121894.
for photoelectrochemical water splitting: review, Int. J. Hydrog. Energy 46 [50] C. Fan, C. Chen, J. Wang, X. Fu, Z. Ren, G. Qian, Z. Wang, Black hydroxylated
(2021) 4998–5024. titanium dioxide prepared via ultrasonication with enhanced photocatalytic
[31] T.S. Rajaraman, S.P. Parikh, V.G. Gandhi, Black TiO2: a review of its properties and activity, Sci. Rep. 5 (2015) 11712.
conflicting trends, Chem. Eng. J. 389 (2020) 123918. [51] A. Hankin, F.E. Bedoya-Lora, J.C. Alexander, A. Regoutz, G.H. Kelsall, Flat band
[32] X. Huang, X. Gao, Q. Xue, C. Wang, R. Zhang, Y. Gao, Z. Han, Impact of oxygen potential determination: avoiding the pitfalls, J. Mater. Chem. A 7 (2019)
vacancies on TiO2 charge carrier transfer for photoelectrochemical water split­ 26162–26176.
ting, Dalton Trans. 49 (2020) 2184–2189. [52] X. Ma, X. Li, M. Mai, D. Lin, H. Zhou, L. Zhang, J. Li, Q. Li, D. Chen, Observation of
[33] X. Liang, Q. He, J. Zhang, X. Ding, Y. Gao, W. Chen, K.H.L. Zhang, C.Y. Haw, suppressed photocurrent of plasmonic Au on TiO2 by a double light beam
Enhanced photo-carrier transportation at semiconductor/electrolyte interface of method, Int. J. Hydrog. Energy 46 (2021) 5045–5052.
TiO2 photoanode by oxygen vacancy engineering, Appl. Surf. Sci. 597 (2022) [53] R.Van de Krol, M. Graetzel, Photoelectrochemical Hydrogen Production,
153744. Springer, New York, 2012.
[34] K.H. Kim, C.W. Choi, S. Choung, Y. Cho, S. Kim, C. Oh, K.S. Lee, C.L. Lee, K. Zhang, [54] S. Corby, R.R. Rao, L. Steier, J.R. Durrant, The kinetics of metal oxide photoanodes
J.W. Han, S.Y. Choi, J.H. Park, Continuous oxygen vacancy gradient in TiO2 pho­ from charge generation to catalysis, Nat. Rev. Mater. 6 (2021) 1136–1155.
toelectrodes by a photoelectrochemical-driven “self-purification” process, Adv. [55] J. You, L. Zhang, L. He, B. Zhang, Photocatalytic degradation of methyl orange on
Energy Mater. 12 (2022) 2103495. ZnO-TiO2/SO42- heterojunction composites, Opt. Mater. 131 (2022) 112737.
[35] J. Wu, Y. Tao, C. Zhang, Q. Zhu, D. Zhang, G. Li, Activation of chloride by oxygen [56] V.Q. Hieu, T.K. Phung, T.Q. Nguyen, A. Khan, V.D. Doan, V.A. Tran, V.T. Le,
vacancies-enriched TiO2 photoanode for efficient photoelectrochemical treat­ Photocatalytic degradation of methyl orange dye by Ti3C2-TiO heterojunction
ment of persistent organic pollutants and simultaneous H2 generation, J. Hazard. under solar light, Chemosphere 276 (2021) 130154.
Mater. 443 (2023) 130363. [57] J.C. Wang, H.H. Lou, Z.H. Xu, C.X. Cui, Z.J. Li, K. Jiang, Y.P. Zhang, L.B. Qu, W. Shi,
[36] Q. Zhang, D. Chen, Q. Song, C. Zhou, D. Li, D. Tian, D. Jiang, Holey defected TiO2 Natural sunlight driven highly efficient photocatalysis for simultaneous de­
nanosheets with oxygen vacancies for efficient photocatalytic hydrogen pro­ gradation of rhodamine B and methyl orange using I/C codoped TiO2 photo­
duction from water splitting, Surf. Interfaces 23 (2021) 100979. catalyst, J. Hazard. Mater. 360 (2018) 356–363.
[37] R. Fernández-Climent, S. Giménez, M. García-Tecedor, The role of oxygen vacancies [58] D.A.P. Velásquez, F.L.N. Sousa, T.A.S. Soares, A.J. Caires, D.V. Freitas, M. Navarro,
in water splitting photoanodes, Sustain. Energy Fuels 4 (2020) 5916–5926. G. Machado, Boosting the performance of TiO2 nanotubes with ecofriendly
[38] Y. Wang, J. Zhang, M.S. Balogun, Y. Tong, Y. Huang, Oxygen vacancy-based metal AgIn5Se8 quantum dots for photoelectrochemical hydrogen generation, J. Power
oxides photoanodes in photoelectrochemical water splitting, Mater. Today Sources 506 (2021) 230165.
Sustain. 18 (2022) 100118. [59] X. Ma, M. Mai, H. Lin, L. Zeng, J. Zhang, H. Zhou, D. Chen, A novel electrochemical
[39] J. Yu, B. Wang, Effect of calcination temperature on morphology and photo­ method for simultaneous measurement of real-time potentials and photo­
electrochemical properties of anodized titanium dioxide nanotube arrays, Appl. current of various photoelectrochemical systems, J. Power Sources 415 (2019)
Catal. B 94 (2010) 295–302. 99–104.

10

You might also like