Next Article in Journal
Preparation of Fenton Catalysts for Water Treatment
Previous Article in Journal
Structure-Sensitive Behavior of Supported Vanadia-Based Catalysts for Combustion of Soot
Previous Article in Special Issue
PtCu Nanoparticle Catalyst for Electrocatalytic Glycerol Oxidation: How Does the PtCu Affect to Glycerol Oxidation Reaction Performance by Changing pH Conditions?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emergent CuWO4 Photoanodes for Solar Fuel Production: Recent Progress and Perspectives

1
School of Energy and Chemical Engineering, Ulsan National Institute of Science and Technology (UNIST), 50 UNIST-gil, Ulsan 44919, Republic of Korea
2
Laboratory of Photonics and Interfaces, Institute of Chemical Sciences and Engineering, École Polytechnique Fédérale de Lausanne, 1015 Lausanne, Switzerland
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(11), 1408; https://doi.org/10.3390/catal13111408
Submission received: 23 September 2023 / Revised: 23 October 2023 / Accepted: 24 October 2023 / Published: 30 October 2023
(This article belongs to the Special Issue Theme Issue in Honor of Prof. Dr. Jae Sung Lee)

Abstract

:
Solar fuel production using a photoelectrochemical (PEC) cell is considered as an effective solution to address the climate change caused by CO2 emissions, as well as the ever-growing global demand for energy. Like all other solar energy utilization technologies, the PEC cell requires a light absorber that can efficiently convert photons into charge carriers, which are eventually converted into chemical energy. The light absorber used as a photoelectrode determines the most important factors for PEC technology—efficiency, stability, and the cost of the system. Despite intensive research in the last two decades, there is no ideal material that satisfies all these criteria to the level that makes this technology practical. Thus, further exploration and development of the photoelectode materials are necessary, especially by finding a new promising semiconductor material with a suitable band gap and photoelectronic properties. CuWO4 (n-type, Eg = 2.3 eV) is one of those emerging materials that has favorable intrinsic properties for photo(electro)catalytic water oxidation, yet it has been receiving less attention than it deserves. Nonetheless, valuable pioneering studies have been reported for this material, proving its potential to become a significant option as a photoanode material for PEC cells. Herein, we review recent progress of CuWO4-based photoelectrodes; discuss the material’s optoelectronic properties, synthesis methods, and PEC characteristics; and finally provide perspective of its applications as a photoelectrode for PEC solar fuel production.

Graphical Abstract

1. Introduction

The climate change due to rapidly increasing atmospheric CO2 concentration has become an imminent threat to the sustainability of human beings’ lives on earth [1]. Solar fuel production is considered an effective strategy to address this urgent global issue, which harvests solar energy and converts and stores it as chemical energy [2]. Three technologies are available for solar fuel production: the photovoltaic cell–electrolyzer (PV-EC) combination is highly efficient but expensive; particulate photocatalysis (PC) is low cost but also low efficiency; the photoelectrochemical (PEC) cell is positioned in the middle of these two technologies, having modest efficiency and cost [3,4]. All the three technologies need to be developed further to create a highly efficient but low cost practical system.
In particular, the PEC device can have a diverse configuration, ranging from a buried junction to a semiconductor-liquid junction (SCLJ) for photovoltage generation [3], and, thus, can utilize a variety of light absorber materials from PV grade materials to stabilize metal oxide semiconductors [4]. Since the long-term stability (>10 years) of aqueous electrolytes is essential for the PEC cells, SCLJ devises are the most staple and general of those used [5]. Also, fundamental studies of the charge transfer mechanism have been focused on SCLJ type PEC cells. For instance, the reaction order for PEC SCLJ water oxidation with Fe2O3 photoanode was found to be in the third order in respect to the hole concentration with an activation energy (Ea) of 60 meV, which is different from those for typical electrochemical water oxidation (the first order, Ea > 300 mV) [6]. Such uniqueness makes further study and the development of the SCLJ PEC cell more interesting and worthwhile.
The PEC water splitting starts from the absorption of photons with energy greater than the band gap energy (Eg) of the semiconductor (SC) comprising the photoelectrode [7]:
SC + hν (≥Eg) → ecb + hvb+: Light absorption
Electrons in the conduction band (CB) (ecb) with an energy level above 0.0 V vs. RHE diffuse to the surface and reduce water to generate hydrogen on a hydrogen evolution co-catalyst (HEC). Holes in the valence band (VB) (hvb+) with an energy level below 1.23 V vs. RHE also diffuse to the surface and oxidize water to generate oxygen, often on an oxygen evolution co-catalyst (OEC). This process leads to the overall water-splitting reaction and produces hydrogen from water using solar energy.
4H+ + 4ecb → 2H2: Hydrogen evolution reaction (HER)
2H2O + 4hvb+ → O2 + 4H+: Oxygen evolution reaction (OER)
Overall water splitting: 2H2O → 2H2 + O2, ∆G° = 238 kJ mol−1
Theoretically, a semiconductor can split water if its Eg is greater than 1.23 eV (water dissociation energy) and its CB and VB edge energy levels are more negative and more positive than the water reduction (0.0 VRHE) and oxidation (1.23 VRHE) potentials, respectively [8]. In reality, a much larger Eg than 1.23 eV is needed to overcome the overpotentials of the reactions (~0.05 V for HER and ~0.25 V for OER); thus, a total >1.5 V is needed [7,8]. In addition, the rate of reaction is much smaller than the rate of light absorption because of dominant electron–hole recombination, which is the main energy loss process.
General guidelines for selecting metal oxide light absorbers that could be used as photocatalysts for water splitting have often been suggested in previous reviews [7,9,10]. Metal oxide should have adequate energy levels for VBM and CBM, as well as chemical inertness against self-oxidation and a reduction in electrolytes. The VBM is mainly composed of the O 2p orbital of bonding state, while the CBM is often composed of d0 (Ti4+, Zr4+, Nb5+, Ta5+, V5+, W6+, and Ce4+) and d10 (Zn2+, In3+, Ga3+, Ge4+, Sn4+, and Sb5+) compounds. However, the majority these semiconductors have large band gaps (3.0 eV~3.2 eV) and only absorb UV light. For example, titanates (TiO2, SrTiO3) have proper band alignments for overall water splitting with high quantum yields, but they can only absorb UV light, which limits the solar-to-hydrogen (STH) conversion efficiency to a very low level (the reported highest value is 0.65% [11]).
This generated demand for metal oxide semiconductors that are active under visible light (Eg < 3.0 eV). Thus, WO3 (Eg = 2.8~2.6 eV) and Fe2O3 (Eg = 2.1 eV) have been extensively studied for generating photocurrent densities (Jph) above 4.0 mA/cm2 @ 1.23 VRHE under one sun irradiation (100 mW/cm2) [12].
As such a performance of the binary oxides was not satisfactory, tertiary metal oxides were introduced to form adequate electronic structures with smaller band gaps. One of the most successful cases is BiVO4 with VBM composed of Bi 6s and O 2p, which gives a respectable Eg of 2.4 eV. With a relatively high diffusion length and charge mobility, the state-of-the art BiVO4 photoanode records a Jph above 6.0 mA/cm2 @ 1.23 VRHE, which represents the highest among all metal oxide light absorbers [13]. CuWO4, which is one of the photoactive ternary tungstates (Bi2WO6, which showed PEC water oxidation activity [14], and ZnWO4, which showed pollutant degradation [15] and PEC water oxidation activity [16]), is also a tertiary metal oxide, with CBM composed of W5d and Cu3d and VBM composed of Cu 3d and O 2p, and it could be used as a photocatalyst and a photoanode for water oxidation [17,18]. Although it has a strong indirect band gap property, its smaller band gap (Eg = 2.3 eV) than that of WO3 (Eg = 2.7 eV) is quite attractive. Hence, here we discuss CuWO4’s potential as an emergent photoanode material for solar fuel production.

2. CuWO4 as a Photoanode Material

2.1. Crystal Structure of Photoactive CuWO4

The CuWO4 has a unique crystal structure made of distorted wolframite because W ions in WO3 are partially replaced by Cu ions. W6+ ions and Cu2+ ions are coordinated with O2− ions in octahedral sites (Figure 1a). This structure has a space group of P1¯ under low pressure at room temperature, with a = 4.694 A°, b = 5.830 A°, c = 4.877 A°, α = 91.64°, β = 92.41°, and γ = 82.91°. It is influenced by the Jahn–Teller effect caused by Cu2+ ions. This effect leads to distortions in the [CuO6] octahedral clusters, resulting in d orbital splitting and the breaking of the degeneracy of σ-antibonding orbitals. The unpaired electron in the dx2-y2 orbital of Cu2+ ions, as dictated by the Pauli exclusion principle, creates a mid-gap band state, with greater stabilization observed in Jahn–Teller-elongated Cu2+ ions, where the 3dz2 orbital contains two electrons [19,20].

2.2. Electronic Structure of Photoactive CuWO4

The electronic structure of CuWO4 was elucidated via various spectroscopies and density functional theory (DFT) calculations. The conduction band is primarily characterized by the contribution from Cu 3d orbitals, with the additional participation of W 5d orbitals. Conversely, the valence band exhibits a strong hybridization between O 2p and Cu 3d orbitals, as illustrated in Figure 1b. Thus, CuWO4 has a suitable band position for water oxidation, an indirect band gap of 2.3~2.4 eV (Cu 3d → Cu 3d), and a direct band of 2.6~2.7 eV (O 2p → Cu 3d). Its UV-vis spectrum features an absorption onset around 550 nm, reflecting a band gap energy of 2.3 eV (Figure 1c,d). A gradual rather than sharp absorption onset is characteristic of an indirect band gap. Such an indirect band gap tends to limit the light harvesting efficiency of CuWO4. The absorptivity coefficient α is markedly high at 6600 cm−1 at 400 nm, but it drops rapidly to 1715 cm−1 when the wavelength increases to 500 nm [18,20].
WO3 and CuWO4 share similar electronic structures, though a distinctive feature arises in their indirect band gaps. The CuWO4 possesses an indirect band gap 0.4 eV narrower than that of WO3. This smaller band gap allows CuWO4 to absorb longer-wavelength visible light photons, as shown in Figure 1c. The increased spectral absorption is reflected in the theoretical solar-to-hydrogen (STH) conversion efficiency of CuWO4 higher than 10%. However, the characteristics of the Cu 3d state in CuWO4 result in electron localization, which leads to a constraint in electronic properties, specifically a charge-carrier mobility in the order of ~6 × 10−3 cm2 V−1 s−1 and a diffusion length restricted to approximately 30 nm for photo-excited electrons in the CB. Such limitations inherently hinder the performance of CuWO4 photoanodes, causing a substantial divergence from their theoretical efficiency. Furthermore, CuWO4 exhibits commendable chemical and thermal stability in neutral electrolytes. This notable stability underscores the material’s potential for use in various technological applications [22,23].

3. Synthesis Chemistry and Photoelectrochemical Stability

3.1. Synthesis of CuWO4 Thin Film

J. E. Yourey et al. [21] initially presented CuWO4 as a promising photoanode material in 2011. The material was synthesized through electrodeposition (ED) using an acidic bath containing a mixture of copper and tungsten. The ED method involved potential sweeps from +0.3 V to −0.5 V, carried out over six cycles. The subsequent annealing of this film was conducted at 500 °C for 2 h. This sample showed 0.2 mA/cm2 at 1.23 VRHE in 0.1 M KPi under one sun illumination. Then, J. C. Hill and K. S. Choi [14] produced CuWO4 with an enhanced surface area by depositing an excess of Cu precursor onto electrochemically deposited porous WO3 electrodes. This approach was found to be adaptable for other tungstate materials. The CuWO4 films were thermally treated at 550 °C in air for 6 h and then immersed in a highly acidic solution to purify the deposited samples. This treatment resulted in an 8-micrometer thick film composed of nanoscale CuWO4 particles, attributed to the presence of the thick underlying WO3 film. It was ascertained that a high-temperature annealing process is vital to produce films that are both impurity-free and highly crystalline. They measured the PEC performances under varying pH conditions—strong acid (0.05 M H2SO4), neutral (0.1 M phosphate buffer), and weak alkaline conditions (0.1 M borate buffer)—to observe the highest Jph of 0.13 mA/cm2 at 1.23 VRHE under weak alkaline conditions.
A different method of CuWO4 synthesis was introduced by D. Hu et al. [24] using a solid-phase reaction with thermal assistance, circumventing the need for electrochemical procedures. The WO3 nanoflakes (NFs) fabricated through a hydrothermal method served as a sacrificial template for the CuWO4 synthesis. An aqueous solution of Cu(NO3)2 was meticulously drop-cast onto these WO3 nanoflake arrays and subsequently allowed to dry at a moderate temperature. The detailed fabrication procedure of the CuWO4/FTO photoanode is illustrated in Figure 2a. The transformation of the WO3 and Cu(NO3)2 mixture into CuWO4 was proposed to proceed through the following reactions:
2Cu(NO3)2(s) → 2CuO(s) + 4NO2(g) + O2(g)
WO3(s) + CuO(s) → CuWO4(s)
Upon the conclusion of this thermal process, a gray–yellow composite film of CuO and CuWO4 was produced. The residual CuO was effectively removed by submerging the film in a 0.5 M HCl solution for 30 min, yielding a vivid yellow film. The X-ray diffraction (XRD) analysis initially showed a distinct monoclinic WO3 pattern in a film, with specific diffraction peaks at 23.3° and 24.5°, as shown in Figure 2b. However, following a solid-phase reaction at 550 °C for 2 h and the elimination of excess CuO at the surface, the WO3 pattern disappeared completely, being replaced by a clear XRD pattern of triclinic CuWO4. This suggests the complete transformation of WO3 into CuWO4 at a lower temperature of 550 °C. Figure 2c exhibits representative SEM images of the resultant CuWO4 film, exhibiting that the flake nanostructure inherent to the WO3 template is mostly preserved. Notably, the synthesized CuWO4 NFs were slightly thicker than the original WO3 NFs. They claimed that this small alteration in thickness could be attributed to the thermal treatment and the solid-phase interaction between CuO and WO3. Owing to their distinctive morphological and crystalline properties, the CuWO4 NF films demonstrated higher PEC activity in comparison to previously reported CuWO4 synthesized via the ED method. Such a distinctive structure with a vast surface area would minimize hole transport distance from within CuWO4 to its interface and reduce the grain boundary concentration. A comparative study of CuWO4 NF films prepared under diverse annealing durations revealed that the sample annealed at 550 °C for 2 h showed the best performance, registering a photocurrent of 0.32 mA/cm2 at 1.23 VRHE in a 0.1 M NaBi buffer, as shown in Figure 2c. Z. Zhang et al. [25] studied the hydrothermal synthesis of CuWO4 films on FTO substrates using a seeding layer at a moderate temperature. Using a tungsten-rich reaction solution led to a WO3-CuWO4 composite film, which displayed enhanced PEC performance due to improved charge separation.
Figure 2. Solid-state reaction of WO3 into CuWO4 thin film. (a) Schematic illustration of the preparation procedure of the CuWO4 NF array film on an FTO substrate. Optical images represent a WO3 film, a WO3 film drop-cast with Cu(NO3)2 solution, a mixture film of CuWO4 and CuO, and a CuWO4 film. (b) XRD pattern of CuWO4 film annealed at 450 °C, 550 °C, and 650 °C. (c) SEM images of the WO3 and CuWO4 nanoflakes prepared at 550 °C for 2 h and (d) IV curve of CuWO4 nanoflake photoanode (0.1 M NaBi buffer, pH 9.0) under one simulated sun. Reproduced from ref. [24].
Figure 2. Solid-state reaction of WO3 into CuWO4 thin film. (a) Schematic illustration of the preparation procedure of the CuWO4 NF array film on an FTO substrate. Optical images represent a WO3 film, a WO3 film drop-cast with Cu(NO3)2 solution, a mixture film of CuWO4 and CuO, and a CuWO4 film. (b) XRD pattern of CuWO4 film annealed at 450 °C, 550 °C, and 650 °C. (c) SEM images of the WO3 and CuWO4 nanoflakes prepared at 550 °C for 2 h and (d) IV curve of CuWO4 nanoflake photoanode (0.1 M NaBi buffer, pH 9.0) under one simulated sun. Reproduced from ref. [24].
Catalysts 13 01408 g002
C. M. Tian et al. [18] presented a solution-based synthesis approach that produced thick, high-quality, and polycrystalline CuWO4 films. These films had distinct optoelectronic characteristics and a notable Jph of ~0.5 mA/cm2 at 1.23 VRHE. Later, J.U. Lee et al. [26] developed a greenish, transparent CuWO4 polycrystalline film on FTO using a spin coating method, followed by annealing at 550 °C for 2 h. This synthesis protocol and optical images are shown in Figure 3a,b. The absorption spectrum of the three-layered CuWO4 film in Figure 3c exhibits an absorption edge extended to 510 nm. The Tauc plots reveal indirect and direct band gaps of 2.43 eV and 2.89 eV, respectively, which are similar to the existing literature on polycrystalline CuWO4 films. Moreover, the XRD for the three-layered CuWO4 electrodes in Figure 3d clearly displays the peaks associated with the triclinic CuWO4 phase. The surface and cross-sectional images of the CuWO4 film observed via SEM in Figure 4d exhibit a worm-like grain morphology, forming a porous structure with an approximate thickness of 650 nm and feature sizes of 100–200 nm (Figure 4e) [26]. M.J.d.S. Costa et al. [27] compared various binary metal tungstates (AWO4, A2+ = Fe, Cu, Ni, and Co) as photoanode materials synthesized via a sol–gel method using a complexation of metal alkoxides and an esterification/polymerization reaction. Among AWO4 materials, CuWO4 showed the best PEC oxidation performance, recording a Jph of 30 uA/cm2 @ 0.6 Ag/AgCl, while the position of the flat band potential (Efb) related to photovoltage was the lowest. X. Duan et al. [28] reported a controllable fabrication method of ultrathin CuWO4 films via an automatic ultrasonic spray pyrolysis method. They extensively explored the effects of varying tungsten sources and film thicknesses on the photoelectrochemical (PEC) performances of the synthesized CuWO4 films. In particular, a CuWO4 film derived from ammonium meta tungstate with a thickness of approximately 2.16 μm demonstrated the best PEC performance (41 μA/cm2 at 1.23 VRHE). Then, N. Gaillard et al. [29] reported nanocomposite CuWO4 thin films produced using spray pyrolysis from solutions of copper acetate, ammonium meta tungstate, and MWCNT, resulting in porous, crack-free polycrystalline CuWO4 films with nanoparticle sizes of 10–50 nm. The electrochemical impedance (EIS) tests under one sun illumination showed a 30% reduction in bulk resistance for nanocomposite CuWO4 photoanodes compared to the bare one. The CuWO4 nanocomposite revealed improved a Jph of 0.38 mA/cm2 at 1.63 VRHE in a pH 10 buffer, and MNCNT acted as effective electron collectors throughout the CuWO4 bulk.
Meanwhile, Y. Gao et al. [30] fabricated CuWO4 with a Cu:W ratio of 1:1 via the atomic layer deposition method. Subsequent annealing at 600 °C for 30 min facilitated a solid-state reaction between CuO and WO3, resulting in CuWO4 films with elaborately controlled thicknesses and compositions. This CuWO4 showed 0.11 mA/cm2 at 1.23 VRHE in 0.2 M pH 9 KCl. The chemical vapor deposition method was applied by D. Peeters et al. [23] as another synthesis method that makes it easy to control the stoichiometry of Cu and W in CuWO4 by employing [Cu(etaoac)2] and [W(NtBu)2(dpamd)2] precursors. By manipulating process parameters, they varied Cu-to-W ratios, enabling in-depth studies of the effects of stoichiometry on optical and PEC properties. The CuWO4 of Cu:W = 1:1 showed the best performance (Jph = 0.06 mA cm−2) at 1.23 VRHE under frontside one sun illumination.
C. M. Gonzalez et al. [31] proposed pulsed laser deposition (PLD) onto an insulating substrate for synthesizing CuWO4 thin films. They investigated the temperature dependence of electronic conductivity for CuWO4 films across the 100–500 °C temperature range. Although they did not measure the PEC performance, they showed that CuWO4 can be synthesized via PLD. A. Hrubantova et al. [32] made CuWO4 films via reactive high-power impulse magnetron sputtering (HiPIMS) in an argon/oxygen gas mixture using two pieces of magnetron for Cu and W. They observed that both the composition and crystal structure of the as-deposited and post-annealing films depended on the deposition conditions. The samples synthesized on the FTO compromised the WO3 and CuWO4 or Cu2WO4 phases.
Thus, CuWO4 photoanodes can be synthesized via a variety of methods, but none of them has distinguished itself through outstanding PEC performance. Further exploration of more effective synthesis methods is in order.

3.2. Stability of CuWO4

A fundamental requirement for photoelectrodes is their chemical stability in aqueous solutions under light irradiation. The photostability of CuWO4 photoanodes was assessed in comparison to WO3 (Figure 4a,b) in different electrolytes (KPi and KBi buffer, pH 7) (Figure 4c–f). In KBi buffer at pH 7, CuWO4 photoanodes exhibited remarkable stability, maintaining 93% of their initial photocurrent density over a 12-h period. Conversely, in KPi buffer of the same pH, the photocurrent experienced a 50% decline within the initial 4 h and further deteriorated to only the 15% level after a 12-h period. This stark difference implies that the phosphate anion impairs the stability of CuWO4 in aqueous solutions. The CuWO4 electrodes also were tested under higher irradiance conditions to evaluate the degradation of the electrode. The rate of electrode degradation increased in all buffer solutions. The overall trend remained the same, with CuWO4 photoanodes exhibiting better stability in the KBi buffer compared to the KPi buffer.
The CuWO4 photoanodes exhibit a faradic efficiency approaching unity for water oxidation in an equimolar KBi and NaCl solution at pH 7, underscoring their potential utility in this application. In particular, the chronoamperometry (j-t) profiles for 0.1 M KBi and 0.1 M KBi with 100 mM NaCl remain stable under irradiation of both 100 and 300 mW/cm2 [33]. In contrast, WO3 begins to dissolve within the initial 30 min of illumination. After a 12-h soaking in KPi electrolyte, the WO3 film displayed a substantial reduction in photocurrent from 0.17 to 0.03 mA/cm2. This enhanced stability of the CuWO4 film arises from a boosted covalency in Cu-O bonds, which hinders the acid−base reaction in WO3 at >pH 5. Moreover, such stability is preserved during water splitting, which is distinctly different to the WO3 case that generates peroxo intermediates during the reaction, thus hastening its degradation [17,21].
Figure 4. Stability of CuWO4 photoanode in various conditions. (a) IV curves of CuWO4 and WO3. The blue and green traces show photocurrent density after soaking the electrodes in 0.1 M KPi overnight for CuWO4 and WO3, respectively. The photocurrent densities for both of the fresh photoanodes are shown in red. The black curves represent dark currents. (b) Chronoamperometry curves of water electrolysis with CuWO4 (black) and WO3 (green) at 0.5 V vs. Ag/AgCl in a 0.1 M phosphate buffer (pH 7). Reproduced from ref. [21]. CuWO4 photoanode tested in different pHs (c,d) and electrolytes (e,f). (c,e) IV curves and (d,f) chronoamperometry with an applied potential of 1.23 VRHE. All experiments were conducted under simulated AM 1.5G illumination (100 mW/cm2). Reproduced from ref. [33].
Figure 4. Stability of CuWO4 photoanode in various conditions. (a) IV curves of CuWO4 and WO3. The blue and green traces show photocurrent density after soaking the electrodes in 0.1 M KPi overnight for CuWO4 and WO3, respectively. The photocurrent densities for both of the fresh photoanodes are shown in red. The black curves represent dark currents. (b) Chronoamperometry curves of water electrolysis with CuWO4 (black) and WO3 (green) at 0.5 V vs. Ag/AgCl in a 0.1 M phosphate buffer (pH 7). Reproduced from ref. [21]. CuWO4 photoanode tested in different pHs (c,d) and electrolytes (e,f). (c,e) IV curves and (d,f) chronoamperometry with an applied potential of 1.23 VRHE. All experiments were conducted under simulated AM 1.5G illumination (100 mW/cm2). Reproduced from ref. [33].
Catalysts 13 01408 g004

4. Photoelectrochemical Properties of CuWO4 Photoanodes

As discussed in previous sections, CuWO4 meets important criteria for being a suitable photoanode material to be used in aqueous electrolytes. As an n-type semiconductor, it has an appropriate electronic structure for hole-driven oxidation, with a VBM potential (2.85 VRHE) deep enough to overcome the water oxidation redox potential (1.23 VRHE) and a proper band bending at the semiconductor–electrolyte interface for smooth hole transfer [17]. Its inherent stability in near-neutral electrolytes under light irradiation warrants diverse applications involving hole oxidation.
Before the fabrication of a photoelectrode, it is good practice to test the photoactivity of the material in a simpler system, like photovoltaics or a photocatalyst. CuWO4 was found to conduct photocatalytic water oxidation in an electron scavenger (Ag+ or Fe2+)-containing solution, thus demonstrating that it is a good candidate material for a photoanode. Thus, Z. Wu et al. demonstrated that powder-type CuWO4 showed n-type conductivity and the formation of a Schottky junction via contact potential difference (CPD) analysis, as well as photocatalytic water oxidation capability using an electron scavenger [34].
Studies of fundamental PEC properties indicated that thin films of CuWO4 could successfully form a SCLJ. Mott–Schottky (MS) plots in various electrolytes confirmed n-type conductivity with a flat-band potential (Efb) around 0.5 VRHE and a quasi-Fermi level for the electron (Efn) around 0.63 VRHE. The majority carrier density (ND) from the slope of the MS plot was ~10−19 cm−3, which is comparable to that of BiVO4. The photovoltage (Vph) of CuWO4 was ~0.6 V without the electrocatalyst. The maximum Vph can potentially reach 0.8 V if there is perfect interaction with Efn–E(O2/H2O) with no Fermi level pinning. It is slightly lower than that of BiVO4 (0.05 VRHE–1.23 VRHE = ~1.2 V) but comparable to that of Fe2O3 (0.3 VRHE–1.23 VRHE = 0.9 V) [26,35,36]. Besides establishing photovoltage at SCLJ, the formation of water oxidation intermediates at SCLJ was also confirmed in a similar manner to the cases of Fe2O3 and BiVO4 [35].
A physical model for charge-carrier pathways in CuWO4 under illumination is illustrated in Figure 5a—excitation, mid-gap state trapping, and charge–transfer reactions to the solution. The holes originating from VB possess the potential to directly participate in water oxidation or be localized in an intermediate state. The photoexcited electrons in CB move toward the F-doped SnO2 (FTO) substrate, while some of them recombine with holes or migrate to the surface intermediates. The Nyquist plot for CuWO4 demonstrates two independent charge transfer processes at the surface (Figure 5b). The semicircle known as RC1 is linked to non-Faradaic process in the depletion zone, represented as a parallel relationship between Rtrap and Csc, where Rtrap is the resistance to trapping/detrapping electrons in/out of the mid-gap state, and Csc is the capacitance of the space–charge region. Conversely, RC2 is related to a Faradaic reaction with the solution represented by Rct,mg and Cmg, where Rct,mg is the resistance to charge transfer at the surface/solution interface (to perform water oxidation) from the mid-gap state, and Cmg is the capacitance of the mid-gap state. As the applied voltage increases, the resistance associated with water oxidation (Rct,mg) decreases, while the resistance associated with electron trapping (Rtrap) increases. This behavior suppresses the recombination of photo-induced electrons at the mid-gap and facilitates electron transfer to the charge collector, initiating OER at 0.80 V. The capacitance of the space–charge region (CSC) remains constant at potentials exceeding 0.86 V, indicating the continuous flow of electrons to the charge collector. In contrast, the capacitance of the mid-gap state (Cmg) continues to increase after 0.8 V, representing the charging and discharging of this state as water oxidation proceeds. The CuWO4 showed a remarkable photocurrent increase and unpinning of the Fermi level at 1.06 V. This observation indicates that as the reaction proceeds, electrons from the solution populate the mid-gap state, and the increase in Cmg suggests that hole transfer from the valence band to the mid-gap state becomes rate limiting [37].
The onset potentials of CuWO4 for the PEC oxidation of water and Na2SO3 (a hole scavenger) were found to be around 0.85 VRHE and 0.6 VRHE, respectively (Figure 5c). The observed shift in onset potential between the water and Na2SO3 oxidation processes is potentially due to the suboptimal hole collection during water oxidation at the CuWO4 interface. The EIS analysis was conducted for CuWO4 in H2O and Na2SO3 solution to explain the observation. The derived surface state capacitance (Css) values diminished as the light intensity was reduced from 1 sun to 0.1 sun (Figure 5d). The peak Css at ~1.0 VRHE is close to its onset potential for water oxidation, while IV curves indicated that water oxidation intermediates form at ~0.8 VRHE. Furthermore, the bulk capacitance (Cbulk) values decreased with the applied voltage, but they were essentially invariant at a given potential under different light intensities. Using these Cbulk values, MS plots were formulated in Figure 5e, which gave Efb and ND values comparable to the literature’s values, as described above [35].
Intensity modulated photocurrent spectroscopy (IMPS) was employed to quantify hole transfer kinetics at SCJL [35,38]. A large impedance of hole transfer at SCLJ turned out to be the major huddle of CuWO4 for water oxidation, just like other known metal oxide semiconductors.

5. Modification Strategies of CuWO4 Photoanodes

Despite having only recently emerged, the PEC water oxidation performance of CuWO4 photoanodes has steadily improved through various modification strategies, as summarized in Table 1. These strategies include n-type doping, solid solution, heterojunction, electrocatalyst, and post-treatment, and each of these strategies is discussed in detail in this section.

5.1. Extrinsic/Intrinsic Defect Engineering via Doping

The extrinsic defect engineering conducted via the doping of foreign elements can promote the electron conductivity of CuWO4 by generating additional charge carriers, which can be characterized using the MS equation.
1 C 2 = 2 ( V E f b k T e ) e ε ε 0 N D A 2
where C = the capacitance of the photoanode (metal oxide + electrolyte double layer, etc.), e = the charge of the electron, ε = the dielectric constant of CuWO4, ε0 = the permittivity of the vacuum, V = the applied bias (vs. RHE), Efb = the flat band potential (vs. RHE), k = the Boltzmann constant, ND = the donor density of the n-type semiconductor (cm−3), A = the surface area of photoanode, and T = the temperature (K). There are two kinds of doping methods: one uses a non-isovalent dopant, and the other uses an isovalent dopant. The non-isovalent dopants can enhance electron conductivity by elevating the charge-carrier density of CuWO4, while the isovalent doping also boosts conductivity by improving the charge migration, although it cannot enhance the charge density. The dopant atoms can substitute into Cu2+ or W6+ octahedral sites, and their efficacy is determined by their suitable oxidation state, ionic size, and electronic configuration.
Fe doping was first applied to CuWO4 as extrinsic doping via spray pyrolysis using a FeSO4 precursor [41]. The iron was selected as a suitable dopant due to its prevalent non-toxic characteristics, its similarity in terms of the ionic radii of Fe3+ and Fe2+ to the target Cu2+ ion, and its good compatibility with the crystalline structure of CuWO4. The reaction (8) expresses the defect sites formed by Fe3+ substituting into the CuWO4 lattice, as delineated by the Kröger–Vink notation:
F e 2 O 3 + 2 W O 3 2 W W x + 8 O O x + 1 2 O 2 g + 2 F e C u + 2 e
The CuWO4 doped with 0.3% Fe exhibited a 50% increase in photocurrent density compared to the undoped one at 1.23 VRHE in 0.1 M KPi with an H2O2 hole scavenger. Furthermore, the charge separation efficiency at 1.23 VRHE was 50% higher for the doped sample.
Yttrium was suggested as another extrinsic dopant by S. G. Poggini et al. [43]. Y-doping was conducted via a dip coating method. The Y-doping into CuWO4 thin film led to a minor blue shift in the band gap from that of pristine CuWO4 (2.30 eV) due to a rise in the conduction band position, as supported by DFT calculations. An optimized 5% Y-doped sample showed a photocurrent density 92.5% higher than that of an undoped sample at 1.3 VRHE.
Another effective doping method for the CuWO4 crystal lattice is to generate oxygen vacancies ( V Ö ) through H2 treatment, N2 treatment, or F doping. The H2 treatment increases the charge-carrier density through the partial reduction of the metal oxide:
Intrinsic   doping :   O O x + H 2 ( g ) H 2 O g + V Ö + 2 e
Y. Tang et al. [62] conducted H2 treatment for CuWO4 to induce oxygen deficiency in its crystal lattice (Figure 6a,b). The H2-treated CuWO4 film showed a noticeable color change from bright yellow to dark yellow as the annealing time increased because V Ö can occur at a color center located in the band gap. This color change hints at a possible alteration in the CuWO4 band gap caused by H2 treatment. The UV–visible absorption spectra of bare and modified CuWO4 samples shown in Figure 7a reflect this color change. The 30-min treatment gave the best performance, showing a negative shift of overpotential from 0.35 V to 0.05 V vs. Ag/AgCl, as well as an increased donor density in MS analysis compared to bare CuWO4 (Figure 7b).
The use of nitrogen post-treatment to enhance the PEC performance of CuWO4 by forming oxygen vacancy was suggested by Z. Ma et al. [63]. A nitrogen atmosphere can provide a moderate reductive environment for CuWO4, which neither collapses the crystal structure nor creates a color center in the band gap. As a result, there is no color change in the CuWO4 films. The N2-treated CuWO4 sample annealed at 623 K showed 80 µA/cm2 @ 1.23 VRHE in 0.1 M KPi under one sun illumination.
C. Li and P. Diao [42] introduced a straightforward method for fluorine doping, where F anions were substituted into the crystal lattice of CuWO4 nanoflakes, replacing oxygen atoms and acting as potent electron donors to increase electron density. By modulating the volume of the F precursor solution, the dopant concentration was precisely controlled. The F-doped CuWO4 nanoflakes demonstrated significantly enhanced PEC performance coupled with robust stability in PEC oxygen evolution reactions. The optimized 2.5% F-doping enhanced the photocurrent density of CuWO4 from 0.32 to 0.57 mA/cm2 at 1.23 VRHE.
Figure 6. H2 gas treatment for partial reduction of CuWO4 photoanode. (a) UV-visible absorption spectra of CuWO4 film with H2 gas treatment (300 °C, H2 5%, Ar 95% flow) with different treatment durations. Inset is an image of a CuWO4 film with H2 gas treatment. (b) IV curves of CuWO4 with H2 gas treatment (0.1 M Na2SO4, under simulated one sun). Inset represents the photocurrent densities at an applied bias of 1.01 VAg/AgCl. Reproduced from ref. [62].
Figure 6. H2 gas treatment for partial reduction of CuWO4 photoanode. (a) UV-visible absorption spectra of CuWO4 film with H2 gas treatment (300 °C, H2 5%, Ar 95% flow) with different treatment durations. Inset is an image of a CuWO4 film with H2 gas treatment. (b) IV curves of CuWO4 with H2 gas treatment (0.1 M Na2SO4, under simulated one sun). Inset represents the photocurrent densities at an applied bias of 1.01 VAg/AgCl. Reproduced from ref. [62].
Catalysts 13 01408 g006
Figure 7. CuW1−xMoxO4 solid solution for photoanode. (a) IV curves of CuWO4 and solid solutions (W/Mo = 0.5/0.5 and 0.35/0.65) made via the electrodeposition (ED) of the precursor and thermal conversion (0.1 M KPi, pH 7.0, under simulated one sun). (b) IPCE under applied voltage of 1.61 VRHE, (c) Mott–Schottky plots of CuWO4 and CuW0.35Mo0.65O4 obtained in 0.1 M KPi (pH 7), and (d) band structures of CuWO4, Type-III CuMoO4, and CuW0.5Mo0.5O4 solid solutions, aligned by Mo (W) 4s, 4p (5s, 5p) core states (left) and projected density of states (PDOS) for a CuW0.5Mo0.5O4 solid solution (right), obtained from DFT calculations (a: the Cu-O state, b: the Cu-O state with underestimated the local spin density approximation (LDA + U), c: the hybrid W/Mo 5d/4d and O 2p states). Reproduced from ref. [45]. (e) Images of CuW1−xMoxO4 film made via spray pyrolysis. (f) IV curves (0.1 M KPi pH 7.0, under simulated one sun, 0.1 mol% H2O2 as hole scavenger). (g) Mott–Schottky plots of CuW1−xMoxO4 film. Reproduced from ref. [46].
Figure 7. CuW1−xMoxO4 solid solution for photoanode. (a) IV curves of CuWO4 and solid solutions (W/Mo = 0.5/0.5 and 0.35/0.65) made via the electrodeposition (ED) of the precursor and thermal conversion (0.1 M KPi, pH 7.0, under simulated one sun). (b) IPCE under applied voltage of 1.61 VRHE, (c) Mott–Schottky plots of CuWO4 and CuW0.35Mo0.65O4 obtained in 0.1 M KPi (pH 7), and (d) band structures of CuWO4, Type-III CuMoO4, and CuW0.5Mo0.5O4 solid solutions, aligned by Mo (W) 4s, 4p (5s, 5p) core states (left) and projected density of states (PDOS) for a CuW0.5Mo0.5O4 solid solution (right), obtained from DFT calculations (a: the Cu-O state, b: the Cu-O state with underestimated the local spin density approximation (LDA + U), c: the hybrid W/Mo 5d/4d and O 2p states). Reproduced from ref. [45]. (e) Images of CuW1−xMoxO4 film made via spray pyrolysis. (f) IV curves (0.1 M KPi pH 7.0, under simulated one sun, 0.1 mol% H2O2 as hole scavenger). (g) Mott–Schottky plots of CuW1−xMoxO4 film. Reproduced from ref. [46].
Catalysts 13 01408 g007

5.2. Formation of Solid Solution with Mo

Of the modification strategies developed to improve the PEC performance of CuWO4 photoanodes, solid solution with Mo was the most effective and innovative [45,46,48]. The Mo6+ ion can easily replace W6+ in the CuWO4 crystal structure because they have similar ionic radii and the same valance. The incorporation of Mo reduces the band gap and augments the light harvesting in the visible light region. The reduced band gap for CuW1−XMoXO4 is caused by the shift of the CBM to a low energy as Mo orbitals are mixed with Cu and W orbitals. This shift can push Efb below the water reduction potential and make the overall water splitting impossible, but the PEC water oxidation efficiency improves significantly owing to extended light harvesting.
J. C. Hill et al. [45] demonstrated, for the first time, CuW1−XMoXO4 photoelectrode prepared using ED and thermal conversion methods. They identified an optimal photoelectrode of CuW0.35Mo0.65O4 with a reddish color that showed the highest PEC performance (Figure 7a). To ascertain that the highest photocurrent comes from extended light harvesting, the incident photon-to-current conversion efficiency (IPCE) of CuWO4 and CuW0.35Mo0.65O4 are presented in Figure 7b. The photocurrent onset for CuWO4 was 560 nm, whereas it shifted to 620 nm for CuW0.35Mo0.65O4. This demonstrates that the increased photon absorption between 560 and 620 nm was directly translated into elevated photocurrents. In addition, CuW0.35Mo0.65O4 showed substantially superior IPCEs across the entire visible light region. The MS plots of CuWO4 and CuW0.35Mo0.65O4 were displayed to discern their Efb values (Figure 7c). Both samples exhibit marginal frequency-dependent variations, but on an aggregate scale, CuW0.35Mo0.65O4 possesses a more positive Efb than CuWO4. This implies that the augmented photocurrents of CuW0.35Mo0.65O4 were caused by an increase in light absorption rather than Efb. Moreover, the slopes of the MS plots for both samples appear similar at each frequency, suggesting that the Efb was shifted by the CBM change rather than the variation in charge-carrier density. The band structures of CuWO4, Type-III CuMoO4, and CuW0.5Mo0.5O4 solid solutions were calculated via DFT. The VBMs of all samples predominantly comprise hybridized Cu 3d and O 2p states. But the CBMs of CuW0.5Mo0.5O4 and Type-III CuMoO4 are mainly composed of hybridized Mo 4d and O 2p states, while the CBM of CuWO4 consists of hybridized W 5d and O 2p states, as shown in Figure 7d.
Meanwhile, CuW1−XMoXO4 was synthesized via spray pyrolysis by Q. Liang et al. [46], who simply adjusted the ratio by adding an Mo source instead of a W source. As the Mo ratio increased, the color of CuW1−XMoXO4 changed to reddish, and when the Mo ratio was higher than 35%, Cu3Mo2O9 impurity began to appear. Fortunately, the number of impurities was small up to 50% Mo content but increased rapidly beyond that level. These impurities can create a color center within the band and darken the color of CuW1−XMoXO4 samples (Figure 7e). The PEC performances of CuW1−XMoXO4 samples were measured in 0.1 M phosphate buffer with or without H2O2 under simulated one sun, and CuW0.5Mo0.5O4 showed the best performance in both electrolytes. In water oxidation, it generated ~0.47 mA/cm2 at 1.23 VRHE under one sun illumination, which was ~7 times higher than that of bare CuWO4. In H2O2 oxidation, CuW0.5Mo0.5O4 recorded ~0.95 mA/cm2 at 1.23 VRHE, and its onset potential shifted by 80 mV in a cathodic direction due to the enhanced reaction kinetics (Figure 7f). The Efb and charge density (ND) of CuW1−XMoXO4 were calculated from MS plots, as shown in Figure 7g. CuW0.65Mo0.35O4 and CuW0.5Mo0.5O4 showed more positive Efb than CuWO4, while ND values estimated from the slopes of the MS plots were higher for bare CuWO4. The authors claimed that CuW0.5Mo0.5O4 had enhanced conductivity and mitigated charge recombination.
A. Polo et al. [48] produced a CuW0.5Mo0.5O4 photoelectrode by stacking layers using a spin coating method. They found that a three-layered sample was optimal, and using four or more layered samples led to electron mobility issues, reducing the performance. The three-layered CuW0.5Mo0.5O4 generated a photocurrent of ~0.25 mA/cm2 at 1.23 VRHE in 0.1 M K3BO3, which was remarkably higher than CuWO4 prepared via a similar method. Utilizing a similar spin coating method, K. Wang et al. [64] achieved a gradient distribution of Mo within the CuWO4 film while doping the surface layer with Ni. The Mo incorporation caused the CBM to shift to a higher energy level, while the Ni doping induced the VBM to shift to a slightly lower energy, promoting efficient charge separation and alleviating charge recombination. As a result, the fully modified CuWMoO4 revealed 0.85 mA/cm2 @ 1.23 VRHE. Lastly, J. Yang et al. [47] made CuW1−XMoXO4 nanoflake films using hydrothermal and thermal solid-state reactions. They first synthesized Mo-WO3 nanoflakes via a solid-state reaction, with desired amounts of Mo and W placed on the WO3 seed layer. Then, a Cu precursor was deposited on it and heated to obtain a CuW1−XMoXO4 NFs film. The CuW0.68Mo0.32O4 exhibited the best performance of ~0.62 mA/cm2 at 1.23 VRHE in 0.1 M sodium phosphate buffer solution.

5.3. Heterojunction Electron Transfer Layer

An effective heterojunction electron transfer layer (ETL) should possess the following characteristics: (i) enhanced electron transfer properties compared to CuWO4, (ii) a type II staggered band alignment facilitating spontaneous electron transfer while blocking hole transfer, and (iii) a multidimensional high-aspect-ratio structure accompanied by porosity, enabling significant CuWO4 loading while maintaining a modest thickness [65]. WO3 has been commonly used as an ETL to improve the bulk charge separation efficiency for many photoanodes, such as Fe2O3 [66] and BiVO4 [67], because it has an appropriate band gap (~2.7 eV), a high electron mobility (~10 cm2V−1S−1), and a long hole diffusion length (~800 nm) [68]. Moreover, the synthesis method of CuWO4 from WO3 is well known, and CuWO4/WO3 samples of various structures have also been reported.
D. Wang et al. [49] synthesized CuWO4/WO3 porous thin films through magnetron sputtering, employing a polymer templating approach. Utilizing a dip coating process, a Cu precursor with polymer surfactant was deposited on a WO3 film. Then, CuWO4/WO3 was made via annealing at 550 °C for 2 h (Figure 8a). Figure 8b shows a cross-sectional SEM image of the CuWO4/WO3 composite, which clearly reveals respective WO3 and CuWO4 layers. The WO3 seed layer became a little thinner as CuWO4 of a thickness of 600 nm was formed. The CuWO4/WO3 composite exhibited a remarkably higher PEC performance of ~0.45 mA/cm2 at 1.2 VRHE compared to either the WO3 or CuWO4 photoanodes, as shown in Figure 8c. The IPCE for CuWO4/WO3 thin film at 1.2 VRHE, as shown in Figure 8d, was much higher than those of the WO3 and CuWO4 films. Although CuWO4 can absorb photons up to 540 nm, both samples revealed very low IPCEs between 470 and 540 nm, indicating that the charge carriers generated from CuWO4 in this 470~540 nm range contribute little to the photocurrent generation. The CBs of CuWO4 and WO3 are positioned at +0.2 eV and +0.4 eV, respectively, whereas the VBs of CuWO4 and WO3 are positioned at +2.4 eV and +3.0 eV, respectively. When they form a CuWO4/WO3 heterojunction, the photo-induced holes from WO3 migrate to CuWO4, while electrons travel in the opposite direction to the FTO charge collector, as shown in Figure 8e.
T. Wang et al. [50] produced urchin-like nanoarray CuWO4/WO3 films through a one-step hydrothermal reaction by controlling the reaction time. The optimized CuWO4/WO3 nanoarray film demonstrated an onset potential of 0.6 VRHE and a photocurrent density of 0.48 mA/cm2 at 1.23 VRHE in pH 7 solution. An enhanced PEC performance of the film can be attributed to its distinctive urchin-like nanoarray structure that provides a large surface area, in addition to facile charge separation by forming an effective CuWO4/WO3 heterojunction. I. Rodríguez-Gutiérrez et al. also synthesized CuWO4/WO3 photoelectrodes from WO3 nanorod seed [39]. Through various electrochemical analyses, it was shown that WO3 was partially or completely converted to CuWO4, depending on the annealing temperature. It was found that only CuWO4 was present in the sample that was annealed at temperatures of 650 °C or higher. However, the highest efficiency (~0.3 mA/cm2 @ 1.23 VRHE in 0.1 M phosphate buffer) was observed for the sample annealed at 450 °C, which contained a significant amount of WO3. In addition, CuWO4 served as a protective layer for the WO3 material that is not generally stable in neutral aqueous solutions.
Meanwhile, another common ETL material, namely SnO2, was also applied to CuWO4 film, which was synthesized via a spin coating method [26]. Although the CBM of SnO2 is located at a higher energy than that of CuWO4, its high electronic conductivity allows electrons to easily move to SnO2, and the VBM of SnO2 prevents the hole transfer from CuWO4. As a result, the CuWO4/SnO2 sample showed a photocurrent of ~0.1 mA/cm2 (at 1.23 VRHE in 0.1 M KPi) and a bulk charge separation efficiency twice as high as that of bare at 1.23 VRHE. The IPCE of CuWO4/SnO2 showed a diminished disparity in front/back side illumination conditions compared to bare CuWO4. This suggests a more effective electron transfer for the CuWO4/SnO2 via the hole-blocking effect. In addition, SnO2 ETL can inhibit the electron–hole recombination at the interface of the light absorber and charge collector.

5.4. Surface Modification via Co-Catalyst Loading

The surface modifications of CuWO4 usually target the enhancement of charge transfer from a semiconductor to an electrolyte and stability under sun illumination. A simple but most effective way to modify the surface properties is to deposit an electrocatalyst on the semiconductor’s surface. The electrocatalyst (termed co-catalyst on photoelectrode) plays multiple roles, such as (i) reducing the activation energy required for water oxidation, (ii) passivating the surface state or defects on the semiconductor, and (iii) storing photogenerated holes to promote a water oxidation reaction that requires multi-hole transfer [69]. The last two functions distinguish the co-catalyst on the photoanode from the typical electrocatalyst using only the first function.
A cobalt phosphate (Co-pi) is the best-known co-catalyst for oxygen evolution reaction (OER), as it is highly effective with various photoanodes. Figure 9a shows the band alignment and working mechanism when Co-Pi is loaded on CuWO4. This scheme illustrates that Co-Pi effectively passivates the surface state of CuWO4 to improve hole transfer efficiency and increase photovoltage [26,70]. S. Chen et al. [52] reported that Co-Pi could not induce a significant negative shift in the onset potential, but Co-Pi/CuWO4 exhibited a higher photocurrent density than bare CuWO4 by 86% at 1.23 VRHE in 0.1 M KPi, as shown in Figure 9b. EIS data of bare CuWO4 at 0.3 VRHE were fitted into an equivalent circuit model comprising a single RC circuit related to the semiconductor–electrolyte interface (Figure 9c). In contrast, EIS data of Co-Pi/CuWO4 aligned closely with an equivalent circuit model featuring two RC circuits. These circuits can be associated with the electron transport between CuWO4 and CoPi layers, as well as at the semiconductor–electrolyte interface. As a result, the Co-Pi/CuWO4 appears to enhance electron transfer both at the surface of CuWO4 and within the electrode. The Co-Pi was introduced via various methods, such as cyclic voltammetry (CV) [53] and photo-assisted electrodeposition (PED) [26], and both methods formed an amorphous coating layer on a CuWO4 particle. In addition, the Co-Pi/CuWO4 was more stable in electrolytes than bare CuWO4 because the Co-Pi layer could cover the defects of the electrode surface and inhibit the charge accumulation on the surface.
Co3O4 was introduced by C. M. Tian et al. [35] as an OER co-catalyst that increased the surface charge separation efficiency by reducing the onset potential of CuWO4 by ~100 mV. Through LSV and EIS analyses, however, they showed that too-thick co-catalyst layer created recombination sites where holes could not pass through, consequently leading to a decrease in PEC performance. The CuWO4 can make the p-n junction with a p-type sulfide electrocatalyst [44]. This junction induced the photogenerated electron and hole transfer to the CuWO4 bulk and p-type sulfide electrocatalyst, respectively. The photocurrent of the CuWO4/p-type sulfide increased and effectively reduced the onset potential by suppressing charge recombination. K. M. Nam et al. [55] compared Mn2O3 and MnPO co-catalysts in terms of enhancing the efficiency of a CuWO4 photoanode. The OER catalytic performance of MnPO was relatively higher when it was applied to CuWO4/WO3 heterojunction. Unfortunately, the onset potential remained the same, but the stability and efficiency of MnPO/CuWO4/WO3 increased to a reasonable extent. The NiWO4 has a distorted wolframite crystal structure similar to CuWO4 and can form a type II heterojunction with CuWO4 due to its higher energy CB and VB levels than those of CuWO4, reducing electron–hole recombination and improving IPCE [56]. P. Shadabipour et al. [57] deposited a well-known Ni- and Fe-based co-catalyst on CuWO4, but the OER efficiency of Ni0.75Fe0.25Oy/CuWO4 did not significantly increase because photoexcited holes did not migrate efficiently and the reaction occurred preferentially on the CuWO4 surface rather than on the co-catalyst.
The Ni-Pi synthesized on CuWO4 via a drop-casting method dramatically increased the stability of CuWO4 by interfering with charge accumulation [59]. In addition, it reduced charge recombination to greatly increase PEC performance. Moreover, the authors calculated the decay time of the charges in CuWO4 and Ni-Pi/CuWO4 through a time-resolved photoluminescence (TRPL)-based approach and found that Ni-Pi/CuWO4 had a decay time of 400 μs, which was five times longer than that of bare CuWO4. M. Davi et al. [58] fabricated a MnNCN/CuWO4 electrode through ED and drop-casting. Interestingly, the surface of MnNCN converted to MnPOx, forming a core–shell structure during water oxidation. However, the electrode showed a relatively small increase in activity, and the onset potential did not change. Noble metals were also introduced as co-catalysts for CuWO4. R. Salimi et al. [60] synthesized Ag-functionalized CuWO4/WO3 using a polyvinyl pyrrolidone (PVP)-assisted sol–gel (PSG) method. This photoanode showed a photocurrent of ~0.2 mA/cm2 at 1.23 VRHE in 0.1 M phosphate buffer due to Ag functionalization, which enhanced charge transfer and separation efficiency and mitigated charge recombination. An IrCo-Pi co-catalyst was deposited on CuWO4 NFs via the ED method to mitigate interfacial electron–hole recombination. The IrCo(9:1)-Pi/CuWO4 exhibited a photocurrent of ~0.54 mA/cm2 at 1.23 VRHE. Moreover, the co-catalyst-modified photoelectrode was more robust than bare CuWO4 [61]. In conclusion, the strategy of co-catalyst has been studied in a variety of ways and is still being actively explored. Unfortunately, despite these research efforts, an optimized co-catalyst material that gives outstanding performance has not yet been found, which requires continued research.

6. Conclusions

CuWO4 possesses a suitable band gap (2.3 eV) for solar water splitting, which positions it as a promising photoanode with a theoretical STH conversion efficiency exceeding 10%. Since its utilization as a photoanode began in 2011, the performance of CuWO4 photoanodes has been steadily improved through advanced synthesis methods and modification strategies. The most noticeable event in the development of CuWO4 photoanodes was the formation of a solid solution with Mo, thereby reducing the band gap to ~2.1 eV, which resulted in a significant enhancement in solar light harvesting and water oxidation efficiency. In addition, modification strategies proven effective for other well-developed photoelectrodes have been applied to the CuWO4 photoanodes. For example, forming a heterojunction with WO3 or SnO2 provides an ETL for effective charge transfer. Meanwhile, doping an Mn+ (n ≥ 3) ion into the Cu octahedral site or using gas treatment to generate oxygen vacancies increases the bulk charge-carrier density. Lastly, co-catalyst loading on CuWO4 offers multiple benefits, such as increasing photovoltage, improving surface hole transfer kinetics, and enhancing chemical stability. A fundamental understanding of the CuWO4 material and its working mechanism under PEC reaction conditions attained via DFT calculations and spectroscopy has facilitated these technical developments.
Despite these steady advancements, the performance of the CuWO4 photoanodes still lags far behind the theoretical expectation. This shortfall comes from intrinsically undesirable PEC properties of CuWO4, like low charge separation efficiency, short hole diffusion length, sluggish surface kinetics, and low bulk conductivity. This signifies the pressing need for the development of an optimal synthesis method and effective modification strategies for CuWO4 in the future. Additionally, strategies used to maximize efficiency by combining CuWO4 with another photoanode with a smaller band gap or integrating it with a high-efficiency photocathode for overall water-splitting systems could also be pursued. The enhancement of CuWO4 and its relevant variant will greatly enlarge research and application opportunities (e.g., the commercialization of PEC solar overall water splitting) for solar energy utilization technology, just as BiVO4 and Fe2O3 did for this field. For instance, future research, such as the valorization of organic molecular via partial oxidation, can be conducted using CuWO4 and other tungstates, which will have different and interesting chemistry compared to those of other materials. Also, detailed analysis of the optoelectronic property of CuWO4 will provide much information that is useful for solving the aforementioned problems realized in the current state of development. After this problem has been dealt with, CuWO4 will surely be able to take the place of BiVO4 due to its similarity in terms of operating conditions (near neutral) and elemental composition (ternary metal oxide), which will benefit the community studying photoelectrochemical cells in the same way as previous go-to materials. More specifically, owing to its appropriate band gap that can achieve STH above 10%, it can be possible candidate for a practical solar overall water-splitting PEC cell.

Author Contributions

Conceptualization, J.U.L. and J.H.K.; investigation, J.U.L.; data curation, J.H.K.; writing—original draft preparation, J.U.L. and J.H.K.; writing—review and editing, J.S.L.; supervision, J.H.K. and J.S.L.; project administration, J.H.K.; funding acquisition, J.S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Climate Change Response Project (NRF2019M1A2A2065612) and the Brainlink Project (NRF2022H1D3A3A01081140), funded by the Ministry of Science and ICT of Korea via National Research Foundation, and by research funds from Hanhwa Solutions Chemicals (2.220990.01) and UNIST (1.190013.01). This work was also supported by the Institute for Basic Science (IBSR019-D1). J.H.K. gratefully acknowledges the support from the Basic Science Research Program funded by the Ministry of Education (NRF2021R1A6A3A14039651, NRF-2019R1A6A3A01096197).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Davis, S.J.; Lewis, N.S.; Shaner, M.; Aggarwal, S.; Arent, D.; Azevedo, I.L.; Benson, S.M.; Bradley, T.; Brouwer, J.; Chiang, Y.-M.; et al. Net-zero emissions energy systems. Science 2018, 360, eaas9793. [Google Scholar] [CrossRef]
  2. Lewis, N.S. Research opportunities to advance solar energy utilization. Science 2016, 351, aad1920. [Google Scholar] [CrossRef]
  3. Jacobsson, T.J.; Fjällström, V.; Edoff, M.; Edvinsson, T. Sustainable solar hydrogen production: From photoelectrochemical cells to PV-electrolyzers and back again. Energy Environ. Sci. 2014, 7, 2056–2070. [Google Scholar] [CrossRef]
  4. Yao, T.; An, X.; Han, H.; Chen, J.Q.; Li, C. Photoelectrocatalytic materials for solar water splitting. Adv. Energy Mater. 2018, 8, 1800210. [Google Scholar] [CrossRef]
  5. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  6. Mesa, C.A.; Francas, L.; Yang, K.R.; Garrido-Barros, P.; Pastor, E.; Ma, Y.; Kafizas, A.; Rosser, T.E.; Mayer, M.T.; Reisner, E.; et al. Multihole water oxidation catalysis on haematite photoanodes revealed by operando spectroelectrochemistry and DFT. Nat. Chem. 2020, 12, 82–89. [Google Scholar] [CrossRef] [PubMed]
  7. Sivula, K.; Van De Krol, R. Semiconducting materials for photoelectrochemical energy conversion. Nat. Rev. Mater. 2016, 1, 15010. [Google Scholar] [CrossRef]
  8. Walter, M.G.; Warren, E.L.; McKone, J.R.; Boettcher, S.W.; Mi, Q.; Santori, E.A.; Lewis, N.S. Solar Water Splitting Cells. Chem. Rev. 2010, 110, 6446–6473. [Google Scholar] [CrossRef]
  9. Wang, Q.; Domen, K. Particulate Photocatalysts for Light-Driven Water Splitting: Mechanisms, Challenges, and Design Strategies. Chem. Rev. 2020, 120, 919–985. [Google Scholar] [CrossRef]
  10. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  11. Takata, T.; Jiang, J.; Sakata, Y.; Nakabayashi, M.; Shibata, N.; Nandal, V.; Seki, K.; Hisatomi, T.; Domen, K. Photocatalytic water splitting with a quantum efficiency of almost unity. Nature 2020, 581, 411–414. [Google Scholar] [CrossRef] [PubMed]
  12. Sarnowska, M.; Bienkowski, K.; Barczuk, P.J.; Solarska, R.; Augustynski, J. Highly Efficient and Stable Solar Water Splitting at (Na)WO3 Photoanodes in Acidic Electrolyte Assisted by Non-Noble Metal Oxygen Evolution Catalyst. Adv. Energy Mater. 2016, 6, 1600526. [Google Scholar] [CrossRef]
  13. Park, Y.; McDonald, K.J.; Choi, K.-S.J.C.S.R. Progress in bismuth vanadate photoanodes for use in solar water oxidation. Chem. Soc. Rev. 2013, 42, 2321–2337. [Google Scholar] [CrossRef] [PubMed]
  14. Hill, J.C.; Choi, K.-S. Synthesis and characterization of high surface area CuWO4 and Bi2WO6 electrodes for use as photoanodes for solar water oxidation. J. Mater. Chem. A 2013, 1, 5006–5014. [Google Scholar] [CrossRef]
  15. Bathula, B.; Eadi, S.B.; Lee, H.-D.; Yoo, K. ZnWO4 nanorod-colloidal SnO2 quantum dots core@shell heterostructures: Efficient solar-light-driven photocatalytic degradation of tetracycline. Environ. Res. 2023, 228, 115851. [Google Scholar] [CrossRef]
  16. Babu, B.; Peera, S.G.; Yoo, K. Fabrication of ZnWO4-SnO2 Core-Shell Nanorods for Enhanced Solar Light-Driven Photoelectrochemical Performance. Inorganics 2023, 11, 213. [Google Scholar] [CrossRef]
  17. Lhermitte, C.R.; Bartlett, B.M. Advancing the chemistry of CuWO4 for photoelectrochemical water oxidation. Acc. Chem. Res. 2016, 49, 1121–1129. [Google Scholar] [CrossRef]
  18. Tian, C.M.; Jiang, M.; Tang, D.; Qiao, L.; Xiao, H.Y.; Oropeza, F.; Hofmann, J.P.; Hensen, E.; Tadich, A.; Li, W.; et al. Elucidating the electronic structure of CuWO4 thin films for enhanced photoelectrochemical water splitting. J. Mater. Chem. A 2019, 7, 11895–11907. [Google Scholar] [CrossRef]
  19. Souza, E.L.S.; Sczancoski, J.C.; Nogueira, I.C.; Almeida, M.A.P.; Orlandi, M.O.; Li, M.S.; Luz, R.A.S.; Filho, M.G.R.; Longo, E.; Cavalcante, L.S. Structural evolution, growth mechanism and photoluminescence properties of CuWO4 nanocrystals. Ultrason. Sonochem. 2017, 38, 256–270. [Google Scholar] [CrossRef]
  20. Lalić, M.V.; Popović, Z.S.; Vukajlović, F.R. Ab initio study of electronic, magnetic and optical properties of CuWO4 tungstate. Comput. Mater. Sci. 2011, 50, 1179–1186. [Google Scholar] [CrossRef]
  21. Yourey, J.E.; Bartlett, B.M. Electrochemical deposition and photoelectrochemistry of CuWO4, a promising photoanode for water oxidation. J. Mater.Chem. 2011, 21, 7651–7660. [Google Scholar] [CrossRef]
  22. Lalić, M.V.; Popović, Z.S.; Vukajlović, F.R. Electronic structure and optical properties of CuWO4: An ab initio study. Comput. Mater. Sci. 2012, 63, 163–167. [Google Scholar] [CrossRef]
  23. Peeters, D.; Mendoza Reyes, O.; Mai, L.; Sadlo, A.; Cwik, S.; Rogalla, D.; Becker, H.W.; Schütz, H.M.; Hirst, J.; Müller, S.; et al. CVD-grown copper tungstate thin films for solar water splitting. J. Mater. Chem. A 2018, 6, 10206–10216. [Google Scholar] [CrossRef]
  24. Hu, D.; Diao, P.; Xu, D.; Xia, M.; Gu, Y.; Wu, Q.; Li, C.; Yang, S. Copper (II) tungstate nanoflake array films: Sacrificial template synthesis, hydrogen treatment, and their application as photoanodes in solar water splitting. Nanoscale 2016, 8, 5892–5901. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, Z.; Xiao, C.; Li, S. CuWO4 films grown via seeding-hydrothermal method for photoelectrochemical water oxidation. Mater. Lett. 2018, 232, 25–28. [Google Scholar] [CrossRef]
  26. Lee, J.U.; Kim, J.H.; Kang, K.; Shin, Y.S.; Kim, J.Y.; Kim, J.H.; Lee, J.S. Bulk and surface modified polycrystalline CuWO4 films for photoelectrochemical water oxidation. Renew. Energy 2023, 203, 779–787. [Google Scholar] [CrossRef]
  27. dos Santos Costa, M.J.; Lima, A.E.B.; Ribeiro, E.P.; dos Santos Costa, G.; Longo, E.; da Luz, G.E.; Cavalcante, L.S.; da Silva Santos , R. Transition metal tungstates AWO4 (A2+ = Fe, Co, Ni, and Cu) thin films and their photoelectrochemical behavior as photoanode for photocatalytic applications. J. Appl. Electrochem. 2023, 53, 1349–1367. [Google Scholar] [CrossRef]
  28. Duan, X.; Xu, C.; El Nahrawy, A.M.; Chen, J.; Zhu, Z.; Wang, J.; Liang, Q.; Cao, F. Ultrasonic Spray Pyrolysis-Assisted Fabrication of Ultrathin CuWO4 Films with Improved Photoelectrochemical Performance. ChemNanoMat 2022, 8, e202100419. [Google Scholar] [CrossRef]
  29. Gaillard, N.; Chang, Y.; DeAngelis, A.; Higgins, S.; Braun, A. A nanocomposite photoelectrode made of 2.2 eV band gap copper tungstate (CuWO4) and multi-wall carbon nanotubes for solar-assisted water splitting. Int. J. Hydrogen Energy 2013, 38, 3166–3176. [Google Scholar] [CrossRef]
  30. Gao, Y.; Zandi, O.; Hamann, T.W. Atomic layer stack deposition-annealing synthesis of CuWO4. J. Mater. Chem. A 2016, 4, 2826–2830. [Google Scholar] [CrossRef]
  31. Gonzalez, C.M.; Dunford, J.L.; Du, X.; Post, M.L. Characterization of carrier states in CuWO4 thin-films at elevated temperatures using conductometric analysis. J. Solid State Chem. 2013, 201, 35–40. [Google Scholar] [CrossRef]
  32. Hrubantova, A.; Hippler, R.; Wulff, H.; Cada, M.; Gedeon, O.; Jiricek, P.; Houdkova, J.; Olejnicek, J.; Nepomniashchaia, N.; Helm, C.A.; et al. Copper tungsten oxide (CuxWOy) thin films for optical and photoelectrochemical applications deposited by reactive high power impulse magnetron co-sputtering. J. Appl. Phys. 2022, 132, 215301. [Google Scholar] [CrossRef]
  33. Yourey, J.E.; Pyper, K.J.; Kurtz, J.B.; Bartlett, B.M. Chemical stability of CuWO4 for photoelectrochemical water oxidation. J. Phys. Chem. C 2013, 117, 8708–8718. [Google Scholar] [CrossRef]
  34. Wu, Z.; Zhao, Z.; Cheung, G.; Doughty, R.M.; Ballestas-Barrientos, A.R.; Hirmez, B.; Han, R.; Maschmeyer, T.; Osterloh, F.E. Role of Surface States in Photocatalytic Oxygen Evolution with CuWO4 Particles. J. Electrochem. Soc. 2019, 166, H3014. [Google Scholar] [CrossRef]
  35. Gao, Y.; Hamann, T.W. Elucidation of CuWO4 surface states during photoelectrochemical water oxidation. J. Phys. Chem. Lett. 2017, 8, 2700–2704. [Google Scholar] [CrossRef]
  36. Mayer, M.T. Photovoltage at semiconductor–electrolyte junctions. Curr. Opin. Electrochem. 2017, 2, 104–110. [Google Scholar] [CrossRef]
  37. Pyper, K.J.; Yourey, J.E.; Bartlett, B.M. Reactivity of CuWO4 in photoelectrochemical water oxidation is dictated by a midgap electronic state. J. Phys. Chem. C 2013, 117, 24726–24732. [Google Scholar] [CrossRef]
  38. Gao, Y.; Hamann, T.W. Quantitative hole collection for photoelectrochemical water oxidation with CuWO4. Chem. Commun. 2017, 53, 1285–1288. [Google Scholar] [CrossRef]
  39. Rodríguez-Gutiérrez, I.; Djatoubai, E.; Rodríguez-Pérez, M.; Su, J.; Rodríguez-Gattorno, G.; Vayssieres, L.; Oskam, G. Photoelectrochemical water oxidation at FTO|WO3@CuWO4 and FTO|WO3@CuWO4|BiVO4 heterojunction systems: An IMPS analysis. Electrochim. Acta 2019, 308, 317–327. [Google Scholar] [CrossRef]
  40. Peng, B.; Li, C.; Yue, C.; Diao, P. Sacrificial Template Synthesis of Copper Tungstate: Influence of Preparing Conditions on Morphology and Photoactivity. Int. J. Electrochem. Sci. 2019, 14, 2574–2588. [Google Scholar] [CrossRef]
  41. Bohra, D.; Smith, W.A. Improved charge separation via Fe-doping of copper tungstate photoanodes. Phys. Chem. Chem. Phys. 2015, 17, 9857–9866. [Google Scholar] [CrossRef] [PubMed]
  42. Li, C.; Diao, P. Fluorine doped copper tungsten nanoflakes with enhanced charge separation for efficient photoelectrochemical water oxidation. Electrochim. Acta 2020, 352, 136471. [Google Scholar] [CrossRef]
  43. González-Poggini, S.; Sánchez, B.; Colet-Lagrille, M. Enhanced Photoelectrochemical Activity of CuWO4 Photoanode by Yttrium Doping. J. Electrochem. Soc. 2023, 170, 066512. [Google Scholar] [CrossRef]
  44. Ikeue, K.; Ueno, T. Photoelectrochemical water oxidation properties of Mo-doped CuWO4: Effect of p-type sulfide loading and annealing. Mater. Lett. 2023, 348, 134690. [Google Scholar] [CrossRef]
  45. Hill, J.C.; Ping, Y.; Galli, G.A.; Choi, K.-S. Synthesis, photoelectrochemical properties, and first principles study of n-type CuW1−xMoxO4 electrodes showing enhanced visible light absorption. Energy Environ. Sci 2013, 6, 2440–2446. [Google Scholar] [CrossRef]
  46. Liang, Q.; Guo, Y.; Zhang, N.; Qian, Q.; Hu, Y.; Hu, J.; Li, Z.; Zou, Z. Improved water-splitting performances of CuW1−xMoxO4 photoanodes synthesized by spray pyrolysis. Sci. China Mater. 2018, 61, 1297–1304. [Google Scholar] [CrossRef]
  47. Yang, J.; Li, C.; Diao, P. Molybdenum doped CuWO4 nanoflake array films as an efficient photoanode for solar water splitting. Electrochim. Acta 2019, 308, 195–205. [Google Scholar] [CrossRef]
  48. Polo, A.; Nomellini, C.; Grigioni, I.; Dozzi, M.V.; Selli, E. Effective Visible Light Exploitation by Copper Molybdo-Tungstate Photoanodes. ACS Appl. Energy Mater. 2020, 3, 6956–6964. [Google Scholar] [CrossRef]
  49. Wang, D.; Bassi, P.S.; Qi, H.; Zhao, X.; Gurudayal; Wong, L.H.; Xu, R.; Sritharan, T.; Chen, Z. Improved Charge Separation in WO3/CuWO4 Composite Photoanodes for Photoelectrochemical Water Oxidation. Materials 2016, 9, 348. [Google Scholar] [CrossRef]
  50. Wang, T.; Fan, X.; Gao, B.; Jiang, C.; Li, Y.; Li, P.; Zhang, S.; Huang, X.; He, J. Self-Assembled Urchin-like CuWO4/WO3 Heterojunction Nanoarrays as Photoanodes for Photoelectrochemical Water Splitting. ChemElectroChem 2021, 8, 125–134. [Google Scholar] [CrossRef]
  51. Ye, W.; Chen, F.; Zhao, F.; Han, N.; Li, Y. CuWO4 nanoflake array-based single-junction and heterojunction photoanodes for photoelectrochemical water oxidation. ACS Appl. Mater. Interfaces 2016, 8, 9211–9217. [Google Scholar] [CrossRef]
  52. Chen, S.; Hossain, M.N.; Chen, A. Significant enhancement of the photoelectrochemical activity of CuWO4 by using a cobalt phosphate nanoscale thin film. ChemElectroChem 2018, 5, 523–530. [Google Scholar] [CrossRef]
  53. Li, C.; Guo, B.; Peng, B.; Yue, C.; Diao, P. Copper Tungstate (CuWO4) Nanoflakes Coupled with Cobalt Phosphate (Co-Pi) for Effective Photoelectrochemical Water Splitting. Int. J. Electrochem. Sci. 2019, 14, 9017–9029. [Google Scholar] [CrossRef]
  54. Rosa, W.S.; Rabelo, L.G.; Tiveron Zampaulo, L.G.; Gonçalves, R.V. Ternary Oxide CuWO4/BiVO4/FeCoOx Films for Photoelectrochemical Water Oxidation: Insights into the Electronic Structure and Interfacial Band Alignment. ACS Appl. Mater. Interfaces 2022, 14, 22858–22869. [Google Scholar] [CrossRef] [PubMed]
  55. Nam, K.M.; Cheon, E.A.; Shin, W.J.; Bard, A.J. Improved Photoelectrochemical Water Oxidation by the WO3/CuWO4 Composite with a Manganese Phosphate Electrocatalyst. Langmuir 2015, 31, 10897–10903. [Google Scholar] [CrossRef] [PubMed]
  56. Fan, L.; Sunarso, J.; Zhang, X.; Xiong, X.; He, L.; Luo, L.; Wang, F.; Fan, Z.; Wu, C.; Han, D.; et al. Regulating the hole transfer from CuWO4 photoanode to NiWO4 electrocatalyst for enhanced water oxidation performance. Int. J. Hydrogen Energy 2022, 47, 20153–20165. [Google Scholar] [CrossRef]
  57. Shadabipour, P.; Raithel, A.L.; Hamann, T.W. Charge-Carrier Dynamics at the CuWO4/Electrocatalyst Interface for Photoelectrochemical Water Oxidation. ACS Appl. Mater. Interfaces 2020, 12, 50592–50599. [Google Scholar] [CrossRef]
  58. Davi, M.; Mann, M.; Ma, Z.; Schrader, F.; Drichel, A.; Budnyk, S.; Rokicinska, A.; Kustrowski, P.; Dronskowski, R.; Slabon, A. An MnNCN-Derived Electrocatalyst for CuWO4 Photoanodes. Langmuir 2018, 34, 3845–3852. [Google Scholar] [CrossRef]
  59. Xiong, X.; Fan, L.; Chen, G.; Wang, Y.; Wu, C.; Chen, D.; Lin, Y.; Li, T.; Fu, S.; Ren, S. Boosting water oxidation performance of CuWO4 photoanode by surface modification of nickel phosphate. Electrochim. Acta 2019, 328, 135125. [Google Scholar] [CrossRef]
  60. Salimi, R.; Sabbagh Alvani, A.A.; Mei, B.T.; Naseri, N.; Du, S.F.; Mul, G. Ag-Functionalized CuWO4/WO3 nanocomposites for solar water splitting. New J. Chem. 2019, 43, 2196–2203. [Google Scholar] [CrossRef]
  61. Li, C.; Diao, P. Boosting the Activity and Stability of Copper Tungsten Nanoflakes toward Solar Water Oxidation by Iridium-Cobalt Phosphates Modification. Catalysts 2020, 10, 913. [Google Scholar] [CrossRef]
  62. Tang, Y.; Rong, N.; Liu, F.; Chu, M.; Dong, H.; Zhang, Y.; Xiao, P. Enhancement of the photoelectrochemical performance of CuWO4 films for water splitting by hydrogen treatment. Appl. Surf. Sci 2016, 361, 133–140. [Google Scholar] [CrossRef]
  63. Ma, Z.; Linnenberg, O.; Rokicinska, A.; Kustrowski, P.; Slabon, A. Augmenting the Photocurrent of CuWO4 Photoanodes by Heat Treatment in the Nitrogen Atmosphere. J. Phys. Chem. C 2018, 122, 19281–19288. [Google Scholar] [CrossRef]
  64. Wang, K.; Chen, L.; Liu, X.; Li, J.; Liu, Y.; Liu, M.; Qiu, X.; Li, W. Gradient surficial forward Ni and interior reversed Mo-doped CuWO4 films for enhanced photoelectrochemical water splitting. Chem. Eng. J. 2023, 471, 144730. [Google Scholar] [CrossRef]
  65. Kim, J.H.; Lee, J.S. Elaborately modified BiVO4 photoanodes for solar water splitting. Adv. Mater. 2019, 31, 1806938. [Google Scholar] [CrossRef] [PubMed]
  66. Sivula, K.; Formal, F.L.; Grätzel, M. WO3-Fe2O3 Photoanodes for Water Splitting: A Host Scaffold, Guest Absorber Approach. Chem. Mater. 2009, 21, 2862–2867. [Google Scholar] [CrossRef]
  67. Shi, X.; Zhang, K.; Shin, K.; Ma, M.; Kwon, J.; Choi, I.T.; Kim, J.K.; Kim, H.K.; Wang, D.H.; Park, J.H. Unassisted photoelectrochemical water splitting beyond 5.7% solar-to-hydrogen conversion efficiency by a wireless monolithic photoanode/dye-sensitised solar cell tandem device. Nano Energy 2015, 13, 182–191. [Google Scholar] [CrossRef]
  68. Bignozzi, C.A.; Caramori, S.; Cristino, V.; Argazzi, R.; Meda, L.; Tacca, A. Nanostructured photoelectrodes based on WO3: Applications to photooxidation of aqueous electrolytes. Chem. Soc. Rev. 2013, 42, 2228–2246. [Google Scholar] [CrossRef]
  69. Wang, D.; Li, R.; Zhu, J.; Shi, J.; Han, J.; Zong, X.; Li, C. Photocatalytic Water Oxidation on BiVO4 with the Electrocatalyst as an Oxidation Cocatalyst: Essential Relations between Electrocatalyst and Photocatalyst. J. Phys. Chem. C 2012, 116, 5082–5089. [Google Scholar] [CrossRef]
  70. Nellist, M.R.; Qiu, J.; Laskowski, F.A.; Toma, F.M.; Boettcher, S.W. Potential-sensing electrochemical AFM shows CoPi as a hole collector and oxygen evolution catalyst on BiVO4 water-splitting photoanodes. ACS Energy Lett. 2018, 3, 2286–2291. [Google Scholar] [CrossRef]
Figure 1. (a) Crystal and electronic structure of photoactive CuWO4. (b) Crystal structure of triclinic wolframite CuWO4 and its calculated projected density of state using hybrid DFT, reproduced from ref. [18]. (c) UV−vis spectrum of WO3 and CuWO4 thin film (inset 1: picture image of CuWO4/FTO) (inset 2: Tauc plot of WO3 and CuWO4 with trend line for optical direct band gap), (d) experimentally determined electronic structure of WO3 and CuWO4 thin film made via electrodeposition and thermal conversion on FTO, reproduced from ref. [21].
Figure 1. (a) Crystal and electronic structure of photoactive CuWO4. (b) Crystal structure of triclinic wolframite CuWO4 and its calculated projected density of state using hybrid DFT, reproduced from ref. [18]. (c) UV−vis spectrum of WO3 and CuWO4 thin film (inset 1: picture image of CuWO4/FTO) (inset 2: Tauc plot of WO3 and CuWO4 with trend line for optical direct band gap), (d) experimentally determined electronic structure of WO3 and CuWO4 thin film made via electrodeposition and thermal conversion on FTO, reproduced from ref. [21].
Catalysts 13 01408 g001
Figure 3. Polycrystalline CuWO4 film synthesized via a sol–gel coating method: (a) schematics of CuWO4 film coatings on FTO using 1.0 M Cu and W precursors, (b) optical images of the films, (c) UV-vis spectrum and Tauc plot (insets), (d) XRD pattern (blue square indicates pattern of FTO, F doped SnO2 as substrate), and (e) a SEM image of CuWO4/FTO. Reproduced from ref. [26].
Figure 3. Polycrystalline CuWO4 film synthesized via a sol–gel coating method: (a) schematics of CuWO4 film coatings on FTO using 1.0 M Cu and W precursors, (b) optical images of the films, (c) UV-vis spectrum and Tauc plot (insets), (d) XRD pattern (blue square indicates pattern of FTO, F doped SnO2 as substrate), and (e) a SEM image of CuWO4/FTO. Reproduced from ref. [26].
Catalysts 13 01408 g003
Figure 5. Formation of semiconductor–liquid junction for CuWO4 in an aqueous electrolyte. (a) Proposed physical model for charge-carrier pathways in CuWO4: excitation, mid-gap state trapping, and charge–transfer reactions to solution. W represents the depletion width, and the red arrows indicate charge–transfer processes. (b) Nyquist plot for EIS data measured in 0.5 M KBi and 0.2 M KCl at pH 7, under one sun illumination, and 0.96 VRHE. Solid circles in Nyquist plot is marked with frequency of measurement applied to it. Reproduced from ref. [37]. (c) IV curves of CuWO4/FTO in pH 9.0 KBi buffer under one sun illumination. (d) Css obtained via fitting EIS under 0.1 sun (blue triangles) and 1 sun (red circles) illumination in H2O electrolyte, fitted with equivalent circuit for EIS data with two capacitive features (inset). (e) Mott–Schottky plots of CuWO4 immersed in H2O electrolyte, measured in the dark (open squares), under 0.1 sun (blue triangles) and 1 sun (red circles) illumination, fitted with a Randel circuit (inset). Reproduced from ref. [35].
Figure 5. Formation of semiconductor–liquid junction for CuWO4 in an aqueous electrolyte. (a) Proposed physical model for charge-carrier pathways in CuWO4: excitation, mid-gap state trapping, and charge–transfer reactions to solution. W represents the depletion width, and the red arrows indicate charge–transfer processes. (b) Nyquist plot for EIS data measured in 0.5 M KBi and 0.2 M KCl at pH 7, under one sun illumination, and 0.96 VRHE. Solid circles in Nyquist plot is marked with frequency of measurement applied to it. Reproduced from ref. [37]. (c) IV curves of CuWO4/FTO in pH 9.0 KBi buffer under one sun illumination. (d) Css obtained via fitting EIS under 0.1 sun (blue triangles) and 1 sun (red circles) illumination in H2O electrolyte, fitted with equivalent circuit for EIS data with two capacitive features (inset). (e) Mott–Schottky plots of CuWO4 immersed in H2O electrolyte, measured in the dark (open squares), under 0.1 sun (blue triangles) and 1 sun (red circles) illumination, fitted with a Randel circuit (inset). Reproduced from ref. [35].
Catalysts 13 01408 g005
Figure 8. Effects of CuWO4/WO3 type II heterojunction. (a) Schematics of the formation of a CuWO4/WO3 heterojunction film and (b) its SEM image. (c) IV curves (solid line: under illumination, dotted line: under dark) and (d) IPCE at an applied bias of 1.20 VRHE for CuWO4, WO3, and CuWO4/WO3 photoanodes in 0.5 M Na2SO4 under one sun. (e) Tentative energy diagram of the CuWO4/WO3 type II heterojunction based on UV-vis spectra and the Mott–Schottky plot. Reproduced from ref. [49].
Figure 8. Effects of CuWO4/WO3 type II heterojunction. (a) Schematics of the formation of a CuWO4/WO3 heterojunction film and (b) its SEM image. (c) IV curves (solid line: under illumination, dotted line: under dark) and (d) IPCE at an applied bias of 1.20 VRHE for CuWO4, WO3, and CuWO4/WO3 photoanodes in 0.5 M Na2SO4 under one sun. (e) Tentative energy diagram of the CuWO4/WO3 type II heterojunction based on UV-vis spectra and the Mott–Schottky plot. Reproduced from ref. [49].
Catalysts 13 01408 g008
Figure 9. Efficacy of Co-Pi for CuWO4 photoanode. (a) Working principle of CuWO4 photoanodes with and without a Co-Pi co-catalyst. (b) IV curves and (c) Nyquist plots at an applied bias of 1.23 VRHE in 0.1 M KPi, pH 7.0 under simulated one sun. Red and blue solid line are for from equivalent circuit. Reproduced from ref. [52].
Figure 9. Efficacy of Co-Pi for CuWO4 photoanode. (a) Working principle of CuWO4 photoanodes with and without a Co-Pi co-catalyst. (b) IV curves and (c) Nyquist plots at an applied bias of 1.23 VRHE in 0.1 M KPi, pH 7.0 under simulated one sun. Red and blue solid line are for from equivalent circuit. Reproduced from ref. [52].
Catalysts 13 01408 g009
Table 1. Summary of progress made for CuWO4 as a photoanode material.
Table 1. Summary of progress made for CuWO4 as a photoanode material.
StrategiesReported SubstancesTypical Effects
Synthesis methodElectrodeposition [14,21]
Thermal conversion [39,40]
Sol–gel method [26,29,35]
Hydrothermal [25]
Electrodeposition (Jph 0.2 mA/cm2 at 1.23 VRHE) [21]
ALD [30]Thermal conversion (Jph 0.33 mA/cm2 at 1.23 VRHE) [24]
CVD [23]
Ultrasonic spray pyrolysis [28]
Impulse magnetron co-sputtering [32]
PLD [31]
Sol–gel (Jph 0.5mA/cm2 at 1.23 VRHE, 1.0 mA/cm2 at 1.23 VRHE for SA oxidation) [18]
Sol–gel (Jph 0.07mA/cm2 at 1.23 VRHE, 0.15 mA/cm2 at 1.23 VRHE for SA oxidation) [26]
DopingFe3+ [41]0.3% Fe:CuWO4/FTO, ~1.5 times greater ηbulk (Jph 0.5 mA/cm2 at 1.23 VRHE for SA oxidation) [41]
F [42]
Y3+ [43]
Mo6+ [44]
Mo solid solutionMo6+ [45,46,47,48]Red shift of photoresponse (from 550 nm to 600 nm) [45]
CuW0.35Mo0.65O4/FTO (Jph 1.0 mA/cm2 at 1.23 VRHE for SA oxidation) [46]
Heterojunction and electron transfer layerWO3 (flat) [49]
WO3 (nanorod) [39]
WO3 (urchin like) [50]
SnO2 [26]
BiVO4 [51]
CuWO4/flat WO3/FTO. ~4 times increment (Jph 0.55 mA/cm2 at 1.23 VRHE) [49]
ElectrocatalystCo-Pi [26,52,53]
Co3O4 [18]
FeCoOx [54]
MnPO4 [55]
NiWO4 [56]
NiFeOx [57]
P-type sulfide (MoS2, NbS2, NiSx) [44]
MnNCN [58]
Ni-Pi [59]
Ag [60]
Co-Pi/CuWO4/FTO, 30% increment (Jph 0.4 mA/cm2 at 0.6 VAg/AgCl) [52]
IrCo-Pi [61]
Post-treatmentH2 treatment [62] H2 treatment (300 °C), 3-time increment (Jph 0.6 mA/cm2 at 1.01 VAg/AgCl) [62]
SA: sacrificial agent. ηbulk = bulk charge separation efficiency.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lee, J.U.; Kim, J.H.; Lee, J.S. Emergent CuWO4 Photoanodes for Solar Fuel Production: Recent Progress and Perspectives. Catalysts 2023, 13, 1408. https://doi.org/10.3390/catal13111408

AMA Style

Lee JU, Kim JH, Lee JS. Emergent CuWO4 Photoanodes for Solar Fuel Production: Recent Progress and Perspectives. Catalysts. 2023; 13(11):1408. https://doi.org/10.3390/catal13111408

Chicago/Turabian Style

Lee, Jin Uk, Jin Hyun Kim, and Jae Sung Lee. 2023. "Emergent CuWO4 Photoanodes for Solar Fuel Production: Recent Progress and Perspectives" Catalysts 13, no. 11: 1408. https://doi.org/10.3390/catal13111408

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop