Next Article in Journal
Asymmetric Hydrogenation of 1-aryl substituted-3,4-Dihydroisoquinolines with Iridium Catalysts Bearing Different Phosphorus-Based Ligands
Previous Article in Journal
Ordered Mesoporous Carbon as a Support of Pd Catalysts for CO2 Electrochemical Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosting the Activity and Stability of Copper Tungsten Nanoflakes toward Solar Water Oxidation by Iridium-Cobalt Phosphates Modification

Key Laboratory of Aerospace Materials and Performance (Ministry of Education), School of Materials Science and Engineering, Beihang University, Beijing 100191, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(8), 913; https://doi.org/10.3390/catal10080913
Submission received: 26 July 2020 / Revised: 6 August 2020 / Accepted: 7 August 2020 / Published: 10 August 2020
(This article belongs to the Section Photocatalysis)

Abstract

:
Severe interfacial electron–hole recombination greatly limits the performance of CuWO4 photoanode towards the photoelectrochemical (PEC) oxygen evolution reaction (OER). Surface modification with an OER cocatalyst can reduce electron–hole recombination and thus improve the PEC OER performance of CuWO4. Herein, we coupled CuWO4 nanoflakes (NFs) with Iridium–cobalt phosphates (IrCo-Pi) and greatly improved the photoactivity of CuWO4. The optimized photocurrent density for CuWO4/IrCo-Pi at 1.23 V vs. reversible hydrogen electrode (RHE) rose to 0.54 mA∙cm−2, a ca. 70% increase over that of bare CuWO4 (0.32 mA∙cm−2). Such improved photoactivity was attributed to the enhanced hole collection efficiency, which resulted from the reduced charge-transfer resistance via IrCo-Pi modification. Moreover, the as-deposited IrCo-Pi layer well coated the inner CuWO4 NFs and effectively prevented the photoinduced corrosion of CuWO4 in neutral potassium phosphate (KPi) buffer solution, eventually leading to a superior stability over the bare CuWO4. The facile preparation of IrCo-Pi and its great improvement in the photoactivity make it possible to design an efficient CuWO4/cocatalyst system towards PEC water oxidation.

Graphical Abstract

1. Introduction

Photoelectrochemical (PEC) water splitting is a green way to convert solar power into storable chemical fuels and thus addresses the rising demand for renewable energy [1,2]. The overall course includes both the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER), which occur on the photoanode and photocathode, respectively [1]. Between them, the OER, with a more sluggish kinetic character, is generally considered as the rate-limiting step [3]. The solar-to-chemical conversion efficiency for PEC OER undoubtedly depends on the employed photoanode material. Beginning with TiO2 as an efficient photoanode in 1972 [4], some other binary transition-metal oxides, such as ZnO [5,6], WO3 [7,8,9,10], and Fe2O3 [11,12,13], have been found active toward PEC OER. Among them, WO3 has attracted intensive attention due to its nontoxic composition, high hole mobility property (~10 cm2∙V−1∙s−1) [14], and long hole diffusion length (~150 nm) [14]. However, the drawbacks of WO3, such as relatively large bandgap (ca. 2.7 eV) [7,8,9] and narrow pH range for stable work (pH < 4), restrict its large-scale application [15]. To address such limitations, it is feasible to associate another transition-metal oxide with WO3 to generate a WO3-based ternary oxide, as the introduced metallic component can regulate the bandgap structure and bonding state of WO3 without damaging its main merits [16,17,18,19].
In the family of WO3-based ternary oxides, CuWO4, as a promising photoanode material, has played an increasingly important role in PEC water oxidation over recent years [19,20,21,22,23,24,25,26,27,28,29]. Due to the contribution of Cu(3d) orbitals to the valance band, the bandgap value of CuWO4 is reduced by ca. 0.4 eV compared to WO3, resulting in a red-shift of the optical-absorption edge from 475 nm to 550 nm [19]. While WO3 suffers severe corrosion in a pH > 4 aqueous solution due to its acidic oxide nature [15], CuWO4 exhibits excellent stability over a wider pH range (≤9.5) [19,20,21,23,24,25,28,29]. The strong covalency of Cu-O binding in CuWO4 inhibits the generation of soluble tungstate [19]. Despite these merits, CuWO4 still shows relatively lower photoactivity than the common photoanodes, including its binary counterpart WO3. Much effort has been directed to reveal the limiting factors in water oxidation for CuWO4 and then improve its solar-to-chemical conversion efficiency [24,25,26,27,28,29].
As reported previously, CuWO4 suffers from severe electron–hole recombination at the CuWO4/solution interface owing to the slow OER kinetics [21,22,27], thus limiting its PEC OER performance. However, in the presence of hole scavengers, such as CH3OH [19], H2O2 [25], or Na2SO3 [29], CuWO4 will show a much higher photoactivity, suggesting huge room for improvement in its PEC OER activity. Generally, surface modification with an OER cocatalyst is a valid method to alleviate the interfacial electron–hole recombination, since the cocatalyst can efficiently capture the photogenerated holes and then accelerate the PEC OER [5,8,30,31]. Nevertheless, Hamann [32] found that decorating CuWO4 with common cocatalysts usually led to comparable or even worse PEC performance. To date, the two previous studies concerning cocatalyst-modified CuWO4 reported that only mild enhancement in the photoactivity was achieved via manganese carbodiimide (MnNCN) [26] and manganese phosphate (Mn-Pi) [27] modification. Highly-enhanced performance can be potentially achieved via noble-metal-based OER catalysts, since they generally outperform the noble-metal-free counterparts whether as electrocatalysts alone or as cocatalysts combined with semiconductors [33]. However, the high cost prohibits their further application in industry.
Ir-Pi, with its noble-metal composition, shows the highest OER activity in the family of transition metal phosphate (TMP) [34]. Following the merit of TMP, IrPi is easy to prepare and possesses a superior stability in neutral potassium phosphate buffer solution (KPi) [34], both favoring its widespread applicability in OER. In our previous work, we doped the low-cost Co into pure Ir-Pi via a cyclic voltammetry (CV) co-deposition method, and the obtained Iridium–cobalt phosphates (IrCo-Pi) shows not only a reduced usage of Ir but also an enhanced OER activity over the pristine Ir-Pi due to the synergetic effect between Ir and Co [35]. Therefore, using IrCo-Pi to modify CuWO4 would be a possible approach to achieve a highly improved PEC OER performance for the CuWO4 photoanode while reducing the use of high-cost Ir.
Herein, we add IrCo-Pi on the surface of CuWO4 nanoflakes (NFs) via a facile CV co-deposition method to alleviate the interfacial charge recombination. The obtained CuWO4/IrCo-Pi composites exhibit a highly-boosted PEC OER activity and an improved stability in neutral KPi solution over the bare CuWO4. We reveal that the greatly-enhanced hole-collection efficiency, which originates from the reduced charge-transfer resistance, accounts for the advanced PEC OER activity. The results presented here, as far as we know, stand as the best enhancement ever achieved on cocatalyst-modified CuWO4. This work demonstrates that it is feasible to design an efficient CuWO4/cocatalyst system towards PEC water oxidation.

2. Results and Discussion

2.1. Morphology, Chemical Composition, Crystal Structure and Bonding State of CuWO4 and CuWO4/IrCo-Pi

As shown in Figure 1a,b, both CuWO4 and CuWO4/IrCo-Pi exhibit similar network structures, suggesting that the deposition process imposed negligible impact on the morphology. Nevertheless, the roughening of the surface and the increase in the average thickness of nanoflakes (See the thickness distribution in Supplementary Figure S1a,b) reveal that a ca. 5 nm-thick film has been deposited on CuWO4 NFs. The elemental signals of Ir, Co, and P shown in Supplementary Figure S1c confirm the successful integration of IrCo-Pi with CuWO4. Moreover, the three elements are evenly distributed throughout the entire NF (Figure 1c). Such a uniform distribution is attributed to the merit of CV co-deposition method, since it ensures a fast-alternate deposition of multiple components within a certain potential window [35]. The crystalline structures of CuWO4 and CuWO4/IrCo-Pi are shown in Figure 1d,e. Well-resolved lattice fringes appear in both samples, and the lattice spaces of 0.376 nm and 0.487 nm are assigned to the (0-11) and (001) facet of triclinic CuWO4 phase, respectively. Moreover, the formation of a uniform amorphous IrCo-Pi layer outside the single-crystalline CuWO4 nanoflake proves the successful preparation of CuWO4/IrCo-Pi composite, and the thickness of the layer (ca. 5 nm) matches well the results shown in Supplementary Figure S1a,b. In line with the high-resolution transmission electron microscopy (HRTEM) tests, no signal of crystalline peaks related to IrCo-Pi was collected in the X-ray diffraction (XRD) patterns of CuWO4/IrCo-Pi (Figure 1f) and FTO/IrCo-Pi (Supplementary Figure S1d).
X-ray photoelectron spectroscopic (XPS) measurements were conducted to explore the elemental constitutions and bonding states of CuWO4/IrCo-Pi with different Ir-to-Co ratio (denoted as RIr-to-Co). The obtained survey spectra in Supplementary Figure S2 reveal that Cu, W, Ir, Co, P and O are the major compositions of CuWO4/IrCo-Pi, further demonstrating the successful combination of IrCo-Pi with CuWO4. As shown above (Figure 1c), CV co-deposition enables the two elements (Ir and Co) to be uniformly mixed at the microscopic scale, which creates an environment in which synergistic effect between adjacent atoms can occur. To probe such issue, the high-resolution spectra of Co 2p and Ir 4f in CuWO4/IrCo-Pi were examined, and the results are shown in Figure 2. All the Co-containing samples exhibit typical Co 2p1/2 (ca. 795.7 eV) and Co 2p3/2 (ca. 780.7 eV) peaks with the satellite peaks next to them, indicative of the co-existence of Co2+ and Co3+ [5,36], and the peak positions agree well with the previously-reported values [5,36,37]. As for Ir, the characteristic peaks of Ir 4f5/2 (ca. 65.4 eV) and Ir 4f7/2 (ca. 62.4 eV) are also presented in all the Ir-containing samples [34,35]. The mutual effect between Ir and Co is reflected by the shift of peak position. In detail, the Co 2p1/2 and Co 2p3/2 peaks shift negatively with the increase of Ir contents, whereas the Ir 4f5/2 and Ir 4f7/2 peaks shift positively when the content of Co grows. The shift of the XPS peaks reveals a regional electron transfer from Ir to Co along the Ir-O-Co bond in IrCo-Pi, which accelerates the rate-limiting step of the OER for IrCo-Pi and finally results in an enhanced OER activity of IrCo-Pi over pristine Ir-Pi [35]. Further, the actual values of RIr-to-Co were also quantified and the detailed results are summarized in Supplementary Table S1.

2.2. Optical Absorption Property of IrCo-Pi, CuWO4 and CuWO4/IrCo-Pi

Once an OER cocatalyst is deposited on a photoanode, it will cause the light scatter and absorption, and thereby inevitably influence the light-absorbing property of the photoanode [37,38]. Accordingly, a suitable cocatalyst should be as optically transparent as possible in the UV and visible-light regions. Herein, UV-vis measurements were carried out on all the samples. As shown in Figure 3a, IrCo-Pi only shows a light-grey color after 80 cycles of deposition, and its absorption intensity is negligible relative to that of CuWO4. Upon combination with IrCo-Pi (Figure 3b), all the CuWO4/IrCo-Pi samples exhibit a nearly similar performance as bare CuWO4 in term of absorption intensity. The results above testify that IrCo-Pi, with its weak light absorption capacity, is an ideal OER cocatalyst candidate.

2.3. Photoactivity and Photostability of CuWO4 and CuWO4/IrCo-Pi toward PEC OER

The photocurrent density vs. applied potential (Jph-V) curves for CuWO4 and CuWO4/IrCo-Pi with 30 deposition cycles are shown in Figure 4a. The relationship between the photocurrent density and deposition cycle are shown in Supplementary Figure S3. Given that the dark current density (Jdark) for CuWO4/IrCo-Pi is no longer inappreciable over the whole potential range (see Supplementary Figure S4a,b), Jph herein is the photocurrent density obtained by deducting Jdark from the total (Jtotal) under illumination. As shown, Co-Pi only brings about mild improvement in PEC performance while obvious enhancement is obtained via Ir-Pi modification, revealing that Ir-Pi has a much higher activity as an OER cocatalyst than Co-Pi. Although the increased proportion of Ir-Pi in IrCo-Pi hybrids leads to better enhancement effect, such an effect does not merely increase monotonically with the growth of RIr-to-Co, on the ground that the mutual effect between Ir and Co can enhance the inherent activity of IrCo-Pi [35]. When Ir-Pi gradually becomes the main composition in IrCo-Pi hybrids (the nominal and actual RIr-to-Co exceed 5:1 and 1:1, respectively, see Supplementary Table S1), IrCo-Pi starts to outperform Ir-Pi. The optimized RIr-to-Co is 9:1 under which the mutual effect is demonstrated to be the strongest due to the nearly 1:1 ratio of Ir-to-Co [35].
As shown in Figure 4b, with the aid of IrCo-Pi with the optimized RIr-to-Co (hereafter, denoted as IrCo(9:1)-Pi), the photocurrent density at 1.23 V vs. RHE rises to 0.54 mA∙cm−2, a ca. 70% increase over that of CuWO4 (0.32 mA∙cm−2). However, no obvious shift of the OER onset potential is observed, which is also found in other reports on cocatalyst-modified CuWO4 [26,39,40,41]. This indicates that the charge-transfer resistance around the onset potential range is quite large and thus even an OER cocatalyst fails to effectively accelerate the charge transfer. The comparison between our study and previous ones has been made and the results are summarized in Table 1. We can see that the results herein represent the highest level ever achieved on cocatalyst-modified single-phase CuWO4. Note that, in Ref. [40,41], although the authors named the photoanode CuWO4, it actually was, as the authors mentioned in the articles, a mixture of WO3 and CuWO4 based on the XRD data where unique WO3 peaks at ca. 33.2° can be clearly seen. According to the prior study, in a cocatalyst-modified WO3/CuWO4 composite system, WO3 contributes largely to the shift in the photocurrent density [27]. Besides, the incident photon-to-current efficiency (IPCE) data in Figure 4c show that CuWO4/IrCo(9:1)-Pi exhibits higher values than CuWO4 within the whole wavelength range, signifying that it can convert the incident light more efficiently.
By comparing the results herein with those in our previous work [35], we find that the optimized RIr-to-Co would vary based on the number of deposition circles, the substrates for deposition. Despite this, we still would like to highlight that IrCo-Pi, with its ease of preparation, reduced usage of high-cost Ir and superior performance over pure Ir-Pi, can enrich the photoanode/cocatalyst systems.
The stability during solar water oxidation is another vital factor to evaluate a photoanode. Herein, the long-term durability tests were conducted on both CuWO4 and CuWO4/IrCo(9:1)-Pi in a neutral KPi solution. As shown in Figure 4d, CuWO4/IrCo(9:1)-Pi shows a higher photocurrent density than CuWO4, manifesting that IrCo-Pi improves the PEC OER activity of CuWO4. Over a 4-h continuous illumination, CuWO4 reveals a medium decrease in the photocurrent density, consistent with the prior reports where CuWO4 is demonstrated to be slightly unstable in KPi solution [19,21,28]. This instability is closely related to the phosphate anion, not pH [21]. Upon using a potassium borate (KBi) solution (same pH and concentration), CuWO4 would display a superior photostability (See Figure S4c) [21,28,29]. CuWO4/IrCo(9:1)-Pi, in contrast, shows a nearly constant photocurrent density under the same conditions. The enhanced stability can be ascribed to the two following reasons: First, the as-deposited IrCo-Pi layer well coats CuWO4 (see Figure 1b,e) and then prevents the photoinduced corrosion of CuWO4 in KPi solution. Second, IrCo-Pi itself shows an excellent stability in KPi solution [35], ensuring long-term protection of the inner CuWO4 NFs.

2.4. Interfacial Charge-Transfer Behavior of CuWO4 and CuWO4/IrCo-Pi

The primary aim of IrCo-Pi modification is to reduce the interfacial electron–hole recombination. Herein, photocurrent transient measurements were conducted on CuWO4 and CuWO4/IrCo(9:1)-Pi under chopped illumination to study such recombination. As shown in Figure 5a and Supplementary Figure S5a, sharp anodic/cathodic photocurrent spikes appear on both samples when the light is suddenly on or off, indicative of interfacial charge recombination [42,43,44]. In detail, upon illumination, the photoinduced holes transport to the CuWO4/solution interface and then charge the surface states (non-Faradaic current) or oxidize H2O (Faradaic current) [44,45,46,47]. Both procedures contribute to the initial anodic current. Then the non-Faradaic current decays rapidly owing to the electron–hole recombination [46,47]. Therefore, water oxidation finally dominates the whole process and the steady-state current density agrees well with the value shown in J-V curves (See Figure 4a) [47]. When the light is off, the trapped holes at the surface states would be reduced by the electrons, giving rise to the cathodic spike [42,43]. The current eventually decreases to ca. 0 mA∙cm−2, since no reaction occurs at this potential without illumination (see Supplementary Figure S4a). The charge-recombination rate can be reflected by the decay time (D) of the transient spike, which is calculated via the following equation [42]:
D = (ItIs)/(IinIs)
where It is the current at time t, Is is the steady-state current, and Iin is the initial spike current. The time t at which In D = −1 is generally used to evaluate and compare the transient decay time [42,43,46]. As seen from Figure 5b, CuWO4/IrCo(9:1)-Pi shows a relatively longer decay time than bare CuWO4, suggesting a lower electron–hole recombination rate for CuWO4/IrCo(9:1)-Pi. This can be ascribed to the fact that the photogenerated holes can be efficiently captured by IrCo-Pi and then quick consumed in OER catalyzed by IrCo-Pi. Accordingly, the presence of IrCo-Pi on the surface of CuWO4 significantly accelerates the separation of photoinduced electron–hole pairs and hence boosts the performance for PEC OER.
To quantify the hole-collection efficiency (ηhc) at the CuWO4/solution interface, PEC measurements were carried out in the presence of hole scavenger Na2SO3, which can consume the photogenerated holes rapidly and thus improve ηhc to ca. 100% [48]. Due to the nearly complete suppression of surface recombination, J-V curve for Na2SO3 oxidation (Supplementary Figure S5b) displays a more negative OER onset potential and a much higher photocurrent density over the whole potential window. Accordingly, ηhc is defined by the following equation [48]:
ηhc = Jph-H2O/Jph-Na2SO3
where Jph-H2O and Jph-Na2SO3 refer to the photocurrent density for H2O and Na2SO3 oxidation, respectively. As shown in Figure 5c, although ηhc gradually increases with the applied potential, it only shows a value of ca. 55% even at 1.23 V vs. RHE, indicating that near half of the photogenerated holes on the surface are consumed via the useless recombination. With IrCo(9:1)-Pi modification, ηhc is apparently improved in the whole potential range and shifts to ca. 90% at 1.23 V vs. RHE, suggesting that the vast majority of photogenerated holes are captured by IrCo-Pi and then participate the OER.
Electrochemical impedance spectroscopy (EIS) measurements were conducted under illumination to explore the reasons for the enhanced photoactivity. As shown in Figure 5d, only one semicircle is presented for both the bare CuWO4 and cocatalyst-modified CuWO4, therefore, a simple but typical Randel’s circuit (inset in Figure 5d) was used to interpret the obtained EIS data [21,40,41,49,50]. The equivalent circuit consists of the series resistance (Rs), the constant phase element (CPE) and the charge-transfer resistance (Rct). The fitting results are summarized in Supplementary Table S2. Among all the samples, CuWO4/IrCo(9:1)-Pi exhibits the smallest Rct value (ca. 1114 Ω∙cm2), suggesting the fastest interfacial charge-transfer kinetics. The reduced charge-transfer barrier can boost the electron–hole separation on the surface and thus reduce the useless recombination, finally leading to an enhanced hole-collection efficiency.

2.5. The Interfacial Charge-Transfer Process for CuWO4/IrCo-Pi

The charge-transfer process at the photoanode/solution interface for CuWO4/IrCo(9:1)-Pi is illustrated in Figure 6. Upon illumination, holes are generated in the valence band of CuWO4 located at ca. 2.80 V vs. RHE [22], and hence they have sufficient driving force to oxidize Co2+ and Ir3+ to Co3+/4+ and Ir4+/5+, respectively [34,36]. Both the Co2+ and Ir3+ oxidation via photogenerated holes are kinetically faster than the direct H2O oxidation, and therefore the holes are transferred rapidly with the aid of IrCo-Pi and the electron–hole recombination is suppressed effectively. The obtained Co3+/4+ and Ir4+/5+ show high activity toward OER [34,36], and importantly, they would be reduced back to Co2+ and Ir3+ as the OER proceeds. Such oxidation-reduction cycle ensures a lasting enhancement effect in the photoactivity via IrCo-Pi. Therefore, through an alternative IrCo-Pi involved pathway for photoinduced holes, the interfacial charge recombination is considerably reduced, leading to an enhanced hole collection efficiency and an eventually improved photoactivity.

3. Materials and Methods

3.1. Chemicals and Substrates

Tungstic acid (H2WO4), polyvinyl alcohol ((C2H4O)n), hydrogen peroxide (H2O2), acetonitrile (C2H3N), urea (H2NCONH2), oxalic acid (H2C2O4), hydrochloric acid (HCl), and acetic acid (C2H4O2) were purchased from Beijing Chemical Works. Cupric nitrate trihydrate (Cu(NO3)2·3H2O) was purchased from Sinopharm Chemical Reagent Co., Ltd. Cobalt(II) chloride hexahydrate (CoCl2∙6H2O, 99.9%), iridium(III) chloride hydrate (IrCl3∙xH2O, 99.9%), potassium dihydrogen phosphate (KH2PO4, 99.5%), and dipotassium hydrogen phosphate (K2HPO4, 99.5%) were purchased from Alfa Aesar Chemicals (Co., Ltd. Shanghai, China). Fluorine-doped tin oxide (FTO) substrates (8 Ω∙cm2, transmittance of 80%) were purchased from Asahi Glass (Co., Ltd. Tokyo, Japan). All the chemicals used herein were of analytical grade without further purification. All the aqueous solutions herein were prepared using deionized water (18 MΩ∙cm).

3.2. Preparation of CuWO4 and CuWO4/IrCo-Pi

CuWO4 NFs were synthesized using a sacrificial template method introduced by our group [24], where WO3 NFs prepared via a reported hydrothermal approach [7] served as the templates. IrCo-Pi was coupled with CuWO4 NFs via a CV co-deposition method within a potential window from 0.65 V to 1.85 V vs. RHE at a scan rate of 0.05 V∙s−1 for different cycles. The solution for deposition consisted of 0.1 M KH2PO4/K2HPO4 (KPi) (pH 7) and a 0.5 mM mixture of CoCl2 and IrCl3. The nominal molar ratio of Ir to Co (RIr-to-Co) was altered within a predefined range. For convenience, RIr-to-Co = 1:0 or 0:1 refers to pure Ir-Pi or Co-Pi, respectively.

3.3. Characterization

X-ray diffraction (XRD) data were recorded by a D/MAX-2500 advance powder X-ray diffractometer (Rigaku Co., Tokyo, Japan) with Cu Kα radiation. High-resolution transmission electron microscopy (HRTEM) data were collected by a field emission JEM-2100F microscope (JEOL Ltd., Tokyo, Japan) under an accelerating voltage of 200 kV. The morphologies and elemental compositions were analyzed using a S-4800 field emission scanning electron microscope (FESEM) (Hitachi Ltd., Tokyo, Japan) equipped with an energy dispersive X-ray analysis (EDXA) system under an accelerating voltage of 5 kV. UV–vis diffuse reflection spectra were gathered via a 3600 UV–vis–NIR spectrophotometer (Shimadzu Co., Kyoto, Japan). X-ray photoelectron spectroscopic (XPS) spectra were recorded via an Escalab 250Xi (Thermo Fisher Scientific Co., Waltham, MA, USA) using an Al-monochromatic X-ray with a power of 200 W. All the binding energies herein have been calibrated against C 1s at 284.8 eV.

3.4. Electrochemical and PEC Measurements

All the electrochemical and PEC measurements were carried out on a CHI 660E work station (CH Instruments Co., Austin, TX, USA) equipped with a three-electrode configuration. The bare CuWO4 or CuWO4/cocatalyst, a Pt mesh, and a saturated calomel electrode (SCE) acted as the working, auxiliary, and reference electrode, respectively. A 300 W Xe lamp equipped with an AM 1.5 G filter (PLS-SXE300, Beijing Perfectlight Co., Ltd. Beijing, China) was employed as the light source. The samples were immersed into the electrolyte with a contact area of 0.21 cm2 under front-side illumination and the light intensity was to 100 mW∙cm−2. The impedance spectra obtained in the electrochemical impedance spectroscopy (EIS) tests were fitted and analyzed via Zview software. The data of incident photon-to-current efficiency (IPCE) were obtained by recording the current density at 1.23 V vs. RHE under a certain monochromatic light (wavelength range: 350–600 nm, step: 5 nm). The detailed values were calculated via Equation (3) below [51]:
IPCE = (1240J)/(λIlight) × 100%
where λ is the wavelength of the incident light (nm), J is the current density (mA∙cm−2) under a certain monochromatic light, and Ilight is the intensity of the incident light (mW∙cm−2). The electrolyte was a 0.1 M KPi buffer solution (pH 7) with or without adding 0.1 M Na2SO3. All potentials herein were against SCE and converted to RHE via the Equation (4) below [24]:
ERHE = ESCE + 0.059*pH + 0.244

4. Conclusions

In summary, IrCo-Pi was electrodeposited on CuWO4 NFs to alleviate the interfacial electron–hole recombination and thereby enhance the photoactivity of CuWO4. After the CuWO4 NFs are modified with IrCo(9:1)-Pi, the photocurrent density of the obtained composite photoanode at 1.23 V vs. RHE increases from 0.32 mA∙cm−2 to 0.54 mA∙cm−2. Such an improvement stands as the highest level ever achieved on cocatalyst-modified single-phase CuWO4. The high photoactivity is attributed to the greatly enhanced hole collection efficiency, which results from the reduced interfacial charge-transfer resistance. In addition, CuWO4/IrCo(9:1)-Pi exhibits a superior photostability in neutral KPi solution over the bare CuWO4. This work reveals that it is feasible to obtain a highly-enhanced PEC OER performance for CuWO4 via cocatalyst modification.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/8/913/s1, Figure S1: The detailed thickness distribution of (a) CuWO4 NFs and (b) CuWO4/IrCo-Pi NFs. (c) EDS spectrum of CuWO4/IrCo-Pi. (d) XRD patterns of FTO, FTO/Co-Pi, FTO/Ir-Pi, and FTO/IrCo-Pi, Figure S2: XPS survey of (a) CuWO4/Co-Pi (b) CuWO4/IrCo-Pi and (c) CuWO4/Ir-Pi, Figure S3: (a) The photocurrent density vs. applied potential (Jph-V) curves of CuWO4 and CuWO4/IrCo(9:1)-Pi with varied deposition cycles. (b)The corresponding Jdark-V and Jtotal-V curves of CuWO4 and CuWO4/IrCo(9:1)-Pi with varied deposition cycles, Figure S4: (a) Jdark-V and Jtotal-V curves of CuWO4 and CuWO4/IrCo-Pi with varied RIr-to-Co. (b) The enlarged Jdark-V curves of CuWO4 and CuWO4/IrCo-Pi within the potential range from 1.40 V to 1.80 V vs. RHE. (c) Current density-time responses of CuWO4 at 1.23 V vs. RHE under continuous illumination, Figure S5: (a) Transient photocurrent density curves for CuWO4 and CuWO4/IrCo(9:1)-Pi at 1.05 V vs. RHE for multiple light-on/light-off cycles.(b) Jph-V curves of CuWO4 and CuWO4/IrCo(9:1)-Pi obtained in the presence or absence of hole scavenger (0.1 M Na2SO3), Table S1: The nominal and actual RIr-to-Co for CuWO4/IrCo-Pi samples, Table S2: The fitting values for Rs, Rct and CPE using the equivalent circuit displayed in the inset of Figure 5d.

Author Contributions

Conceptualization, Methodology, Formal analysis, Writing—original draft preparation, C.L. Conceptualization, Writing—review and editing, Supervision, P.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number (51672017, 51872015).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  2. 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] [PubMed]
  3. Hurst, J.K. In pursuit of water oxidation catalysts for solar fuel production. Science 2010, 328, 315–316. [Google Scholar] [CrossRef] [PubMed]
  4. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  5. Steinmiller, E.M.P.; Choi, K.S. Photochemical deposition of cobalt-based oxygen evolving catalyst on a semiconductor photoanode for solar oxygen production. Proc. Natl. Acad. Sci. USA 2009, 106, 20633–20636. [Google Scholar] [CrossRef] [Green Version]
  6. Kang, Z.; Si, H.; Zhang, S.; Wu, J.; Sun, Y.; Liao, Q.; Zhang, Z.; Zhang, Y. Interface engineering for modulation of charge carrier behavior in ZnO photoelectrochemical water splitting. Adv. Funct. Mat. 2019, 29, 1808032–1808050. [Google Scholar] [CrossRef]
  7. Su, J.; Feng, X.; Sloppy, J.D.; Guo, L.; Grimes, C.A. Vertically aligned WO3 nanowire arrays grown directly on transparent conducting oxide coated glass: Synthesis and photoelectrochemical properties. Nano Lett. 2011, 11, 203–208. [Google Scholar] [CrossRef]
  8. Jin, T.; Diao, P.; Xu, D.; Wu, Q. High-aspect-ratio WO3 nanoneedles modified with nickel-borate for efficient photoelectrochemical water oxidation. Electrochim. Acta 2013, 114, 271–277. [Google Scholar] [CrossRef]
  9. Hu, D.; Diao, P.; Xu, D.; Wu, Q. Gold/WO3 nanocomposite photoanodes for plasmonic solar water splitting. Nano Res. 2016, 9, 1735–1751. [Google Scholar] [CrossRef]
  10. Jin, T.; Diao, P.; Wu, Q.; Xu, D.; Hu, D.; Xie, Y.; Zhang, M. WO3 nanoneedles/α-Fe2O3/cobalt phosphate composite photoanode for efficient photoelectrochemical water splitting. Appl. Catal. B: Environ. 2014, 148-149, 304–310. [Google Scholar] [CrossRef]
  11. Wu, Q.; Xu, D.; Xue, N.; Liu, T.; Xiang, M.; Diao, P. Photo-catalyzed surface hydrolysis of iridium(iii) ions on semiconductors: A facile method for the preparation of semiconductor/IrOx composite photoanodes toward oxygen evolution reaction. Phys. Chem. Chem. Phys. 2017, 19, 145–154. [Google Scholar] [CrossRef] [PubMed]
  12. Xi, L.; Lange, K.M. Surface modification of hematite photoanodes for improvement of photoelectrochemical performance. Catalysts 2018, 8, 497–511. [Google Scholar] [CrossRef] [Green Version]
  13. Fu, Y.; Dong, C.L.; Zhou, W.; Lu, Y.R.; Huang, Y.C.; Liu, Y.; Guo, P.; Zhao, L.; Chou, W.C.; Shen, S. A ternary nanostructured α-Fe2O3/Au/TiO2 photoanode with reconstructed interfaces for efficient photoelectrocatalytic water splitting. Appl. Catal. B Environ. 2020, 260, 118206–118214. [Google Scholar] [CrossRef]
  14. Butler, M.A. Photoelectrolysis and physical properties of the semiconducting electrode WO3. J. Appl. Phys. 1977, 48, 1914–1920. [Google Scholar] [CrossRef]
  15. Liu, X.; Wang, F.; Wang, Q. Nanostructure-based WO3 photoanodes for photoelectrochemical water splitting. Phys. Chem. Chem. Phys. 2012, 14, 7894–7911. [Google Scholar] [CrossRef] [PubMed]
  16. Cho, I.S.; Kwak, C.H.; Kim, D.W.; Lee, S.; Hong, K.S. Photophysical, photoelectrochemical, and photocatalytic properties of novel SnWO4 oxide semiconductors with narrow band gaps. J. Phys. Chem. C 2009, 113, 10647–10653. [Google Scholar] [CrossRef]
  17. Kölbach, M.; Pereira, I.J.; Harbauer, K.; Plate, P.; Höflich, K.; Berglund, S.P.; Friedrich, D.; van de Krol, R.; Abdi, F.F. Revealing the performance-limiting factors in α-SnWO4 photoanodes for solar water splitting. Chem. Mat. 2018, 30, 8322–8331. [Google Scholar] [CrossRef] [Green Version]
  18. Ke, J.; Adnan Younis, M.; Kong, Y.; Zhou, H.; Liu, J.; Lei, L.; Hou, Y. Nanostructured ternary metal tungstate-based photocatalysts for environmental purification and solar water splitting: A review. Nano-Micro Lett. 2018, 10, 69–95. [Google Scholar] [CrossRef] [Green Version]
  19. Yourey, J.E.; Bartlett, B.M. Electrochemical deposition and photoelectrochemistry of CuWO4, a promising photoanode for water oxidation. J. Mat. Chem. 2011, 21, 7651–7660. [Google Scholar] [CrossRef]
  20. 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]
  21. 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]
  22. 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]
  23. 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]
  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]
  25. 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. Mat. Interfaces 2016, 8, 9211–9217. [Google Scholar] [CrossRef]
  26. 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]
  27. 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]
  28. 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]
  29. Li, C.; Diao, P. Fluorine doped copper tungsten nanoflakes with enhanced charge separation for efficient photoelectrochemical water oxidation. Electrochim. Acta 2020, 352, 136471–136479. [Google Scholar] [CrossRef]
  30. Peng, B.; Xia, M.; Li, C.; Yue, C.; Diao, P. Network structured CuWO4/BiVO4/Co-Pi nanocomposite for solar water splitting. Catalysts 2018, 8, 663–671. [Google Scholar] [CrossRef] [Green Version]
  31. Lo Vecchio, C.; Trocino, S.; Campagna Zignani, S.; Baglio, V.; Carbone, A.; Díez García, M.I.; Contreras, M.; Gómez, R.; Aricò, A.S. Enhanced photoelectrochemical water splitting at hematite photoanodes by effect of a NiFe-oxide co-catalyst. Catalysts 2020, 10, 525–541. [Google Scholar] [CrossRef]
  32. Hamann, T. Electron dynamics at the CuWO4/electrocatalyst interface for photoelectrochemical water oxidation. In Proceedings of the Processes at the Semiconductor Solution Interface, 235th ECS Meeting, Dallas, TX, USA, 26–30 May 2019; p. 1226. [Google Scholar] [CrossRef]
  33. Xu, X.T.; Pan, L.; Zhang, X.; Wang, L.; Zou, J.J. Rational design and construction of cocatalysts for semiconductor-based photo-electrochemical oxygen evolution: A comprehensive review. Adv. Sci. 2019, 6, 1801505–1801539. [Google Scholar] [CrossRef]
  34. Irshad, A.; Munichandraiah, N. High catalytic activity of amorphous Ir-Pi for oxygen evolution reaction. ACS Appl. Mat. Interfaces 2015, 7, 15765–15776. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, Z.; Lin, Z.; Diao, P. Hybrids of iridium–cobalt phosphates as a highly efficient electrocatalyst for the oxygen evolution reaction in neutral solution. Chem. Commun. 2019, 55, 3000–3003. [Google Scholar] [CrossRef]
  36. Kanan, M.W.; Nocera, D.G. In situ formation of an oxygen-evolving catalyst in neutral water containing phosphate and Co2+. Science 2008, 321, 1072–1075. [Google Scholar] [CrossRef] [Green Version]
  37. Jeon, T.H.; Choi, W.; Park, H. Cobalt–phosphate complexes catalyze the photoelectrochemical water oxidation of BiVO4 electrodes. Phys. Chem. Chem. Phys. 2011, 13, 21392–21401. [Google Scholar] [CrossRef] [Green Version]
  38. Abdi, F.F.; van de Krol, R. Nature and light dependence of bulk recombination in Co-Pi-catalyzed BiVO4 photoanodes. J. Phys. Chem. C 2012, 116, 9398–9404. [Google Scholar] [CrossRef]
  39. 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]
  40. 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]
  41. 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, 135125–135133. [Google Scholar] [CrossRef]
  42. Tafalla, D. Kinetic approach to the photocurrent transients in water photoelectrolysis at n-TiO2 electrodes. J. Electrochem. Soc. 1990, 137, 1810–1815. [Google Scholar] [CrossRef]
  43. Hagfeldt, A.; Lindström, H.; Södergren, S.; Lindquist, S.E. Photoelectrochemical studies of colloidal TiO2 films: The effect of oxygen studied by photocurrent transients. J. Electroanal. Chem. 1995, 381, 39–46. [Google Scholar] [CrossRef]
  44. Le Formal, F.; Sivula, K.; Grätzel, M. The transient photocurrent and photovoltage behavior of a hematite photoanode under working conditions and the influence of surface treatments. J. Phys. Chem. C 2012, 116, 26707–26720. [Google Scholar] [CrossRef]
  45. Kim, D.W.; Riha, S.C.; DeMarco, E.J.; Martinson, A.B.F.; Farha, O.K.; Hupp, J.T. Greenlighting photoelectrochemical oxidation of water by iron oxide. ACS Nano 2014, 8, 12199–12207. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, H.; Cheng, C. Three-dimensional FTO/TiO2/BiVO4 composite inverse opals photoanode with excellent photoelectrochemical performance. ACS Energy Lett. 2017, 2, 813–821. [Google Scholar] [CrossRef]
  47. Gao, L.; Li, F.; Hu, H.; Long, X.; Xu, N.; Hu, Y.; Wei, S.; Wang, C.; Ma, J.; Jin, J. Dual modification of a BiVO4 photoanode for enhanced photoelectrochemical performance. ChemSusChem 2018, 11, 2502–2509. [Google Scholar] [CrossRef]
  48. Gao, Y.; Hamann, T.W. Quantitative hole collection for photoelectrochemical water oxidation with CuWO4. Chem. Commun. 2017, 53, 1285–1288. [Google Scholar] [CrossRef]
  49. Klahr, B.; Gimenez, S.; Fabregat-Santiago, F.; Bisquert, J.; Hamann, T.W. Photoelectrochemical and impedance spectroscopic investigation of water oxidation with “Co–Pi”-coated hematite electrodes. J. Am. Chem. Soc. 2012, 134, 16693–16700. [Google Scholar] [CrossRef] [Green Version]
  50. Klahr, B.; Giménez, S.; Fabregat-Santiago, F.; Bisquert, J.; Hamann, T.W. Electrochemical and photoelectrochemical investigation of water oxidation with hematite electrodes. Energy Environ. Sci. 2012, 5, 7626–7636. [Google Scholar] [CrossRef] [Green Version]
  51. Chen, Z.; Jaramillo, T.F.; Deutsch, T.G.; Kleiman-Shwarsctein, A.; Forman, A.J.; Gaillard, N.; Garland, R.; Takanabe, K.; Heske, C.; Sunkara, M.; et al. Accelerating materials development for photoelectrochemical hydrogen production: Standards for methods, definitions, and reporting protocols. J. Mat. Res. 2011, 25, 3–16. [Google Scholar] [CrossRef]
Figure 1. Typical top-view scanning electron microscope (SEM) images of (a) CuWO4 and (b) CuWO4/IrCo-Pi. (c) Energy dispersive spectrum (EDS) elemental mappings of Ir, Co, P and O in CuWO4/IrCo-Pi. HRTEM images of (d) CuWO4 and (e) CuWO4/IrCo-Pi. (f) XRD patterns of CuWO4 and CuWO4/IrCo-Pi.
Figure 1. Typical top-view scanning electron microscope (SEM) images of (a) CuWO4 and (b) CuWO4/IrCo-Pi. (c) Energy dispersive spectrum (EDS) elemental mappings of Ir, Co, P and O in CuWO4/IrCo-Pi. HRTEM images of (d) CuWO4 and (e) CuWO4/IrCo-Pi. (f) XRD patterns of CuWO4 and CuWO4/IrCo-Pi.
Catalysts 10 00913 g001
Figure 2. High-resolution XPS spectra of (a) Co 2p and (b) Ir 4f for CuWO4/IrCo-Pi with different RIr-to-Co.
Figure 2. High-resolution XPS spectra of (a) Co 2p and (b) Ir 4f for CuWO4/IrCo-Pi with different RIr-to-Co.
Catalysts 10 00913 g002
Figure 3. UV-vis absorption spectra of (a) IrCo-Pi and CuWO4, (b) CuWO4/IrCo-Pi with varied deposition cycles. The insets in (a,b) show the optical images of FTO/IrCo-Pi and CuWO4/IrCo-Pi with varied deposition cycles, respectively. (From left to right, 0, 40, 60, and 80 cycles).
Figure 3. UV-vis absorption spectra of (a) IrCo-Pi and CuWO4, (b) CuWO4/IrCo-Pi with varied deposition cycles. The insets in (a,b) show the optical images of FTO/IrCo-Pi and CuWO4/IrCo-Pi with varied deposition cycles, respectively. (From left to right, 0, 40, 60, and 80 cycles).
Catalysts 10 00913 g003
Figure 4. (a) Jph-V curves of CuWO4 and CuWO4/IrCo-Pi with different RIr-to-Co. (b) Jph at 1.23 V vs. RHE for CuWO4 and CuWO4/IrCo-Pi with different RIr-to-Co. (c) IPCE of CuWO4 and CuWO4/IrCo(9:1)-Pi at 1.23 V vs. RHE over the wavelength range from 350 nm to 600 nm. (d) Current density-time responses of CuWO4 and CuWO4/IrCo(9:1)-Pi at 1.23 V vs. RHE under continuous illumination.
Figure 4. (a) Jph-V curves of CuWO4 and CuWO4/IrCo-Pi with different RIr-to-Co. (b) Jph at 1.23 V vs. RHE for CuWO4 and CuWO4/IrCo-Pi with different RIr-to-Co. (c) IPCE of CuWO4 and CuWO4/IrCo(9:1)-Pi at 1.23 V vs. RHE over the wavelength range from 350 nm to 600 nm. (d) Current density-time responses of CuWO4 and CuWO4/IrCo(9:1)-Pi at 1.23 V vs. RHE under continuous illumination.
Catalysts 10 00913 g004
Figure 5. (a) Transient photocurrent density curves for CuWO4 and CuWO4/IrCo(9:1)-Pi at 1.05 V vs. RHE for one light-on/light-off cycle. (b) The corresponding transient decay time for each photoanode. (c) ηhc for CuWO4 and CuWO4/IrCo(9:1)-Pi. (d) Nyquist plots of bare CuWO4, CuWO4/Co-Pi, CuWO4/Ir-Pi and CuWO4/IrCo(9:1)-Pi at 1.05 V vs. RHE within the frequency range of 10−1–104 Hz under illumination. The inset is the corresponding equivalent circuit.
Figure 5. (a) Transient photocurrent density curves for CuWO4 and CuWO4/IrCo(9:1)-Pi at 1.05 V vs. RHE for one light-on/light-off cycle. (b) The corresponding transient decay time for each photoanode. (c) ηhc for CuWO4 and CuWO4/IrCo(9:1)-Pi. (d) Nyquist plots of bare CuWO4, CuWO4/Co-Pi, CuWO4/Ir-Pi and CuWO4/IrCo(9:1)-Pi at 1.05 V vs. RHE within the frequency range of 10−1–104 Hz under illumination. The inset is the corresponding equivalent circuit.
Catalysts 10 00913 g005
Figure 6. The schematic illustration of the charge-transfer pathway at the photoanode/solution interface for CuWO4/IrCo-Pi.
Figure 6. The schematic illustration of the charge-transfer pathway at the photoanode/solution interface for CuWO4/IrCo-Pi.
Catalysts 10 00913 g006
Table 1. PEC OER performances of cocatalyst-modified CuWO4 or WO3/CuWO4 photoanodes in this work and previous literatures [26,27,39,40,41].
Table 1. PEC OER performances of cocatalyst-modified CuWO4 or WO3/CuWO4 photoanodes in this work and previous literatures [26,27,39,40,41].
ReferencePhotoanodeCocatalystJph (mA∙cm−2) at 1.23 V vs. RHE
BeforeAfterIncrement
This workCuWO4IrCo(9:1)-Pi0.320.5470%
This workCuWO4Ir-Pi0.320.5056%
[26]CuWO4MnNCN0.0220.0336%
[27]CuWO4Mn-Pi0.0580.07529%
[39]WO3/CuWO4Mn-Pi0.380.5850%
[40]WO3/CuWO4Co-Pi0.200.3785%
[41]WO3/CuWO4Ni-Pi0.400.6050%
The intensity of the incident light is 100 mW∙cm−2. The electrolyte used in this work and all the cited references is 0.1 M KPi solution with a pH value of 7.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/catal10080913

AMA Style

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(8):913. https://doi.org/10.3390/catal10080913

Chicago/Turabian Style

Li, Chao, and Peng Diao. 2020. "Boosting the Activity and Stability of Copper Tungsten Nanoflakes toward Solar Water Oxidation by Iridium-Cobalt Phosphates Modification" Catalysts 10, no. 8: 913. https://doi.org/10.3390/catal10080913

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