Metal–organic framework-derived nano-CoS-enhanced photoelectrochemical water splitting performance of the BiVO4 photoanode

Dongqin Li a, Xinchao Chen b, Mingshi Shao b, Shushi Hou b, Jingyi Guo c, Haoyu Jin c, Hao Yang d, Guikui Chen *c and Yongchao Huang *b
aInstitute of Quality Standard and Monitoring Technology for Agro-products of Guangdong Academy of Agricultural Sciences, Guangdong Provincial Key Laboratory of Quality & Safety Risk Assessment for Agro-products, Guangzhou 510640, China
bInstitute of Environmental Research at Greater Bay Area, Key Laboratory for Water Quality and Conservation of the Pearl River Delta, Ministry of Education, Guangzhou University, Guangzhou, 510006, China. E-mail: huangych@gzhu.edu.cn
cThe College of Natural Resources and Environment of South China Agricultural University, Guangzhou 510642, China. E-mail: guikuichen@scau.edu.cn
dGuangxi Key Laboratory of Electrochemical Energy Materials, School of Chemistry & Chemical Engineering, Guangxi University, Nanning, 530004, P. R. China

Received 17th March 2025 , Accepted 4th June 2025

First published on 5th June 2025


Abstract

The utilization of metal–organic frameworks (MOFs) as oxygen evolution reaction (OER) co-catalysts for BiVO4-based photoelectrochemical (PEC) water splitting is limited by their intrinsic low conductivity and low availability of accessible catalytic metal sites. In this work, we transformed the cobalt-based imidazole-based ZIF-67 catalyst into a MOF-derived CoS cocatalyst through a two-step vulcanization process to construct a novel M-CoS/BiVO4 photoanode. The resulting M-CoS/BiVO4 photoanode demonstrated a photocurrent density of 5.22 mA cm−2 at 1.23 V versus the reversible hydrogen electrode under AM 1.5G light irradiation. Both the experimental results and theoretical calculations indicated that the built-in electric fields of the CoS/BiVO4 heterojunction effectively suppressed charge recombination in the bulk system. Furthermore, the MOF-derived CoS provided active sites with larger specific surface areas and increased the electronic conductivity, which effectively enhanced the charge separation and water oxidation kinetics, promoting the PEC water splitting performance. These findings indicate the potential of this new method in designing highly efficient photoanodes from MOFs.


1. Introduction

Photoelectrochemical (PEC) water splitting is one of the promising strategies for producing clean and sustainable “green” hydrogen fuel, which can play an important role in solving the energy crisis.1,2 The oxygen evolution reaction (OER) occurring on the photoanode for water oxidation, involving a four-electron process, is thermodynamically unfavorable, and it is the rate-determining step.3–5 As a promising photoanode candidate, bismuth vanadate (BiVO4) has received much attention because of its high theoretical photocurrent of 7.5 mA cm−2 (under standard AM 1.5 light irradiation) and solar-to-hydrogen conversion efficiency of 9.2%.6–12 Unfortunately, the low carrier mobility (0.044 cm2 V−1 s−1) and short hole diffusion distance (70 nm) result in severe photogenerated charge recombination in BiVO4.13,14 In addition, the slow water oxidation kinetics (1.4 s−1) limits the amount of photogenerated current and efficiency.15

In recent decades, metal–organic framework (MOF)-based OER co-catalysts have been used to increase the photoelectrochemical water oxidation properties of BiVO4 photoanodes, passivating the surface states for suppressing charge recombination and exposing more active sites for decreasing the activation energy of the OER.16–23 MOFs consist of transition metal ions/clusters and organic ligands, with the characteristics of an orderly structure, uniformly distributed active sites, ordered atom arrangement, and high stability.24,25 For instance, a cobalt-based zeolitic imidazolate framework (Co-agZIF-62) was introduced on NiO/BiVO4 photoanodes for promoting charge transfer and separation and accelerating the interfacial reaction kinetics, resulting in a photocurrent density of 5.34 mA cm−2 at 1.23 V vs. RHE.26 Furthermore, crystalline NiFc-MOFs were integrated into BiVO4 photoanodes, resulting in a high photocurrent density of 4.34 mA cm−2 at 1.23 V vs. RHE.27 Although various MOFs have been applied to BiVO4 photoanodes for PEC reactions with good results, the inherent low conductivity and instability of MOFs impede achieving a better photocurrent.28 Therefore, researchers have modified MOFs as precursors to other substances via carbonization or oxidation, increasing their electrical conductivity.29 Musa et al.30 annealed MOFs at high temperatures to obtain MOF-derived RuO2/N,S-TiO2 heterojunctions, which demonstrated excellent photocatalytic activity for hydrogen generation. Moreover, MOF-derived Co3O4/BiVO4 was obtained through calcination of Co-MOF/BiVO4 in air, while the performance was still far from satisfactory (<2.5 mA cm−2).31,32 High temperature calcination of MOFs is a common strategy to improve the electrical conductivity of MOFs, which destroys the overall framework of the MOFs and then reduces their surface area.31,33 Therefore, a gentler modification method is needed to improve the electrical conductivity of MOFs and to provide a greater active area for reactions. Based on the above analysis, we propose obtaining small-structured CoS by the vulcanization of a MOF on BiVO4 to enhance the PEC performance. The improvement in performance should arise as CoS has a low Gibbs free energy for hydrogen adsorption and the edge atoms are the catalytic active sites. Furthermore, CoS has high electrical conductivity and can effectively boost electron transfer.

Herein, to prove our concept, a typical MOF (here, cobalt-imidazole-based ZIF-67) was simply integrated onto the BiVO4 photoanode. Subsequently, the ZIF-67/BiVO4 electrode was simply vulcanized to obtain the nanostructured CoS/BiVO4 photoanode (M-CoS/BiVO4). Such a structure could not only form a CoS/BiVO4 heterojunction, but could also protects the stability of CoS through the framework of the MOF. In this design, the vulcanization process would not damage the structure of the MOF, effectively retaining a larger surface area and providing more active sites. In addition, the generated CoS have smaller particle sizes. The resulting M-CoS/BiVO4 photoanode demonstrated a high photocurrent density of 5.22 mA cm−2 at 1.23 V vs. RHE in potassium borate buffer. Experimental study and theoretical calculations revealed that the formation of the CoS/BiVO4 heterojunction, the larger specific surface area and the more active sites could all effectively enhance the charge separation and water oxidation kinetics, thereby promoting the PEC water splitting performance. This study provides a new strategy for preparing MOF derivatives combined with semiconductors as photoanodes to promote PEC water oxidation.

2. Experimental section

Synthesis of M-CoS/BiVO4

Bismuth vanadate was obtained according to our previous reports.8,34 In brief, at room temperature, 30 mL aqueous solution containing 1 mmol Co(NO3)3·6H2O was added to 30 mL aqueous solution containing 8 mmol 2-methylimidazole and the mixture was then stirred for 10 min. The prepared BiVO4 was put into the above mixed solution for 3 min, and then the electrode was removed and washed with deionized water and dried in an oven at 60 °C for 60 min. Subsequently, the electrode was immersed in 0.03 M sodium sulfide solution for 10 min and then washed with deionized water and dried. Finally, the photoanode and 1 mg per mL thioacetamide ethanol solution were placed in a Teflon-lined stainless steel autoclave and heated at 120 °C for 5 h in an oven. Afterwards, after cooling to room temperature the final electrode was washed with ethanol and dried overnight in an oven at 60 °C and was denoted as M-CoS/BiVO4. The CoS/BiVO4 photoanode was prepared in the same way, except that 2-methylimidazole was not added.

Characterization

The phase of the samples were identified by X-ray diffraction (Supernova, Rigaku). The morphology was collected by SEM (Gemini SEM 500). High-resolution TEM images and EDS mapping were recorded on an FEI Tecnai G2F30 instrument. Raman spectroscopy analysis was performed with a 532 nm laser (inVia Qontor, RENISHAW). XPS (Thermo-VG Scientific, ESCALAB250) was carried out to analyze the valence states of the elements. UV-vis absorbance spectra and PL spectra were obtained using a Shimadzu UV-3600i Plus spectrometer and FLS1000 system from Edinburgh Instruments.

PEC performance measurements

All the PEC tests were carried out under simulated AM 1.5G irradiation with a 100 mW cm−2 light intensity from a xenon lamp (PLS-FX300HU, Beijing Perfect light). A three-electrode cell was used, with the BiVO4 electrode used as the working electrode, the Ag/AgCl electrode as the reference electrode, and Pt as the auxiliary electrode. The electrolyte was 0.5 M potassium borate solution (pH 9.5). The irradiation area of the working electrode is 1 cm2. Additionally, the light passes through the insulating side of the FTO substrate.

3. Result and discussion

The nanostructures of the prepared BiVO4 and M-CoS/BiVO4 photoanodes were studied by scanning electron microscopy (SEM). As displayed in Fig. 1a, the pristine BiVO4 sample displayed the accumulation of worm-like structures with a smooth surface. After introducing the MOF-derived CoS on the BiVO4 surface, flocculation appeared on the surface (Fig. 1b). Furthermore, in-depth information on the morphology of M-CoS/BiVO4 was obtained by transmission electron microscopy (TEM) analysis. As can be seen from Fig. 1c, there was a thin layer at the edge of BiVO4. The high-resolution TEM (HRTEM) image shows a lattice fringe spacing of 0.308 nm, which corresponds to the (121) crystal plane of BiVO4 (JCPDS No. 14-0688) (Fig. 1d).35 The surface of BiVO4 was covered with a layer of amorphous material, namely about 10 nm CoS particles enwrapping it. This shows that the vulcanization process converted the Co2+ in ZIF to CoS and preserved the MOF framework. Furthermore, the TEM-EDS (Fig. 1e) and SEM-EDS (Fig. S1) element mappings of M-CoS/BiVO4 both confirmed the uniform distribution of Co and S elements on the surface of BiVO4, suggesting the MOF-derived had been CoS successfully grown on the surface of the BiVO4 photoanodes.
image file: d5ta02162f-f1.tif
Fig. 1 SEM images of (a) BiVO4 and (b) M-CoS/BiVO4 samples. (c) TEM image, (d) HR-TEM image and (e) EDS elemental mapping images of the M-CoS/BiVO4 sample.

To study the effect of M-CoS on the chemical structure of the electrode, BiVO4 and M-CoS/BiVO4 photoanodes were first analyzed by X-ray diffraction (XRD) measurements. As displayed in Fig. 2a, the XRD peaks of M-CoS/BiVO4 photoanode were not different compared to pristine BiVO4, agreeing well with the standard card of monoclinic BiVO4 (JCPDS#14-0688) and the FTO substrate (SnO2, JCPD#46-1088). No peaks belonging to CoS were observed, which suggests that the CoS was in the amorphous state, as also confirmed by the TEM results (Fig. 1d). Furthermore, the existence of CoS was also identified by the Raman spectra. As displayed in Fig. 2b, pristine BiVO4 showed several peaks at 820, 707, 323, 210 and 120 cm−1. The peaks at 120 and 210 cm−1 were contributed by an external mode.36 The peaks at 366 and 323 cm−1 were assigned to the symmetric (δs) and antisymmetric (δas) bending modes of VO43−, respectively.37 The peaks at 820 and 707 cm−1 were attributed to the symmetric (υs) and antisymmetric (υas) V–O stretching modes, respectively.38 Compared with pristine BiVO4, a new peak at 473 cm−1 appeared for M-CoS/BiVO4, which could be assigned to the Eg mode of CoS,39 indicating the successful construction of the CoS/BiVO4 heterojunction. Subsequently, X-ray photoelectron spectroscopy (XPS) was performed to identify the chemical state and composition of M-CoS/BiVO4 photoanode. As displayed in Fig. 2c, the Bi 4f XPS spectra of pristine BiVO4 displayed two characteristic peaks at binding energies of 158.9 and 164.2 eV, corresponding to Bi 4f7/2 and Bi 4f5/2, respectively, implying the presence of Bi3+ in BiVO4.40,41 The Bi 4f peaks of M-CoS/BiVO4 were slightly shifted to higher binding energies, indicating a strong interaction between BiVO4 and CoS, which plays an important role in photogenerated carrier transfer. Interestingly, compared with BiVO4, the V 2p and O 1s peaks of M-CoS/BiVO4 also moved to higher binding energies. The peak at 532.5 eV for O 1s corresponding to dissociated oxygen species from water molecules (OC) was increased,42 suggesting that M-CoS/BiVO4 had a better hydrophilic ability and was conducive to the OER reaction. Furthermore, S 1s and Co 2p signals could be detected in M-CoS/BiVO4 photoanode, further confirming the presence of CoS (Fig. 2e).43 Moreover, the optical properties of the BiVO4 and M-CoS/BiVO4 photoanodes were recorded by UV-vis absorption spectroscopy (Fig. 2f). The pristine BiVO4 and M-CoS/BiVO4 photoanodes demonstrated a similar light absorption edge at 520 nm, while the light absorption intensity of M-CoS/BiVO4 was much higher than that of BiVO4, suggesting that the CoS could strengthen the light-absorption intensity.


image file: d5ta02162f-f2.tif
Fig. 2 (a) XRD patterns, (b) Raman spectra obtained with a 532 nm laser, (c) Bi 4f, (d) V 2p and O 1s, and (e) S 1s and Co 2p XPS spectra, and (f) UV-vis diffuse reflectance spectra of BiVO4 and M-CoS/BiVO4 photoanodes.

To investigate the effect of the MOF-derived CoS on the surface-active area, the electrochemical surface-active areas (ECSA) of the BiVO4 and M-CoS/BiVO4 photoanodes were calculated from the electrochemical double-layer capacitance (Cdl) obtained through cyclic voltammetry tests (Fig. S2). As shown in Fig. S2, M-CoS/BiVO4 photoanode achieved a higher ECSA value (71.1 μF cm−2) than the BiVO4 photoanode (58.4 μF cm−2), revealing that the active sites were significantly increased due to the exposure of the MOF-driven CoS on the surface of M-CoS/BiVO4 photoanode, favoring better surface oxidation reactions.44

Next, linear-sweep-voltammetry (LSV) was performed to evaluate the PEC water splitting performance of the pristine BiVO4 and M-CoS/BiVO4 photoanodes under AM 1.5G light irradiation and 0.5 M (pH 9.5) borate buffer solution electrolyte (0.05 V s−1 scan rate). The photocurrent density of M-CoS/BiVO4 photoanode at 1.23 V vs. RHE was 5.22 mA cm−2, which was, as expected, much higher than that of BiVO4 (1.95 mA cm−2) and ZIF/BiVO4 (3.68 mA cm−2), confirming that M-CoS could effectively improve the PEC performance of BiVO4 (Fig. 3a). Furthermore, compared with other reported BiVO4-based photoanodes (Table S1), the performance of M-CoS/BiVO4 photoanode was also quite excellent, suggesting M-CoS/BiVO4 is a promising PEC material. To verify the role of the MOF structure, we synthesized CoS/BiVO4 electrodes without adding ligands for the MOF. However, the current density obtained was only 3.07 mA cm−2 at 1.23 V vs. RHE (Fig. S3), confirming that the MOF structure also plays an important role in the performance. Interestingly, we found that the results of the two-step vulcanization process were more significant than those of the separate vulcanization process (Fig. S4). Meanwhile, the influence of the electrolyte pH value on the performance of M-CoS/BiVO4 was also studied. M-CoS/BiVO4 had the highest performance in the pH range at 9.5–11, so we controlled the electrolyte at pH 9.5 for PEC water splitting (Fig. S5). The amounts of hydrogen and oxygen gases evolved from M-CoS/BiVO4 photoanode were tested at 1.23 V vs. RHE and detected by gas chromatography. The total amounts of H2 and O2 reached 569.1 and 285.5 μmol cm−2, respectively. The faradaic efficiency was about 91.5% (Fig. S6).


image file: d5ta02162f-f3.tif
Fig. 3 (a) LSV curves of BiVO4, ZIF-BiVO4 and M-CoS/BiVO4 photoanodes at a 0.05 V s−1 scan rate. (b) Charge-separation efficiency (ηsep), (c) charge transfer-efficiency (ηtran), (d) ABPE curves, (e) IPCE curves, and (f) it curves at 1.23 V vs. RHE of BiVO4 and M-CoS/BiVO4. Conditions: under AM 1.5G light irradiation and 0.5 M borate buffer solution (pH = 9.5).

To analyze the role of M-CoS on the charge separation and charge transfer of BiVO4, LSV was performed with adding 0.5 M Na2SO3 as a hole-sacrificial agent in the electrolyte (Fig. S6). Fig. 3b demonstrates the charge separation of the BiVO4 and M-CoS/BiVO4 photoanodes. The charge-separation efficiency of BiVO4 was 54.0% at 1.23 V vs. RHE, while after loading M-CoS, the charge-separation efficiency reached 84.2%, confirming that the formation of the CoS/BiVO4 heterojunction could effectively enhance the charge separation. Also, as expected, the charge-transfer efficiency of M-CoS/BiVO4 increased, i.e. from 53.0% to 83.4%, revealing that M-CoS/BiVO4 exhibited accelerated carrier-transfer dynamics. Moreover, the applied bias photon-to-current efficiencies (ABPEs) of BiVO4 and M-CoS/BiVO4 were calculated and are displayed in Fig. 3d. Clearly, M-CoS/BiVO4 photoanode showed an excellent photoconversion efficiency and maximum ABPE value of 1.56% at 0.70 V. Additionally, the incident photon-to-current efficiency (IPCE) value of M-CoS/BiVO4 photoanode was much higher than that of BiVO4 over the entire window, suggesting that M-CoS/BiVO4 photoanode had a better photons-to-electrons conversion efficiency (Fig. 3e).45 Except for its performance, the stability of a photoanode under light irradiation is also an issue that needs to be considered in applications.46 The stabilities of pristine BiVO4 and M-CoS/BiVO4 were not satisfactory (Fig. 3f). The current density of M-CoS/BiVO4 decreased to only 3.87 mA cm−2 after 7.6 h under light irradiation, which was mainly due to the instability of CoS under light irradiation. According to our previous research, loading a NiFeOx cocatalyst can effectively enhance the stability of photoanodes.47 Therefore, NiFeOx was loaded on M-CoS/BiVO4 electrode to improve its stability (labeled as NiFeOx/M-CoS/BiVO4). The photocurrent density of NiFeOx/M-CoS/BiVO4 was retained as 4.61 mA cm−2 after 27 h light illumination, indicating that its stability had indeed been greatly improved (Fig. S7).

To gain a deeper understanding of the reasons for the enhanced PEC performance of M-CoS/BiVO4 photoanode, the electron-transport mechanism was studied through measuring the band positions of CoS and BiVO4. The band positions of CoS and BiVO4 are shown in Fig. S8, revealing that a type II heterojunction could be formed between CoS and BiVO4.48 This structure can effectively separate photogenerated electrons and holes, thereby improving the charge-transfer efficiency (Fig. 4a).49 The charge mobility process was studied by chopped chronoamperometry measurements at 1.23 V vs. RHE (Fig. S9) and the transient decay time (τ) could be obtained at the point where ln D = −1, which represents the charge lifetime of the bulk semiconductor.50

image file: d5ta02162f-t1.tif
where It, Iin and Ist are the photocurrent at time t and the initial state and steady state currents, respectively. Impressively, M-CoS/BiVO4 demonstrated a large decay time τ value of 2.3 s, which was higher than that of BiVO4 (1.2 s), which could be ascribed to the holes on the catalyst surface easily transferring to the electrolyte to prolong the carrier lifetime (Fig. 4b). Furthermore, the charge transfer and separation characteristics of photoanodes can be also observed by electrochemical impedance spectroscopy (EIS) (Fig. 4c).51,52 Here, under AM 1.5G light illumination, M-CoS/BiVO4 exhibited a smaller charge-transfer resistance, suggesting that the CoS/BiVO4 heterojunction could boost the charge-transfer ability at the photoanode/electrolyte interface for water oxidation.53


image file: d5ta02162f-f4.tif
Fig. 4 (a) Schematic of the charge transfer at CoS/BiVO4 heterojunction. (b) Plots of ln D versus time, (c) EIS curves under 1.23 V vs. RHE within the 0.1–105 Hz frequency range in 0.5 M KBi buffer solution, (d) time-resolved photoluminescence spectra, (e) normalized decay profiles, (f) OCP-derived carrier lifetimes, (g) M-S plots without light irradiation with 1 kHz frequency in 0.5 M KBi buffer solution, (h) delay of the cathodic photocurrent curves, (i) plot of the charge-storage capability against the applied bias of BiVO4 and M-CoS/BiVO4 photoanodes.

The charge behaviors of the BiVO4 and M-CoS/BiVO4 photoanodes were further examined by steady-state PL spectra.54 As displayed in Fig. S10, pristine BiVO4 displayed a very strong PL peak, which demonstrated the severe electron–hole recombination (whereby holes in the O 2p band recombine with electrons in the V 3d band).55,56 After loading CoS, the intensity of the PL peak was significantly reduced, suggesting that the CoS/BiVO4 heterojunction could reduce the recombination of charge carriers.57 Furthermore, time-resolved photoluminescence spectra were used to analyze the charge-carrier lifetimes of the BiVO4 and M-CoS/BiVO4 photoanodes with a wavelength of 540 nm (Fig. 4d). The average charge lifetimes (τavg) of M-CoS/BiVO4 were calculated to be 3.2 ns, which were higher than that of BiVO4 (1.9 ns), indicating that charge-carrier recombination was effectively suppressed in M-CoS/BiVO4 photoanode.

Next, the normalized open-circuit potential (OCP) decay profiles of the BiVO4 and M-CoS/BiVO4 photoanodes were calculated and are displayed in Fig. 4e. As expected, M-CoS/BiVO4 photoanode demonstrated faster OCP decay compared with BiVO4, confirming that charge was rapidly transferred to the surface and injected into the electrolyte for water oxidation reaction.58 The charge lifetime was calculated based on the following equation:59

image file: d5ta02162f-t2.tif
where τ is the charge lifetime, e is 1.602 × 10−19 C. k is 1.38 × 10−23 J K−1, T is the temperature (K), and dOCP/dt is the derivative of the OCP transient decay. As displayed in Fig. 4f, M-CoS/BiVO4 photoanode (0.052 s) demonstrated smaller carrier lifetime values than BiVO4 (0.13 s), revealing that M-CoS/BiVO4 offered advanced charge separation via the internal electric field. To quantify the charge-carrier concentrations in the BiVO4 and M-CoS/BiVO4 photoanodes, Mott–Schottky (M-S) curves were recorded.60,61 As shown in Fig. 4g, the positive slopes revealed that the BiVO4 and M-CoS/BiVO4 photoanodes were n-type semiconductors. Moreover, the charge-carrier density (Nd) was calculated and the values are displayed in Table S1. The Nd of M-CoS/BiVO4 photoanode reached 1.55 × 1018 cm−3, which was much higher than that of BiVO4 (1.08 × 1018 cm−3), suggesting that the CoS/BiVO4 heterojunction effectively increases the charge-carrier concentration and charge-transfer efficiency. Simultaneously, the flat band potential of M-CoS/BiVO4 was shifts negatively compared with pristine BiVO4, revealing that the band bending at the interface between the photoanode and the electrolyte was sharper and more conducive to water oxidation.

The charge-storage capability of M-CoS/BiVO4 was studied through the transient-state photocurrent method, in which the cathodic photocurrent peaks were measured under chopped illumination. The trapped holes on the surface were reduced when the light was cut off, representing the cathodic current formation. Fig. 4h exhibits the delay of the cathodic photocurrent curves, which is analyzed from the transient cathodic current in Fig. S11. The delay in the steady-state cathode current indicated that the holes were stored on the electrode surface. Therefore, M-CoS/BiVO4 photoanode had more holes in the surface, suggesting its excellent hole-transfer ability. The number of storage holes can be represented by the integration of the transient photocurrent and the steady-state photocurrent.62 The number of storage holes of M-CoS/BiVO4 photoanode was obviously higher than that of BiVO4, especially at low bias (Fig. 4i). The stored holes can eliminate the injection barrier to allow entry of the electrolyte and its participation in water oxidation at high potential, resulting in the reduction of the stored holes.

Next, Kelvin probe force microscopy (KPFM) was performed to understand the mechanism of the enhanced PEC water splitting activity of M-CoS/BiVO4 photoanode through the internal electrical field.63 As shown in Fig. 5a, the surface potential of pristine BiVO4 was measured under dark and light irradiation. It was found that the surface potential of BiVO4 did not significantly change under light irradiation, which indicated severe photogenerated electron–hole pair recombination. As expected, the surface potential of M-CoS/BiVO4 photoanode significantly increased under light irradiation, definitely confirming that the internal electrical field was beneficial for charge separation. Furthermore, to disclose the charge behavior on the surface of the different components (such as BiVO4 and CoS) in M-CoS/BiVO4 photoanode, in situ XPS was carried out with/without light illumination (500 mW, 405 nm, 5 min). The high-resolution XPS spectra of Bi 4f, V 2p and O 1s for pristine BiVO4 under light and without light illumination are displayed in Fig. S12. The peak spectra were not changed, suggesting that severe photogenerated electron–hole pair recombination occurred for pristine BiVO4 after light illumination. Interestingly, both peaks of V 2p and O 1s for M-CoS/BiVO4 photoanode were also unchanged, while the Bi 4f and Co 2p peaks moved to a higher binding energy direction under light irradiation. These experimental results confirm that more photogenerated holes were transferred to the Co sites through the Bi sites under light irradiation.64


image file: d5ta02162f-f5.tif
Fig. 5 Surface potential images of (a) BiVO4 and (b) M-CoS/BiVO4 photoanodes from KPFM measurements in the dark and under light irradiation. XPS spectra of (c) Bi 4f, (d) V 2p and O 1s, (e) Co 2p of M-CoS/BiVO4 photoanode under light irradiation.

Next, density functional theory (DFT) calculations were performed to better understand the role of nano CoS in electron transfer and in Gibbs free energy H2O oxidation.65 First, theoretical structural models of CoS(nano)/BiVO4 and CoS(solid)/BiVO4 were constructed (Fig. S13). The electron density difference (EDD) models of CoS(nano)/BiVO4 and CoS(solid)/BiVO4 are displayed in Fig. 6a and b. The more closely stacked electrons around BiVO4 in CoS(nano)/BiVO4 confirmed that nano CoS could facilitate charge transfer between CoS and BiVO4. Furthermore, the density of states (DOS) of the d-band centers of CoS(nano)/BiVO4 and CoS(solid)/BiVO4 were determined and are displayed in Fig. 6c and d. CoS(solid)/BiVO4 mainly involved the d-band of Co interacting with H2O, while CoS(nano)/BiVO4 involved the d-band of V and Co interacting with H2O. In addition, the d-bands of V and Co were closer in CoS(nano)/BiVO4, which would facilitate electron transport. The standard Gibbs free energy diagrams for the OER were calculated and are displayed in Fig. S14, where the formation of *OOH from *O was the potential rate-determining step with a low overpotential (η) of 180 mV. The introduction of nano CoS boosted the generation of *OOH to enhance the OER kinetics. In addition, the transient electric field distributions of BiVO4 and CoS(nano)/BiVO4 are shown in Fig. 6e. CoS(nano)/BiVO4 displayed the stronger transient electric field, revealing that the nano CoS boosted the charge transfer and water oxidation kinetics, thereby increasing the photocurrent density.


image file: d5ta02162f-f6.tif
Fig. 6 Electron density differences for (a) CoS(nano)/BiVO4 and (b) CoS(solid)/BiVO4. Calculated DOS of the d-band centers for (c) CoS(nano)/BiVO4 and (d) CoS(solid)/BiVO4. (e) Transient electric field distribution simulated using MATLAB simulations of BiVO4 and CoS(nano)/BiVO4 (−0.1 V vs. RHE).

4. Conclusions

This study demonstrated a novel metalorganic framework (MOF)-derived strategy for fabricating CoS/BiVO4 heterostructured photoanodes with exceptional photoelectrochemical performance. Through controlled ZIF-67 loading and subsequent vulcanization treatment, we successfully constructed a hierarchical M-CoS/BiVO4 architecture. The optimized photoanode achieved a record photocurrent density of 5.22 mA cm−2 at 1.23 V versus the reversible hydrogen electrode (RHE) in borate buffer (pH 9.5) under AM 1.5 G illumination. Systematic investigations revealed three synergistic enhancement mechanisms: (1) the strong built-in electric field at the heterointerface promoted the charge-carrier separation efficiency (ηsep = 84.2%), (2) the MOF structure offered a larger specific surface area and more active sites for OER reaction, and (3) CoS enhanced the electrical conductivity and accelerated the surface reaction kinetics for water oxidation. This MOF-mediated synthesis approach opens new avenues for designing efficient photoelectrodes for solar fuel generation.

Data availability

Data supporting the conclusions of this study are included in the manuscript and in the ESI. These data are also available from the corresponding authors upon reasonable request.

Author contributions

Dongqin Li: investigation, writing – original draft, funding acquisition. Xinchao Chen: investigation, methodology. Mingshi Shao: investigation, data curation. Shushi Hou: investigation, data curation. Jingyi Guo: data curation. Haoyu Jin: data curation. Hao Yang: software. Guikui Chen: formal analysis, writing – review & editing. Yongchao Huang: formal analysis, writing – review & editing, resources, supervision, project administration.

Conflicts of interest

There are no conflicts to declare

Acknowledgements

This work was supported by the Guangzhou basic research planning for basic and applied basic research project (2024A04J3256); Special fund for scientific innovation strategy construction of high level Academy of Agriculture Science (R2023YJ-YB3005), Guangdong Provincial Modern Agricultural Industry Common Key Technologies Research and Development Innovation Team Construction Project (Common Key Technologies for Agricultural Product Quality and Safety) under the Agricultural Sector Framework (2024CXTD18).

Notes and references

  1. D. K. Lee and K.-S. Choi, Nat. Energy, 2018, 3, 53–60 CrossRef CAS.
  2. F. Liang, R. van de Krol and F. F. Abdi, Nat. Commun., 2024, 15, 4944 CrossRef CAS PubMed.
  3. X. Hu, J. Huang, Y. Cao, B. He, X. Cui, Y. Zhu, Y. Wang, Y. Chen, Y. Yang, Z. Li and X. Liu, Carbon Energy, 2023, 5, e369 CrossRef CAS.
  4. L. Liu, M. Ruan, C. Wang and Z. Liu, Appl. Catal. B Environ. Energy, 2024, 354, 124117 CrossRef CAS.
  5. Z. Masoumi, M. Tayebi, Q. Zaib, S. A. M. Lari, B. Seo, C.-S. Lim, S. Yu, H.-G. Kim and D. Kyung, Coord. Chem. Rev., 2024, 503, 215641 CrossRef CAS.
  6. M. Tayebi and B.-K. Lee, Catal. Today, 2021, 361, 183–190 CrossRef CAS.
  7. V. Andrei, G. M. Ucoski, C. Pornrungroj, C. Uswachoke, Q. Wang, D. S. Achilleos, H. Kasap, K. P. Sokol, R. A. Jagt, H. Lu, T. Lawson, A. Wagner, S. D. Pike, D. S. Wright, R. L. Z. Hoye, J. L. MacManus-Driscoll, H. J. Joyce, R. H. Friend and E. Reisner, Nature, 2022, 608, 518–522 CrossRef CAS PubMed.
  8. X. Chen, X. Li, Y. Peng, H. Yang, Y. Tong, M.-S. Balogun and Y. Huang, Adv. Funct. Mater., 2025, 35, 2416091 CrossRef CAS.
  9. T. W. Kim and K.-S. Choi, Science, 2014, 343, 990–994 CrossRef CAS PubMed.
  10. B. Li, L. I. Oldham, L. Tian, G. Zhou, S. Selim, L. Steier and J. R. Durrant, J. Am. Chem. Soc., 2024, 146, 12324–12328 CrossRef CAS PubMed.
  11. S. Wang, K. Wan, J. Feng, Y. Yang and S. Wang, J. Mater. Sci. Technol., 2025, 217, 182–220 CrossRef CAS.
  12. C. Xiao, S. Assavachin, W. Hahn, L. Wang, K. van Benthem and F. E. Osterloh, J. Mater. Chem. A, 2025, 13, 7834–7844 RSC.
  13. B. He, Y. Cao, K. Lin, Y. Wang, Z. Li, Y. Yang, Y. Zhao and X. Liu, Angew. Chem., Int. Ed., 2024, 63, e202402435 CrossRef CAS PubMed.
  14. X. Xu, S. Jin, C. Yang, J. Pan, W. Du, J. Hu, H. Zeng and Y. Zhou, Sol. RRL, 2019, 3, 1900115 CrossRef CAS.
  15. T. N. Jahangir, N. Mahar, M. G. Ahmed, N. M. Alakel, A. Al-Ahmed, A. A. Al-Saadi and T. A. Kandiel, J. Mater. Chem. A, 2025, 13, 7860–7870 RSC.
  16. F. He, Y. Liu, X. Yang, Y. Chen, C.-C. Yang, C.-L. Dong, Q. He, B. Yang, Z. Li, Y. Kuang, L. Lei, L. Dai and Y. Hou, Nano-Micro Lett., 2024, 16, 175 CrossRef CAS PubMed.
  17. J.-W. Ji, L.-J. Zhang, H.-C. Wang, Y.-X. Duan, X.-Z. Yue, J.-H. Chen and S.-S. Yi, Chem. Eng. J., 2024, 484, 149597 CrossRef CAS.
  18. J.-B. Pan, B.-H. Wang, J.-B. Wang, H.-Z. Ding, W. Zhou, X. Liu, J.-R. Zhang, S. Shen, J.-K. Guo, L. Chen, C.-T. Au, L.-L. Jiang and S.-F. Yin, Angew. Chem., Int. Ed., 2021, 60, 1433–1440 CrossRef CAS PubMed.
  19. Q. Y. Tey, J. Z. Soo, W. C. Ng and M. N. Chong, J. Mater. Chem. A, 2025, 13, 9005–9038 RSC.
  20. C. W. S. Yeung, V. Andrei, T. H. Lee, J. R. Durrant and E. Reisner, Adv. Mater., 2024, 36, 2404110 CrossRef CAS PubMed.
  21. W. Zhang, R. Li, X. Zhao, Z. Chen, A. W.-K. Law and K. Zhou, ChemSusChem, 2018, 11, 2710–2716 CrossRef CAS PubMed.
  22. S. Zhou, K. Chen, J. Huang, L. Wang, M. Zhang, B. Bai, H. Liu and Q. Wang, Appl. Catal. B Environ., 2020, 266, 118513 CrossRef CAS.
  23. B. He, S. Jia, M. Zhao, Y. Wang, T. Chen, S. Zhao, Z. Li, Z. Lin, Y. Zhao and X. Liu, Adv. Mater., 2021, 33, 2004406 CrossRef CAS.
  24. Y.-J. Dong, J.-F. Liao, Z.-C. Kong, Y.-F. Xu, Z.-J. Chen, H.-Y. Chen, D.-B. Kuang, D. Fenske and C.-Y. Su, Appl. Catal., B, 2018, 237, 9–17 CrossRef CAS.
  25. L. Shi, D. Benetti, F. Li, C. Harnagea, Q. Wei and F. Rosei, J. Mater. Chem. A, 2024, 12, 29989–29997 RSC.
  26. Y. Song, Y. Ren, H. Cheng, Y. Jiao, S. Shi, L. Gao, H. Xie, J. Gao, L. Sun and J. Hou, Angew. Chem., Int. Ed., 2023, 62, e202306420 CrossRef CAS PubMed.
  27. W. Bai, H. Li, G. Peng, J. Wang, A. Li and P. F.-X. Corvini, Appl. Catal. B Environ. Energy, 2024, 352, 124023 CrossRef CAS.
  28. X.-Y. Yang, Z.-W. Chen, X.-Z. Yue, X. Du, X.-H. Hou, L.-Y. Zhang, D.-L. Chen and S.-S. Yi, Small, 2023, 19, 2205246 CrossRef CAS PubMed.
  29. Y. Li, Y. Xia, K. Liu, K. Ye, Q. Wang, S. Zhang, Y. Huang and H. Liu, ACS Appl. Mater. Interfaces, 2020, 12, 25494–25502 CrossRef PubMed.
  30. E. N. Musa, A. K. Yadav, K. T. Smith, M. S. Jung, W. F. Stickle, P. Eschbach, X. Ji and K. C. Stylianou, Angew. Chem., Int. Ed., 2024, 63, e202405681 CAS.
  31. D. Cardenas-Morcoso, R. Ifraemov, M. García-Tecedor, I. Liberman, S. Gimenez and I. Hod, J. Mater. Chem. A, 2019, 7, 11143–11149 RSC.
  32. D. Xu, T. Xia, W. Fan, H. Bai, J. Ding, B. Mao and W. Shi, Appl. Surf. Sci., 2019, 491, 497–504 CrossRef CAS.
  33. S. Zhang, Y. Lu, Q. Ding, Y. Yu, P. Huo, W. Shi and D. Xu, Colloids Surf., A, 2022, 655, 130282 CrossRef CAS.
  34. Y. Wang, Y. Chen, Y. Yun, X. Hong, Y. Huang and H. Ji, Appl. Catal. B Environ. Energy, 2024, 358, 124375 CrossRef CAS.
  35. Y. Chen, X. Li, H. Yang and Y. Huang, Small, 2024, 20, 2402406 CrossRef CAS.
  36. B. Liu, X. Wang, Y. Zhang, K. Wan, L. Xu, S. Ma, R. Zhao, S. Wang and W. Huang, Adv. Energy Mater., 2025, 15, 2403835 CrossRef CAS.
  37. Y. Wang, W. Wang, H. Mao, Y. Lu, J. Lu, J. Huang, Z. Ye and B. Lu, ACS Appl. Mater. Interfaces, 2014, 6, 12698–12706 CrossRef CAS.
  38. K.-H. Ye, Z. Chai, J. Gu, X. Yu, C. Zhao, Y. Zhang and W. Mai, Nano Energy, 2015, 18, 222–231 CrossRef CAS.
  39. J. Huo, J. Wu, M. Zheng, Y. Tu and Z. Lan, Electrochim. Acta, 2016, 187, 210–217 CrossRef CAS.
  40. G. Liu, S. Wang, C. Zhou, Q. Zhao, J. Hu, Z. Gui, Y. Chen, Y. Huang, P. Zhang and F. Wang, Fuel, 2024, 374, 132420 CrossRef CAS.
  41. M. Tayebi, Z. Masoumi, B. Seo, C.-S. Lim, H.-G. Kim and B.-K. Lee, ACS Appl. Energy Mater., 2022, 5, 11568–11580 CrossRef CAS.
  42. S. Wang, J. Hu, Q. Zhao, G. Liu, Z. Gui, X. Liu, Y. Huang, P. Zhang, Y. Chen and F. Wang, ACS Sustainable Chem. Eng., 2024, 12, 11754–11766 CrossRef CAS.
  43. W. T. Hong, M. Han, Y. Li, J. Kim, J. Lee, W. Yang, T. Yu and J. K. Kim, ACS Mater. Lett., 2024, 6, 4168–4174 CrossRef CAS.
  44. L. Wang, Z. Gao, K. Su, N. T. Nguyen, R.-T. Gao, J. Chen and L. Wang, Adv. Funct. Mater., 2024, 34, 2403948 CrossRef CAS.
  45. J. Su, L. Guo, N. Bao and C. A. Grimes, Nano Lett., 2011, 11, 1928–1933 CrossRef CAS PubMed.
  46. R. Lei, Y. Tang, W. Qiu, S. Yan, X. Tian, Q. Wang, Q. Chen, Z. Wang, W. Qian, Q. Xu, S. Yang and X. Wang, Nano Lett., 2023, 23, 11785–11792 CrossRef CAS PubMed.
  47. Y. Chen, J. Huang, J. Zhou, X. Li, H. Yang and Y. Huang, Mater. Today Adv., 2025, 25, 100554 CrossRef CAS.
  48. Q. Wang, L. I. Oldham, A. Giner-Requena, Z. Wang, D. Benetti, S. Montilla-Verdú, R. Chen, D. Du, T. Lana-Villarreal, U. Aschauer, N. Guijarro, J. R. Durrant and J. Luo, J. Am. Chem. Soc., 2024, 146, 34681–34689 CrossRef CAS PubMed.
  49. P. M. Rao, L. Cai, C. Liu, I. S. Cho, C. H. Lee, J. M. Weisse, P. Yang and X. Zheng, Nano Lett., 2014, 14, 1099–1105 CrossRef CAS PubMed.
  50. G. Liu, J. Shi, F. Zhang, Z. Chen, J. Han, C. Ding, S. Chen, Z. Wang, H. Han and C. Li, Angew. Chem., Int. Ed., 2014, 53, 7295–7299 CrossRef CAS PubMed.
  51. R. Zhang, S. Zhang, H. Xiao, J. An, Z. Wang, W. Luo, Y. Feng, B. Lu, P. Du and X. Lu, J. Colloid Interface Sci., 2025, 687, 531–539 CrossRef CAS.
  52. Y. Huang, Z. Guo, H. Liu, S. Zhang, P. Wang, J. Lu and Y. Tong, Adv. Funct. Mater., 2019, 29, 1903490 CrossRef CAS.
  53. S. Wang, T. He, J.-H. Yun, Y. Hu, M. Xiao, A. Du and L. Wang, Adv. Funct. Mater., 2018, 28, 1802685 CrossRef.
  54. B. Kang, M. Bilal Hussain, X. Cheng, C. Peng and Z. Wang, J. Colloid Interface Sci., 2022, 626, 146–155 CrossRef CAS.
  55. Y. Ma, S. R. Pendlebury, A. Reynal, F. Le Formal and J. R. Durrant, Chem. Sci., 2014, 5, 2964–2973 RSC.
  56. Y. Wang, J. Huang, Y. Chen, H. Yang, K.-H. Ye and Y. Huang, J. Colloid Interface Sci., 2024, 672, 12–20 CrossRef CAS.
  57. C. Zhen, H. Zhu, R. Chen, Z. Zheng, F. Fan, B. Li, X. Xu, Y. Du, H.-M. Cheng, K. Domen and G. Liu, J. Am. Chem. Soc., 2024, 146, 28482–28490 CAS.
  58. B. He, Y. Cao, K. Lin, M. Wu, Y. Zhu, X. Cui, L. Hu, Y. Yang and X. Liu, eScience, 2024, 100242 Search PubMed.
  59. M. Zhong, T. Hisatomi, Y. Kuang, J. Zhao, M. Liu, A. Iwase, Q. Jia, H. Nishiyama, T. Minegishi, M. Nakabayashi, N. Shibata, R. Niishiro, C. Katayama, H. Shibano, M. Katayama, A. Kudo, T. Yamada and K. Domen, J. Am. Chem. Soc., 2015, 137, 5053–5060 CrossRef CAS.
  60. X. Chen, C. Zhen, J. Li, J. Qiu, N. Li, N. Jia and G. Liu, Adv. Funct. Mater., 2024, 34, 2409566 CrossRef CAS.
  61. E. Omid Najafabadi, F. Razi Astaraei, M. Tayebi, Z. Masoumi, O. Moradlou, M. M. Momeni, H.-G. Kim, F. Karimian Bahnamiri, M. Khalili and B.-K. Lee, Mater. Chem. Phys., 2023, 309, 128430 CrossRef CAS.
  62. K. Zhang, B. Jin, C. Park, Y. Cho, X. Song, X. Shi, S. Zhang, W. Kim, H. Zeng and J. H. Park, Nat. Commun., 2019, 10, 2001 CrossRef PubMed.
  63. Y. Zhang, L. Xu, B. Liu, X. Wang, T. Wang, X. Xiao, S. Wang and W. Huang, ACS Catal., 2023, 13, 5938–5948 CrossRef CAS.
  64. J. Wiktor, F. Ambrosio and A. Pasquarello, ACS Energy Lett., 2018, 3, 1693–1697 CrossRef CAS.
  65. H. Huang, Z. Zhang, W. Xie, B. Fan, C. Wu, R. Jiang, J. Huang, B. Zhang, Y. Hou and Z. Yu, J. Colloid Interface Sci., 2024, 668, 551–564 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta02162f

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.