Boosting the photoelectrochemical water oxidation performance of bismuth vanadate by ZnCo2O4 nanoparticles

Jingwei Huang Yani Wang Kaiyi Chen Tingting Liu Qizhao Wang

Citation:  Jingwei Huang, Yani Wang, Kaiyi Chen, Tingting Liu, Qizhao Wang. Boosting the photoelectrochemical water oxidation performance of bismuth vanadate by ZnCo2O4 nanoparticles[J]. Chinese Chemical Letters, 2022, 33(4): 2060-2064. doi: 10.1016/j.cclet.2021.08.082 shu

Boosting the photoelectrochemical water oxidation performance of bismuth vanadate by ZnCo2O4 nanoparticles

English

  • The conversion of solar energy into hydrogen by photoelectrochemical (PEC) water splitting is a promising way to produce hydrogen, which can simultaneously solve the current energy shortage and environmental problems caused by fossil fuel combustion [1-3]. In the process of PEC water splitting, water oxidation (producing oxygen) and water reduction (producing hydrogen) reaction are involved. The water oxidation reaction involves a kinetic unfavorable four-electron transfer process, which determines the entire water splitting reaction rate [4, 5]. Researches are mainly focused on photoanodes where water oxidation reaction takes place. Metal oxides such as TiO2 [6, 7], Fe2O3 [8, 9], WO3 [10], ZnO [11] and BiVO4 [12-14]are often used as photoanode materials for PEC water oxidation. Among them, BiVO4 is a very excellent photoanode material with beneficial energy band structure and a band gap of 2.4 eV, enabling it to absorb part of visible light [15-17]. However, the slow migration rate of photogenerated electrons and holes in BiVO4 and sluggish kinetics for water oxidation render BiVO4 far below its theoretical photocurrent density [18, 19]. Strategies such as controlling morphology [20], constructing heterostructures [10] have been used to improve the performance of BiVO4. Besides, loading cocatalyst is an effective way to improve the surface reaction kinetics of BiVO4 for PEC water splitting [21, 22].

    Cobalt-based cocatalysts, such as cobalt oxides [23-25], cobalt hydroxides [26, 27], cobalt phosphide [28, 29], Co-Pi [30, 31], Co-based double hydroxide (LDH) [32], have been intensively studied as water oxidation catalysts due to their outstanding catalytic performance. Specially, ternary spinel MCo2O4 (M = Ni, Mn, Zn, etc.) have improved bearing conductivity and possessing more active sites than Co3O4, resulting in a better water oxidation capability [33-35]. Among them, ZnCo2O4 has a definite structure in which Zn2+ only replaces Co2+ in the tetrahedral sites in Co3O4, rarely affecting the Co3+ active site in the octahedral sites in Co3O4 [36]. Therefore, as compared to Co3O4, the use of ZnCo2O4 could reduce the use of rare and relative expensive Co element without the compromise of catalytic activity.

    In addition to loading cocatalyst, enhancing built-in electric field of a photoanode by constructing a heterojunction structure is also usually adopted. WO3@BiVO4 heterojunction photoanode [37], Fe2O3/BiVO4 heterojunction photoanode [38], g-C3N4@BiVO4 heterojunction photoanode [39] etc. are recently reported to be favorable for the separation of photo-generated carriers due to the formation of built-in electric field between these different semiconductors in heterojunction. However, the water oxidation reaction on surface of these photoanodes is still sluggish. To address this problem, cocatalysts are still needed. For example, NiFe-LDH was used to load on Fe2O3/BiVO4 heterojunction photoanode [40]. WO3/BiVO4 was modified by FeOOH to construct ternary WO3/BiVO4/FeOOH hierarchical photoanode [41]. Considering the complicated fabrication process of above mentioned photoanodes, loading a semiconductor cocatalyst that could enhance the built-in electric field as well as act as cocatalyst is a new guiding ideology to construct a high performance photoanode for water oxidation.

    In this work, ZnCo2O4 nanoparticles, acting as a p-type semiconductor and a cocatalyst simultaneously, were loaded on BiVO4 to form ZnCo2O4/BiVO4 composite photoanode by electrophoretic deposition. The formation of p-n junctions between ZnCo2O4 and BiVO4 is systematically confirmed, which introduces built-in electric field that is beneficial to the separation of photogenerated carriers. This composite photoanode shows high photocurrent for PEC water oxidation due to the improvement in charge separation efficiency and surface reaction efficiency.

    As can be seen from Fig. 1a, BiVO4 presents a coral-like structure composed of porous nanoparticles. The nanoparticles are interconnected on the surface of the FTO to form pores and channels. These abundant channels facilitate the contact between the semiconductor and electrolyte. From Fig. 1b, it can be clearly seen that ZnCo2O4 nanoparticles are uniformly loaded onto the surface of BiVO4 film by the electrophoretic deposition method. TEM image (Fig. 1c) shows that BiVO4 has a pore structure and smooth surface. However, a thin layer of nanoparticles appears on the surface of BiVO4 after loading ZnCo2O4 as shown in Fig. 1d. Figs. 1e and f show the HR-TEM images of ZnCo2O4/BiVO4 photoanode. The spacing of 0.20 nm and 0.48 nm correspond to the (400) plane of ZnCo2O4 (PDF card No. 23-1390) and the (110) plane of BiVO4 (PDF No. 14-0688), respectively, demonstrating the successful combination of ZnCo2O4 nanoparticles and BiVO4 film to form ZnCo2O4/BiVO4 photoanode.

    Figure 1

    Figure 1.  SEM images of (a) BiVO4, (b) ZnCo2O4/BiVO4-35. TEM images of (c) BiVO4, (d) ZnCo2O4/BiVO4-35. (e, f) HR-TEM images of ZnCo2O4/BiVO4-35.

    In order to clearly grasp the crystal structure of the sample, we performed XRD testing on the photoanode material. As shown in Fig. 2a, FTO substrate has the diffraction peaks located at 26.6°, 37.8°, 51.8°, 61.7° and 65.7° corresponding to the (110), (200), (211), (310) and (301) lattice diffractions of SnO2 (JCPDS No. 46–1088), respectively. BiVO4 shows diffraction peaks located at 18.7°, 28.8°, 30.5°, 34.5°, 35.2°, 40.0°, 42.5°, 46.0°, 47.3°, 50.3°, 53.3°, 58.3° and 59.3°, indicating the formation of monoclinic phase BiVO4 (JCPDS: 14-0688). The synthesized ZnCo2O4 powders have diffraction peaks mainly located at 19.0°, 31.2°, 36.8°, 44.7°, 59.3° and 65.1°. These diffraction peak positions correspond to the (111), (220), (311), (400), (511) and (440) crystal planes of the ZnCo2O4 (JCPDS: 23-1390), proving that the ZnCo2O4 nanoparticles were successfully synthesized. The XRD pattern of ZnCo2O4/BiVO4 composite electrode does not show obvious characteristic diffraction peaks of ZnCo2O4 due to small loading amount of ZnCo2O4 nanoparticles on BiVO4. X-ray photoelectron spectroscopy (XPS) test was used to determine the elements status of ZnCo2O4/BiVO4 electrode. It can be seen from Fig. 2b that two peaks of Co 2p in ZnCo2O4 are located 795.0 and 780.0 eV, respectively belonging to the Co 2p1/2 and Co 2p3/2 peaks of Co3+ in the octahedral site, and proving that the cobalt in the compound is +3 valence [36]. The binding energies of the Zn 2p3/2 and Zn 2p1/2 peaks in the compounds are located at 1021.1 eV and 1044.2 eV, indicating that Zn in the ZnCo2O4 is +2 valence in the compound (Fig. 2c) [42]. The O 1s spectrum in the ZnCo2O4/BiVO4 photoanode can be fitted into two diffraction peaks (Fig. 2d). The peak at 531.4 eV in the spectrum is derived from adsorbed OH on the electrode, while the peak at 529.7 eV is assigned to oxide lattice [43].

    Figure 2

    Figure 2.  (a) XRD patterns of FTO substrate, BiVO4, ZnCo2O4 and ZnCo2O4/BiVO4 composites and XPS spectra of (b) Co 2p, (c) Zn 2p, (d) O 1s in ZnCo2O4/BiVO4.

    The light absorption range of ZnCo2O4/BiVO4 photoanode is broadened slightly by supporting ZnCo2O4 cocatalyst (Fig. S1a in Supporting information). The light absorption intensity of the composite is also higher than that of BiVO4, indicating that ZnCo2O4 improves the light absorption capacity of BiVO4. The band gap of BiVO4 is about 2.46 eV, whilst ZnCo2O4/BiVO4 electrode shows an apparent band gap of 2.43 eV, demonstrating ZnCo2O4 can improve the light utilization efficiency of BiVO4 photoanode (Fig. S1b in Supporting information). In order to investigate whether the catalytic active area of BiVO4 film would be changed after loading ZnCo2O4, cyclic voltammetry (CV) tests of BiVO4 and ZnCo2O4/BiVO4 electrodes were carried out at different scan rates as shown in Figs. S1c and e (Supporting information). The catalytic active area of BiVO4 and ZnCo2O4/BiVO4 electrodes can be estimated from double layer capacitances that are linearly related to the slopes of difference values between anode current (J) and cathode current (J) in CVs against the scan rate as shown in Figs. S1d and f (Supporting information). Taking the current difference as y-axis and scanning rate as x-axis, the linear slope is twice over the double layer capacitances [44]. It can be seen from Figs. S1d and f that the slope of ZnCo2O4/BiVO4 is larger than that of BiVO4, demonstrating the catalytic active area of BiVO4 photoanode increases after loading ZnCo2O4.

    The LSV curves of BiVO4 and ZnCo2O4/BiVO4 photoanodes were tested under AM 1.5G simulated sunlight (100 mW/cm2) from back side of the sample. As shown in Fig. 3a, the photocurrent density of BiVO4 is 1.4 mA/cm2 at 1.23 V vs. RHE, while the photocurrent density of 3.0 mA/cm2 is obtained for the ZnCo2O4/BiVO4 photoanode. Besides, the onset potential of BiVO4 photoanode shifts ca. 0.1 V after loading ZnCo2O4. Co3O4/BiVO4 photoanode was also fabricated for comparison using the same electrophoretic deposition method as shown in Fig. 3a, giving a photocurrent of 2.2 mA/cm2 at 1.23 V vs. RHE at same reaction condition. Obviously, ZnCo2O4 shows more higher catalytic activity than that of Co3O4. The photocurrent density of ZnCo2O4/BiVO4 is 2.1 times that of BiVO4. This indicates that the performance of BiVO4 for PEC water splitting can be enhanced by ZnCo2O4. The comparison of ZnCo2O4/BiVO4 photoanode with other BiVO4 and Co based photoanodes is shown in Table S 1 (Supporting information). In order to investigate the effect of deposition time on the catalytic performance of ZnCo2O4/BiVO4, the electrophoretic deposition time of ZnCo2O4 was set be to 15, 20, 25, 30, 35 and 40 s for comparison. These composite photoanodes are named as ZnCo2O4/BiVO4-15, ZnCo2O4/BiVO4-20, ZnCo2O4/BiVO4-25, ZnCo2O4/BiVO4-30, ZnCo2O4/BiVO4-35 and ZnCo2O4/BiVO4-40 accordingly. As demonstrated in Fig. S 2 (Supporting information), the photocurrent density of ZnCo2O4/BiVO4-35 has the highest photocurrent at 1.23 V vs. RHE, indicating that ZnCo2O4/BiVO4-35 has the best PEC water splitting performance. When the deposition time is less than 35 s, small photocurrents were obtained probably due the lower content of ZnCo2O4. However, when the amount of the supported ZnCo2O4 is excessive, the transport of holes to the electrode surface is hindered, leading to low photocurrents. ZnCo2O4/BiVO4-35 was used in all the experiments without further description. Fig. 3b is LSV curves of the electrodes in the absence of light. The initial potential of BiVO4 is about 2.4 V, while the initial potential of the ZnCo2O4/BiVO4 composite photoanode is about 2.1 V. The initial potential of ZnCo2O4/BiVO4 negatively shifts 300 mV compared to that of BiVO4. This indicates that ZnCo2O4 can reduce the overpotential of PEC water splitting. The chopped I-t curves of the photoanodes were tested at 1.23 V vs. RHE. It can be clearly seen from the I-t curves in Fig. 3c that the photocurrent density of the composite ZnCo2O4/BiVO4 photoanode is stable and higher than that of BiVO4 at constant voltage. Fig. 3d shows the LSV curves of BiVO4 and ZnCo2O4/BiVO4 photoanodes tested in 0.5 mol/L Na2SO4 solution with 1 mol/L Na2SO3. As a hole trapping agent, Na2SO3 is easily oxidized by holes. When Na2SO3 is present, the photogenerated holes reaching the surface of BiVO4 and ZnCo2O4/BiVO4 are all involved in the oxidation of Na2SO3. Therefore, the photocurrent density for Na2SO3 oxidation represents the amount of photogenerated holes reaching on the surface of BiVO4 and ZnCo2O4/BiVO4. In the presence of Na2SO3, the photocurrent density of ZnCo2O4/BiVO4 is higher than that of BiVO4, indicating that the amount of photogenerated holes reaching the surface of ZnCo2O4/BiVO4 is higher than that of BiVO4. This means that the ZnCo2O4/BiVO4 heterojunction structure promotes the separation of photogenerated charges. Fig. S 3 (Supporting information) shows the chopped LSV curves of BiVO4 and ZnCo2O4/BiVO4 with chopped light illumination. When the simulated sunlight was used to illuminate the photoanode, the photocurrent density of the BiVO4 and ZnCo2O4/BiVO4 electrodes immediately increased. When the light is blocked, the photocurrent density of the both photoanodes immediately reduced to almost 0 mA/cm2, suggesting that the photoanodes is very sensitive to light. The photocurrent densities of the electrodes are disparate at different voltages. At the same potential, the photocurrent density of the ZnCo2O4/BiVO4 photoanode is distinctly higher than that of the BiVO4 electrode, implying ZnCo2O4 nanoparticles can significantly improve the PEC water oxidation performance of BiVO4 photoanode.

    Figure 3

    Figure 3.  LSV curves of BiVO4, ZnCo2O4/BiVO4 photoelectrodes (a) with and (b) without light irradiation; (c) Chopped I-t curves of BiVO4, ZnCo2O4/BiVO4 photoelectrodes at 1.23 V vs. RHE in 0.5 mol/L Na2SO4. (d) LSV curves of photoanodes for sulfite oxidation (with 1 mol/L Na2SO3).

    The efficiency of PEC water splitting is determined by light capture efficiency, charge separation efficiency in the bulk (ηbulk) and surface reaction efficiency (ηsurface) of photoelectrode. To analyze these efficiencies, the maximum theoretical photocurrents Jabs under AM 1.5G simulated sunlight was calculated from UV diffuse reflection data of BiVO4 and ZnCo2O4/BiVO4 photoanodes (Fig. S4 in Supporting information). ηbulk was calculated through dividing the photocurrent for Na2SO3 oxidation (JNa, Data in Fig. 3d) by Jabs [44]. ηsurface represents the ratio of surface holes participated in the reaction, which is obtained through dividing photocurrent for water oxidation (data in Fig. 3a) by JNa. Fig. 4a shows that ηbulk of ZnCo2O4/BiVO4 reaches 70% at 1.23 V vs. RHE, while the ηbulk of pure BiVO4 is 50%, indicating the loading of ZnCo2O4 can enhance charge separation efficiency of BiVO4. Fig. 4b represents the ηsurface of the ZnCo2O4/BiVO4 photoanode reaches 50% at 1.23 V vs. RHE, while the ηsurface of the BiVO4 photoanode is merely 25%. As a water oxidation cocatalyst, ZnCo2O4 improves the kinetics of water splitting reaction. Based on LSV curve shown in Fig. 3a, applied bias photon-to-current efficiency (ABPE) values can be obtained [45]. In Fig. 4c, the maximum efficiency of the BiVO4 film is 0.1% at 1.1 V vs. RHE, while the ZnCo2O4/BiVO4 photoanode attains a maximum ABPE value of 0.6% at 0.9 V vs. RHE. The maximum efficiency of ZnCo2O4/BiVO4 composite photoanode is 6 times that of BiVO4. This indicates that ZnCo2O4 improves the photoelectrochemical performance of BiVO4. The IPCE values of BiVO4 and ZnCo2O4/BiVO4 photoanodes were tested at 1.23 V vs. RHE. As can be seen from Fig. 4d, the IPCE value of ZnCo2O4/BiVO4 electrode is higher than that of BiVO4 in the wavelength range of 320~510 nm. The IPCE value of ZnCo2O4/BiVO4 electrode can reach 40% at 380 nm. Absorbed photon-to-current conversion efficiency (APCE) of BiVO4 and ZnCo2O4/BiVO4 electrodes were calculated through dividing IPCE values by light absorption values and the resulting APCE values are given in Fig. S 5 (Supporting information). Almost half absorbed photons can be converted to current by ZnCo2O4/BiVO4 electrode, about 20% higher than that of BiVO4, indicating that the light utilization rate of the electrode is improved after loading ZnCo2O4. To confirm the reliability of our tests, the integrations of IPCE of BiVO4 and ZnCo2O4/BiVO4 photoanodes with the AM 1.5G solar current were carried out. As shown in Figs. 4e and f, the photocurrents of BiVO4 and ZnCo2O4/BiVO4 photoanodes obtained by integrating AM 1.5G solar currents are 1.8 and 3.2 mA/cm2, which are approximate equivalent to the photocurrents in LSV measurement, indicating the tests are reliable.

    Figure 4

    Figure 4.  (a) Charge separation efficiency in the bulk (ηbulk) and (b) surface reaction efficiency (ηsurface) of BiVO4 and ZnCo2O4/BiVO4 photoanodes. (c) ABPE curves of pure BiVO4 and ZnCo2O4/BiVO4 photoelectrodes. (d) IPCE values of BiVO4 and ZnCo2O4/BiVO4 photoanodes; Solar photocurrents of (e)BiVO4 and (f) ZnCo2O4/BiVO4 anodes (left ordinate) and integrated photocurrent at 1.23 V vs. RHE (right ordinate).

    The relative magnitude of the charge transfer rate of BiVO4 and ZnCo2O4/BiVO4 photoanodes was investigated by electrochemical impedance spectroscopy (EIS). The Nyquist plots of EIS spectra of BiVO4 and ZnCo2O4/BiVO4 photoanodes are a semi-circular arc whatever in dark conditions or under AM 1.5G simulated sunlight (Figs. S6a and b in Supporting information). The impedance of the circuit can be explained by the simplified Randles circuit model, in which RΩ represents the solution resistance. Rct represents charge-transfer resistance at the interface of semiconductor and electrolyte and Cct represents the capacitance of the bulk BiVO4 [46]. The smaller the diameter of the arc, the stronger the charge transfer capability is. ZnCo2O4/BiVO4 exhibits a smaller arc diameter than BiVO4, indicating that the charge transfer rate in the photoanode increases after the ZnCo2O4 cocatalyst is supported. It is worth noting that the arc diameter is smaller under light irradiation, indicating charge transfer become easy with light irradiation. The PEC water splitting reaction of ZnCo2O4/BiVO4 photoanode was carried out in a closed reactor. Prior to the water splitting reaction, the reactor was purged with N2 for 1 hour to remove air from the reactor. The PEC water splitting reaction was carried out for 3 h at 1.23 V vs. RHE. 1 mL of gas was withdrawn from the reactor every half hour, and the gas was quickly pumped into the gas chromatograph to detect the amount of hydrogen and oxygen produced. After 3 h of reaction, 153.7 µmol of H2 and 79.8 µmol of O2 were produced with the mole ratio of H2 to O2 of 1.9:1 (Fig. S6c in Supporting information). The average Faradaic efficiency of PEC cell for H2 and O2 production (Fig. S6d in Supporting information) is about 90% and 86%, respectively, which indicates that the photoanode ZnCo2O4/BiVO4 has good water splitting efficiency.

    Mott-Schottky tests were performed to study the semiconductor type of ZnCo2O4, BiVO4 and ZnCo2O4/BiVO4 photoanodes. As demonstrated in Fig. 5a, ZnCo2O4 shows a negative slope, indicating that ZnCo2O4 is p-type semiconductor. While BiVO4 photoanode shows positive slope, meaning the n-type property of BiVO4 (Fig. 5b). The intersection of Mott-Schottky curves and x-axis are close to 0.38 V vs. RHE with/without loading ZnCo2O4, demonstrating the flat band potentials (Vfb) of BiVO4 photoanodes are 0.38V.

    Figure 5

    Figure 5.  Mott-Schottky plots of (a) ZnCo2O4 and (b) BiVO4, ZnCo2O4/BiVO4; (c) Input voltage−output current characteristic curves of BiVO4 and ZnCo2O4/BiVO4 anodes. The dotted line is used as a linear reference; (d) Heterojunction illustration and water splitting mechanism of ZnCo2O4/BiVO4 photoanode.

    The Vfb value of BiVO4 is 0.38V. Considering that the Vfb of n-type semiconductor is slightly positive (~0.1−0.3 V) than the CB potential, the conduction band (CB) potential of BiVO4 is about 0.18 V [47]. The band gap of BiVO4 is 2.46 eV (Fig. S1b). According to the relationship between CB and valence band (VB), the VB of BiVO4 can be calculated to be 2.64 V. The p-type semiconductor ZnCo2O4 has a Vfb value of 1.98 V (Fig. 5a). Because VB position of a p-type semiconductor is approximately 0.2 V higher than its Vfb value, the VB position of ZnCo2O4 is 2.18 V. ZnCo2O4 has a band gap width of 2.10 eV, so the CB position of ZnCo2O4 is calculated to be 0.08 V. The staggered energy levels of BiVO4 and ZnCo2O4 allows them to form p-n junction.

    The formation of p-n junctions between BiVO4 and ZnCo2O4 was proved by input voltage-output current test as shown in Fig. 5c. BiVO4 photoanode shows a linear relationship between input voltage and output current due to the ohmic contact between BiVO4 and FTO substrate [48]. On the contrary, ZnCo2O4/BiVO4 anode represents a nonlinear relation between output current and input voltage, indicating the interface contact type between ZnCo2O4 and BiVO4 is p-n junction [49]. Built-in electric field in the p-n junction favors the transfer of holes from BiVO4 to ZnCo2O4 for water oxidation.

    The PEC water splitting mechanism of ZnCo2O4/BiVO4 is proposed as shown in Fig. 5d. BiVO4 and ZnCo2O4 are excited to generate electrons and holes by illuminating from the back of the photoanode. The electrons in conduction band of ZnCo2O4 migrate to the conduction band of BiVO4, and then transfer to the counter electrode under external circuit for H2 evolution reaction. The holes in valence band of BiVO4 transfer to the valence band position of ZnCo2O4 with the assistance of built-in electric field for water oxidation reaction (O2 evolution reaction). After a long-time reaction, the photocurrent of ZnCo2O4/BiVO4 anode presents only a small decrease (Fig. S7 in Supporting information).

    In summary, we have constructed a p-n heterojunction ZnCo2O4/BiVO4 photoanode by a simple electrophoretic deposition method. The formation of p-n junction between BiVO4 and ZnCo2O4 is proved by input voltage−output current test. Compared with BiVO4, this composite photoanode ZnCo2O4/BiVO4 has better photoelectrochemical water splitting performance. The photocurrent density of ZnCo2O4/BiVO4 reaches 3.0 mA/cm2 at 1.23 V vs. RHE, about 2.1 times greater than that of BiVO4. The formation of a p-n heterojunction between the BiVO4 and ZnCo2O4 improves the separation efficiency of carriers while the cocatalyst ZnCo2O4 accelerates the surface reaction kinetics, leading to enhanced charge separation efficiency and surface reaction efficiency. After 3 h reaction, the produced H2 and O2 achieve 153.7 and 79.8 µmol, respectively. All results demonstrate that ZnCo2O4 can boost the photoelectrochemical water oxidation performance of BiVO4.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    This work was financially supported by the National Natural Science Foundation of China (Nos. 21808189 and 21663027), Natural Science Basic Research Fund of Shaanxi Province (No. 2020JZ20), and Fundamental Research Funds for the Central Universities of Chang'an University (No. 300102299304).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2021.08.082.


    1. [1]

      F. Niu, D. Wang, F. Li, et al., Adv. Energy Mater. 10 (2020) 1900399 doi: 10.1002/aenm.201900399

    2. [2]

      S. Singh, H. Chen, S. Shahrokhi, et al., ACS Energy Lett. 5 (2020) 1487-1497 doi: 10.1021/acsenergylett.0c00327

    3. [3]

      Z. Wang, L. Wang, Sci. China Mater. 61 (2018) 806-821. doi: 10.1007/s40843-018-9240-y

    4. [4]

      H. Kaneko, T. Minegishi, K. Domen, Chem. Eur. J. 24 (2018) 5697-5706 doi: 10.1002/chem.201703104

    5. [5]

      Z. Jiang, Q. Chen, Q. Zheng, et al., Acta Phys-Chim. Sin. 37 (2021) 2010059

    6. [6]

      Y. Li, R. Zhang, J. Li, et al., Chinese Chem. Lett. 32 (2021) 1165-1168 doi: 10.1109/nems51815.2021.9451523

    7. [7]

      N. Zheng, X. He, W. Guo, Z. Hu, Chinese Chem. Lett. 32 (2021) 1993-1997 doi: 10.1016/j.cclet.2020.10.039

    8. [8]

      K. Sivula, R. Zboril, F. Le Formal, et al., J. Am. Chem. Soc. 132 (2010) 7436-7444 doi: 10.1021/ja101564f

    9. [9]

      Z. Luo, T. Wang, J. Zhang, et al., Angew. Chem. Int. Ed. 56 (2017) 12878-12882 doi: 10.1002/anie.201705772

    10. [10]

      J. Huang, P. Yue, L. Wang, H. She, Q. Wang, Chin. J. Catal. 40 (2019) 1408-1420 doi: 10.1016/S1872-2067(19)63399-1

    11. [11]

      J. Long, W. Wang, S. Fu, L. Liu, J. Colloid Interf. Sci. 536 (2019) 408-413 doi: 10.1016/j.jcis.2018.10.074

    12. [12]

      J. Huang, T. Liu, R. Wang, et al., J. Colloid Interf. Sci. 570 (2020) 89-98 doi: 10.3390/ijerph18010089

    13. [13]

      H. She, M. Jiang, P. Yue, et al., J. Colloid Interf. Sci. 549 (2019) 80-88 doi: 10.1016/j.jcis.2019.04.038

    14. [14]

      L. Wang, X. Shi, Y. Jia, et al., Chinese Chem. Lett. 32 (2021) 1869-1878 doi: 10.1016/j.cclet.2020.11.065

    15. [15]

      B. Lamm, B.J. Trzesniewski, H. Doscher, W.A. Smith, M. Stefik, ACS Energy Lett. 3 (2018) 112-124 doi: 10.1021/acsenergylett.7b00834

    16. [16]

      T.L. Kim, M. -J. Choi, H.W. Jang, MRS Commun. (2018) 1-14

    17. [17]

      Q. Shi, S. Murcia-López, P. Tang, et al., ACS Catal. 8 (2018) 3331-3342 doi: 10.1021/acscatal.7b04277

    18. [18]

      A.B. Murphy, P.R.F. Barnes, L.K. Randeniya, et al., Int. J. Hydrogen Energy 31 (2006) 1999-2017 doi: 10.1016/j.ijhydene.2006.01.014

    19. [19]

      Z. Li, W. Luo, M. Zhang, J. Feng, Z. Zou, Energy Environ. Sci. 6 (2013) 347-370 doi: 10.1039/C2EE22618A

    20. [20]

      M. Zhou, H.B. Wu, J. Bao, et al., Angew. Chem. Int. Ed. 52 (2013) 8579-8583 doi: 10.1002/anie.201302680

    21. [21]

      J.H. Kim, J.S. Lee, Adv. Mater. 31 (2019) e1806938 doi: 10.1002/adma.201806938

    22. [22]

      Y. Li, X. Hu, J. Huang, et al., Acta Phys. -Chim. Sin. 37 (2021) 2009022

    23. [23]

      Y. Liu, Y. Guo, L.T. Schelhas, M. Li, J.W. Ager, J. Phys. Chem. C 120 (2016) 23449-23457 doi: 10.1021/acs.jpcc.6b08654

    24. [24]

      C.C. Hou, T.T. Li, Y. Chen, W.F. Fu, ChemPlusChem 80 (2015) 1465-1471 doi: 10.1002/cplu.201500058

    25. [25]

      R. Shen, D. Ren, Y. Ding, et al., Sci. China Mater. 63 (2020) 2153-2188 doi: 10.1007/s40843-020-1456-x

    26. [26]

      H. Zhang, B. Chen, H. Jiang, et al., Nanoscale 10 (2018) 12991-12996 doi: 10.1039/C8NR04195D

    27. [27]

      D. Wu, Y. Wei, X. Ren, et al., Adv. Mater. 30 (2018) 1705366 doi: 10.1002/adma.201705366

    28. [28]

      Y. Pan, K. Sun, S. Liu, et al., J. Am. Chem. Soc. 140 (2018) 2610-2618 doi: 10.1021/jacs.7b12420

    29. [29]

      X. Yu, M. Wang, X. Gong, et al., Adv. Energy Mater. 8 (2018) 1802445 doi: 10.1002/aenm.201802445

    30. [30]

      M.W. Kanan, D.G. Nocera, Science 321 (2008) 1072-1075 doi: 10.1126/science.1162018

    31. [31]

      J. Liu, J. Li, M. Shao, M. Wei, J. Mater. Chem. A 7 (2019) 6327-6336 doi: 10.1039/c8ta11573g

    32. [32]

      Z. Chen, Y. Ha, H. Jia, et al., Adv. Energy Mater. 9 (2019) 1803918 doi: 10.1002/aenm.201803918

    33. [33]

      D. Zhao, H. Liu, X. Wu, Nano Energy 57 (2019) 363-370 doi: 10.1016/j.nanoen.2018.12.066

    34. [34]

      S. Sun, Y. Sun, Y. Zhou, et al., Angew. Chem. Int. Ed. 58 (2019) 6042-6047 doi: 10.1002/anie.201902114

    35. [35]

      G.S. Hutchings, Y. Zhang, J. Li, et al., J. Am. Chem. Soc. 137 (2015) 4223-4229 doi: 10.1021/jacs.5b01006

    36. [36]

      T.W. Kim, M.A. Woo, M. Regis, K. -S. Choi, J. Phys. Chem. Lett. 5 (2014) 2370-2374 doi: 10.1021/jz501077u

    37. [37]

      J.R. Ding, K.S. Kim, Chem. Eng. J. 334 (2018) 1650-1656 doi: 10.1016/j.cej.2017.11.130

    38. [38]

      L. Xia, J. Bai, J. Li, et al., Appl. Catal. B: Environ. 204 (2017) 127-133 doi: 10.1016/j.apcatb.2016.11.015

    39. [39]

      N.A. Mohamed, H. Ullah, J. Safaei, et al., J. Phys. Chem. C 123 (2019) 9013-9026 doi: 10.1021/acs.jpcc.9b00217

    40. [40]

      S. Bai, H. Chu, X. Xiang, et al., Chem. Eng. J. 350 (2018) 148-156 doi: 10.1016/j.cej.2018.05.109

    41. [41]

      Z. Ma, H. Hou, K. Song, et al., ChemElectroChem 5 (2018) 3660-3667 doi: 10.1002/celc.201801233

    42. [42]

      Z.Q. Liu, H. Cheng, N. Li, T.Y. Ma, Y.Z. Su, Adv. Mater. 28 (2016) 3777-3784 doi: 10.1002/adma.201506197

    43. [43]

      Y. Tang, R. Wang, Y. Yang, D. Yan, X. Xiang, ACS Appl. Mater. Interfaces 8 (2016) 19446-19455 doi: 10.1021/acsami.6b04937

    44. [44]

      J. Huang, Y. Zhang, Y. Ding, ACS Catal. 7 (2017) 1841-1845 doi: 10.1021/acscatal.7b00022

    45. [45]

      J. Huang, G. Hu, Y. Ding, M. Pang, B. Ma, J. Catal. 340 (2016) 261-269 doi: 10.1016/j.jcat.2016.05.007

    46. [46]

      B. Klahr, S. Gimenez, F. Fabregat-Santiago, J. Bisquert, T.W. Hamann, J. Am. Chem. Soc. 134 (2012) 16693-16700 doi: 10.1021/ja306427f

    47. [47]

      H. Huang, X. Li, J. Wang, et al., ACS Catal. 5 (2015) 4094-4103 doi: 10.1021/acscatal.5b00444

    48. [48]

      Y. Lansari, J. Ren, B. Sneed, et al., Appl. Phys. Lett. 61 (1992) 2554 doi: 10.1063/1.108124

    49. [49]

      S.J. Wang, W.J. Lu, G. Cheng, et al., Appl. Phys. Lett. 94 (2009) 263106 doi: 10.1063/1.3158953

  • Figure 1  SEM images of (a) BiVO4, (b) ZnCo2O4/BiVO4-35. TEM images of (c) BiVO4, (d) ZnCo2O4/BiVO4-35. (e, f) HR-TEM images of ZnCo2O4/BiVO4-35.

    Figure 2  (a) XRD patterns of FTO substrate, BiVO4, ZnCo2O4 and ZnCo2O4/BiVO4 composites and XPS spectra of (b) Co 2p, (c) Zn 2p, (d) O 1s in ZnCo2O4/BiVO4.

    Figure 3  LSV curves of BiVO4, ZnCo2O4/BiVO4 photoelectrodes (a) with and (b) without light irradiation; (c) Chopped I-t curves of BiVO4, ZnCo2O4/BiVO4 photoelectrodes at 1.23 V vs. RHE in 0.5 mol/L Na2SO4. (d) LSV curves of photoanodes for sulfite oxidation (with 1 mol/L Na2SO3).

    Figure 4  (a) Charge separation efficiency in the bulk (ηbulk) and (b) surface reaction efficiency (ηsurface) of BiVO4 and ZnCo2O4/BiVO4 photoanodes. (c) ABPE curves of pure BiVO4 and ZnCo2O4/BiVO4 photoelectrodes. (d) IPCE values of BiVO4 and ZnCo2O4/BiVO4 photoanodes; Solar photocurrents of (e)BiVO4 and (f) ZnCo2O4/BiVO4 anodes (left ordinate) and integrated photocurrent at 1.23 V vs. RHE (right ordinate).

    Figure 5  Mott-Schottky plots of (a) ZnCo2O4 and (b) BiVO4, ZnCo2O4/BiVO4; (c) Input voltage−output current characteristic curves of BiVO4 and ZnCo2O4/BiVO4 anodes. The dotted line is used as a linear reference; (d) Heterojunction illustration and water splitting mechanism of ZnCo2O4/BiVO4 photoanode.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  458
  • HTML全文浏览量:  73
文章相关
  • 发布日期:  2022-04-15
  • 收稿日期:  2021-06-08
  • 接受日期:  2021-08-17
  • 修回日期:  2021-07-21
  • 网络出版日期:  2021-08-21
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章