Chinese Chemical Letters  2019, Vol. 30 Issue (3): 735-738   PDF    
Construction of a novel ZnCo2O4/Bi2O3 heterojunction photocatalyst with enhanced visible light photocatalytic activity
Jun Chena, Jing Zhana,*, Yumeng Zhanga, Yiwei Tangb     
a School of Metallurgy and Environment, Central South University, Changsha 410083, China;
b Guangdong Jiana Energy Technology Co., Ltd., Qingyuan 513056, China
Abstract: A novel ZnCo2O4/Bi2O3 heterojunction photocatalyst was prepared, and the formation of the heterojunction was confirmed via HRTEM. Photocatalytic activity of as-prepared samples was evaluated through photodegradation of malachite green (MG). The degradation results show that the as-prepared 13% ZnCo2O4/Bi2O3 heterojunction photocatalyst exhibits higher activity than pure Bi2O3. The MG degradation rate for the as-prepared catalyst is as high as 94%. The enhanced photocatalytic activity is mainly attributed to the broad photoabsorption and low recombination rate of photogenerated electronhole pairs, which is driven by the photogenerated potential difference formed at the ZnCo2O4/Bi2O3 heterojunction interface.
Keywords: ZnCo2O4/Bi2O3     Photocatalyst     Heterojunction     Semiconductors     Degradation    

In recent years, use of semiconductor photocatalysts in wastewater treatment has attracted an increasing amount of attention [1-4]. A variety of materials have been used as photocatalysts, including oxides, sulfides and nitrides [5]. Among these materials, Bi-based semiconductor oxides, such as Bi2O3 [6], BiFeO3 [7], and BiOCl [8], have been intensively investigated because of their excellent photocatalytic activity in the field of environmental depollution. In addition, Bi2O3 has attracted much attention due to its narrow-band gap, high refractive index, oxygen ion conductivity, and great photoconductivity [9]. However, its high recombination rate of photogenerated electron-hole pairs limits its photocatalytic activity and practical applications. To solve these problems, it is necessary to combine Bi2O3 with other semiconductor materials to construct a heterojunction photocatalyst [10], such as Bi2O3/g-C3N4 [11] or Bi2O3/BiVO4 [12].

ZnCo2O4 is a new spinel oxide semiconductor with a band gap of approximately 1.77 eV, and it can absorb visible light to degrade organic [13]. In its band structure, the valence band (VB) of ZnCo2O4 is the O 2p level, and the conduction bands (CBs) is composed of the Co 3d-eg and Co 3d-t2g levels [14]. This particular band structure is beneficial for internal electron transitions, which are helpful for decreasing the recombination rate of photogenerated electron-hole pairs. It has also been found that ZnCo2O4 is an ideal candidate for constructing an effective heterojunction with other photocatalysts [15]. Inspired by this consideration, it is supposed that the band edges of ZnCo2O4 and Bi2O3 match well with each other and that a constructed heterojunction consisting of two semiconductors should have a higher ability to absorb visible light and faster charge carrier transfer. Thus, a novel ZnCo2O4/Bi2O3 heterojunction photocatalyst will be developed in this study, and the composite's photocatalytic activity in MG degradation under visible light illumination will be investigated for the first time.

The preparation process of the sample is as follows: (1) preparation of the photocatalyst: the ZnCo2O4 precursor was prepared in a previous report [16]. Pure Bi2O3 was purchased from Aladdin Biochemical Technology Co., Ltd. (2) preparation of the ZnCo2O4/Bi2O3 heterojunction photocatalyst: typically, Bi2O3 (2.0 g) and ZnCo2O4 (0.434 g) precursors were mixed and ball milled for 30 min. After drying at 80 ℃, the powder mixture was calcined at 400 ℃ for 1 h to prepare the 13% ZnCo2O4/Bi2O3 heterojunction photocatalyst.

X-ray diffraction patterns were recorded using a D8 Advance (XRD, Rigaku-TTRⅢ) diffractometer with Cu Kα radiation (λ = 0.1546 nm). Transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) were recorded on a Tecnai G2 F20 electron microscope. UV–vis diffuse reflection spectra (DRS) were obtained using a Shimadzu UV2550 spectrophotometer. Photoluminescence (PL) spectroscopy was measured at room temperature using a fluorescence spectrophotometer (Jobin Yvon Fluorolog 3–21). Electrochemical impedance spectroscopy (EIS) was collected at a frequency of 1000 Hz.

The photocatalytic activity of the obtained ZnCo2O4/Bi2O3 was evaluated via MG degradation under visible light irradiation. Before irradiation, 0.12 g of the as-prepared ZnCo2O4/Bi2O3 was added into 60 mL of 10 mg/L MG solution, which was magnetically stirred in the dark for 30 min to establish equilibrium between MG and the catalyst. The concentration of MG was monitored at given time intervals by measuring the maximum UV absorbance at 619 nm after commencing irradiation.

XRD patterns of as-prepared samples are shown in Fig. 1a. The strong diffraction peaks of Bi2O3 at 21.72°, 25.757°, 26.922°, 27.377°, 28°, and 33.241° are in accordance with the (020), (002), (111), (120), (012), and (200) planes of monoclinic Bi2O3 corresponding to JCPDS No. 41-1449, respectively. The XRD pattern of pure ZnCo2O4 is in good agreement with the reported data (JCPDS No. 23-1390). However, the characteristic diffraction peaks of ZnCo2O4 in 13% ZnCo2O4/Bi2O3 heterojunction photocatalyst are very weak because of its low content.

Download:
Fig. 1. (a) XRD patterns of Bi2O3, ZnCo2O4, and 13% ZnCo2O4/Bi2O3. (b) EDX spectrum and (c) TEM and (d) HRTEM images of 13% ZnCo2O4/Bi2O3.

Fig. 1c shows an overview of a typical TEM image of the heterojunction photocatalyst. Different kinds of lattice fringes are clearly seen in Fig. 1d. Specifically, distances of 0.29 and 0.261 nm correspond respectively to the (-241) and (-232) lattice planes of Bi2O3, and distances of 0.3 and 0.321 nm correspond respectively to the (422) and (511) planes of ZnCo2O4, indicating heterojunction was indeed formed between ZnCo2O4 and Bi2O3. In addition, the EDX spectrum of the photocatalyst (Fig. 1b) shows 13% ZnCo2O4/Bi2O3 is only composed of Bi, Zn, Co, and O.

Fig. 2a shows the PL spectra of pure Bi2O3, ZnCo2O4, and 13% ZnCo2O4/Bi2O3 at an excitation wavelength of 320 nm. As shown in Fig. 2a, the emission peak for the 13% ZnCo2O4/Bi2O3 heterojunction photocatalyst has comparatively weaker intensity than that of pure Bi2O3, indicating a lower recombination rate of photogenerated electron-hole pairs [17, 18]. As shown in Fig. 2b, the arc radius of the EIS Nyquist plot of the heterojunction photocatalyst is much smaller than that of the EIS Nyquist plot of Bi2O3, showing that the heterojunction structure can accelerate the photogenerated electron-hole pair separation. Therefore, it is expected that the as-prepared heterojunction photocatalyst will possess fast photogenerated electron-hole pair separation and transfer capability, which are helpful for significantly enhancing photocatalytic activity.

Download:
Fig. 2. (a) PL emission spectra, (b) EIS, (c) UV–vis diffuse reflectance spectra, and (d) band gap energies of Bi2O3, ZnCo2O4, and 13% ZnCo2O4/Bi2O3.

The UV–vis diffuse reflectance spectra (Fig. 2c) shows the absorption edge of pure Bi2O3 is about 454 nm and the absorption edge of the 13% ZnCo2O4/Bi2O3 heterojunction shifts to longer wavelength (about 788 nm), which indicates that there is an increase of the absorption range of visible-light. Moreover, the absorption intensity of the 13% ZnCo2O4/Bi2O3 heterojunction is higher than that of Bi2O3. Fig. 2d shows the band gap energies of asprepared samples. It can be seen from Fig. 2d that the values of the band gap energy for pure Bi2O3, ZnCo2O4, and the 13% ZnCo2O4/Bi2O3 composite are nearly 2.8 eV, 1.26 eV, and 1.5 eV, respectively. Compared with the band gap of the pure Bi2O3 sample, the band gap of the 13% ZnCo2O4/Bi2O3 is lower, which is beneficial for generating more photogenerated charge carriers under visible light irradiation.

Fig. 3a shows photocatalytic rate for MG of the as-prepared samples. As shown in Fig. 3a, the photocatalytic activity of the ZnCo2O4/Bi2O3 heterojunction photocatalyst with different ZnCo2O4 content are higher than that of pure Bi2O3 (degradation rate 30% for MG), especially, the degradation rate for MG of the 13% ZnCo2O4/Bi2O3 heterojunction photocatalyst is up to 94% in 120 min. It is worth noting that, the degradation rate for MG of the 13% ZnCo2O4/Bi2O3 (physically mixed) and pure ZnCo2O4 is up to 42% and 98% in 120 min, respectively. The superior photocatalytic activity of pure ZnCo2O4 can be ascribed to its special structural feature. The particular structure of the mesoporous fibers provides the mesopores/channels with larger surface area can enhance accessibility of photo-generated electrons and holes [16]. In addition, the ZnCo2O4 fibers were built by plenty of nanoparticles and the nanoparticles were interconnected to each other to form stable fibers, which provided more active sites for the photocatalytic reactions [16]. The improving photocatalytic activity of ZnCo2O4/Bi2O3 results from the heterojunction formation of between ZnCo2O4 and Bi2O3 which can facilitate charge carrier transfer, suppress the recombination rate of photogenerated electron-hole pairs, and enhance the absorption and utilization of visible light. Recycling reactions for the photodegradation of MG over 13% ZnCo2O4/Bi2O3 under visible light irradiation were carried out to investigate the stability, and the results are shown in Fig. 3b. The photocatalytic activity of the 13% ZnCo2O4/Bi2O3 sample had no apparent deactivation even after four successive cycles for MG degradation (only about 8% loss). The high stability of 13% ZnCo2O4/Bi2O3 can probably be attributed to the formation of the heterojunction.

Download:
Fig. 3. (a) MG degradation catalyzed by Bi2O3, ZnCo2O4, and ZnCo2O4/Bi2O3 (λ > 420 nm). (b) Cycling degradation curves for 13% ZnCo2O4/Bi2O3. (c) Effects of different scavengers on MG degradation over 13%ZnCo2O4/Bi2O3. (d) Mechanism of charge separation of 13% ZnCo2O4/Bi2O3.

Fig. 3c shows the active species trapping experiments of the 13% ZnCo2O4/Bi2O3 heterojunction photocatalyst using isopropanol (IPA), ammonium (AO), and benzoquinone (BQ) as scavengers for ·OH, h+ and ·O2-. As seen in Fig. 3c, ·O2- plays a major role in MG degradation. Fig. 3d shows a possible photocatalytic reaction mechanism. To explain the photocatalytic degradation mechanism, the VB and CB edge positions of Bi2O3 and ZnCo2O4 were obtained (Fig. S1 in Supporting information). When 13% ZnCo2O4/Bi2O3 is excited under visible light irradiation, the photogenerated electrons on the CB of Bi2O3 transfer to the CB of ZnCo2O4, whereas the photogenerated holes on the VB of ZnCo2O4 migrate to the VB of Bi2O3, and these transfers make electron-hole pair separation more efficient. As a result, the recombination of formed electrons and holes can be inhibited, and this enhances the photocatalytic activity of the 13% ZnCo2O4/Bi2O3 photocatalyst.

In summary, the novel heterojunction photocatalyst ZnCo2O4/Bi2O3 was constructed. Degradation results show that the asprepared 13% ZnCo2O4/Bi2O3 has the highest rate of MG degradation (94%), which is almost 3.1 times higher than that of pure Bi2O3. The heterojunction formation between Bi2O3 and ZnCo2O4 is beneficial for inhibiting the recombination rate of photogenerated electron-hole pairs between Bi2O3 and ZnCo2O4, and improving the absorption and utilization of visible light. Therefore, 13% ZnCo2O4/Bi2O3 may be a promising photocatalyst for removing organic pollutants in wastewater.

Acknowledgments

This work was financially supported by Jiana Foundation of Central South University (No. JNJJ201613) and the National Natural Science Foundation of China (No. 51404306).

Appendix A. Supplementary data

Supplementary material related to this article can be found, in the online version, at doi:https://doi.org/10.1016/j.cclet.2018.08.020.1

References
[1]
G. Zhou, M.F. Wu, Q.J. Xing, et al., Appl. Catal. B:Environ. 220 (2017) 607-614.
[2]
F. Chen, H. Huang, Y. Zhang, T. Zhang, Chin. Chem. Lett. 28 (2017) 2244-2250. DOI:10.1016/j.cclet.2017.09.017
[3]
C. Han, Y. Wang, Y. Lei, et al., Nano Res. 8 (2015) 1199-1209. DOI:10.1007/s12274-014-0600-2
[4]
X. Li, F. Zhang, X. Sun, et al., Chin. Chem. Lett. 29 (2018) 1219-1222. DOI:10.1016/j.cclet.2017.12.024
[5]
X.W. Cheng, Q.F. Cheng, X.Y. Deng, P. Wang, H.L. Liu, Electrochim. Acta 184 (2015) 264-275. DOI:10.1016/j.electacta.2015.10.073
[6]
K.K. Bera, R. Majumdar, M. Chakraborty, S.K. Bhattacharya, J. Hazar. Mater. 352 (2018) 182-191. DOI:10.1016/j.jhazmat.2018.03.029
[7]
X. Sun, Z. Liu, H. Yu, Z. Zheng, D. Zeng, Mater. Lett. 219 (2018) 225-228. DOI:10.1016/j.matlet.2018.02.052
[8]
D. Sánchez-Rodríguez, M.G.M. Medrano, H. Remita, V. Escobar-Barrios, J. Environ. Chem. Eng. 6 (2018) 1601-1612. DOI:10.1016/j.jece.2018.01.061
[9]
Q. Que, Y. Xing, Z. He, et al., Mater. Lett. 209 (2017) 220-223. DOI:10.1016/j.matlet.2017.07.115
[10]
J.G. Guo, Y. Liu, Y.J. Hao, et al., Appl. Catal. B. Environ. 224 (2018) 841-853. DOI:10.1016/j.apcatb.2017.11.046
[11]
L. Zhang, G. Wang, Z. Xiong, H. Tang, C. Jiang, Appl. Surf. Sci. 436 (2018) 162-171. DOI:10.1016/j.apsusc.2017.11.280
[12]
X. Liu, Y. Kang, Mater. Lett. 164 (2016) 229-231. DOI:10.1016/j.matlet.2015.10.137
[13]
H. Guo, J. Chen, W. Weng, Q. Wang, S. Li, Chem. Eng. J. 239 (2014) 192-199. DOI:10.1016/j.cej.2013.11.021
[14]
B. Cui, H. Lin, X.C. Zhao, J.B. Li, W.D. Li, Acta Phys. Chim. Sin. 27 (2011) 2411-2415.
[15]
J. Rashid, M.A. Barakat, R.M. Mohamed, I.A. Ibrahim, J. Photochem. Photobiol. A:Chem. 284 (2014) 1-7. DOI:10.1016/j.jphotochem.2014.03.017
[16]
J. Chen, J. Zhan, E. Lu, et al., Mater. Lett. 220 (2018) 66-69. DOI:10.1016/j.matlet.2018.02.135
[17]
B. Wang, Y. Wang, Y. Lei, et al., Nano Res. 9 (2016) 886-898. DOI:10.1007/s12274-015-0971-z
[18]
Y. Zhou, Y. Lei, D. Wang, et al., Chem. Commun. 51 (2015) 13305-13308. DOI:10.1039/C5CC05156H