Next Article in Journal
Synthesis, APPI Mass-Spectrometric Characterization, and Polymerization Studies of Group 4 Dinuclear Bis(ansa-metallocene) Complexes
Next Article in Special Issue
Microcystis aeruginosa Synergistically Facilitate the Photocatalytic Degradation of Tetracycline Hydrochloride and Cr(VI) on PAN/TiO2/Ag Nanofiber Mats
Previous Article in Journal
Experimental Evaluation of a Membrane Micro Channel Reactor for Liquid Phase Direct Synthesis of Hydrogen Peroxide in Continuous Flow Using Nafion® Membranes for Safe Utilization of Undiluted Reactants
Previous Article in Special Issue
Enhancement Photocatalytic Activity of the Heterojunction of Two-Dimensional Hybrid Semiconductors ZnO/V2O5
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Full-Spectrum Photocatalytic Activity of ZnO/CuO/ZnFe2O4 Nanocomposite as a PhotoFenton-Like Catalyst

State Key Laboratory of Biobased Material and Green Papermaking, Qilu University of Technology, Shandong academy of sciences, Jinan 250353, China
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(11), 557; https://doi.org/10.3390/catal8110557
Submission received: 2 October 2018 / Revised: 15 November 2018 / Accepted: 16 November 2018 / Published: 18 November 2018
(This article belongs to the Special Issue Nanostructured Materials for Photocatalysis)

Abstract

:
Deriving photocatalysts by the calcination of hydrotalcite-like compounds has attracted growing interest for extending their photocatalytic activity to the visible and even near-infrared (NIR) light regions. Herein, we describe the acquisition of a ZnO/CuO/ZnFe2O4 nanocomposite with good photoFenton-like catalytic activity under UV, visible and near-infrared (NIR) light irradiation by optimizing the calcination temperature of the coprecipitation product of Zn2+, Cu2+ and Fe3+. The ZnO/CuO/ZnFe2O4 nanocomposite is composed of symbiotic crystals of ZnO, CuO and ZnFe2O4, which enable the nanocomposite to show absorption in the UV, visible and NIR light regions and to produce a transient photocurrent in the presence of H2O2 under NIR irradiation. The full-spectrum photoFenton-like catalyst shows improved performance for the degradation of methyl orange with an increasing amount of H2O2 and is very stable in the recycling process. We believe that the ZnO/CuO/ZnFe2O4 nanocomposite is a promising full-spectrum photoFenton-like catalyst for the degradation of organic pollutants.

Graphical Abstract

1. Introduction

Hydrotalcite-like compounds (HLCs), or layered double hydroxides (LDHs), are a class of anionic clays with brucite-like sheets consisting of adjustable mixed metal hydroxides and interlayer anions [1,2]. Many divalent cations, such as Mg2+, Zn2+, Co2+, Ni2+, Mn2+, Fe2+ and Cu2+, and trivalent cations, such as Al3+, Fe3+, Cr3+, Mn3+, Ga3+ and In3+, have been reported to be involved in the preparation of HLCs [1,2,3,4,5]. By calcinating the HLCs at high temperatures, mixed metal oxides (MMOs) with heterojunctions at metal oxide interfaces are formed due to dehydration, dehydroxylation and anion decomposition [6]. When photoactive metal oxides with suitable band gap structures are involved, MMOs derived from HLC calcination generally show superior photocatalytic performance compared to single metal oxides as semiconductor photocatalysts due to the extended absorption spectrum of the MMOs and reduced recombination rates for photogenerated carriers [7,8,9]. Deriving HLC-based photocatalysts by calcination has, therefore, attracted growing interest recently from the viewpoint of designing photocatalysts to either improve their photocatalytic performance [10,11,12,13] or extend their photocatalytic activity to the visible light region [14,15,16].
It is well-known that near-infrared (NIR) light accounts for 44% of the harvestable solar energy. Developing full-spectrum photocatalysts active in the NIR region is necessary for the better use of solar energy in photocatalytic processes [17,18,19,20,21,22]. Many strategies have been developed for fabricating NIR-active photocatalysts involving combination with upconversion materials [23], sensitization with NIR-responsive dyes [24], manipulation of defect bands, vacancies, and other photosensitive sites in semiconductors [17,25,26], and the incorporation of narrow-band-gap semiconductors [26]. However, HLC-based photocatalysts with near-infrared (NIR) activity have rarely been reported [27,28,29]. Of the few reported NIR light active photocatalysts derived from HLC calcination, Er3+-doped ZnO/ZnAl2O4 multiphase oxide (MPO) was prepared by the calcination of Zn/Al/Er-HLC and acquired its the NIR light activity from the doping of Er3+ as an upconversion species in ZnO/ZnAl2O4 MPO [27]. Er3+-doped ZnO-CuO-ZnAl2O4-MPO was produced through calcination of Zn/Cu/Al/Er-HLC. It absorbs NIR light because of the incorporation of narrow-band-gap CuO and the doping of Er3+. The excellent NIR photocatalytic performance of Er3+-doped ZnO-CuO-ZnAl2O4-MPO was ascribed to the formation of n–p–n heterojunctions among Er3+-doped n-ZnAl2O4, Er3+-doped p-CuO and Er3+-doped n-ZnO, which greatly suppresses the recombination of photogenerated electron–hole pairs and extends the life-times of the charge carriers [28]. The ZnO/ZnFe2O4 nanocomposite derived from the calcination of Zn/Fe-LDH also absorbs NIR light due to the presence of ZnFe2O4. However, the separation of photogenerated electron–hole pairs under NIR light irradiation depends on the presence of H2O2. Consequently, the ZnO/ZnFe2O4 nanocomposite catalyzes the degradation of organic pollutants under NIR light irradiation as a photoFenton-like catalyst [29].
Coprecipitation is one of the most popular methods for preparing HLCs [1]. In this work, a new full-spectrum photoFenton-like catalyst ZnO/CuO/ZnFe2O4 nanocomposite was developed by calcinating the coprecipitation product of Zn2+, Cu2+ and Fe3+. The obtained ZnO/CuO/ZnFe2O4 nanocomposite was found to show good optical absorption from the UV to NIR light regions due to the coexistence of CuO and ZnFe2O4 and found to possess excellent photocatalytic activities in the presence of H2O2 under UV, visible and NIR light. The ZnO/CuO/ZnFe2O4 nanocomposite has also very stable catalytic activities under all tested light irradiation. This work provides a new idea for developing full-spectrum photoFenton-like catalysts.

2. Results and Discussion

2.1. Preparation and Optimization

The ZnO/CuO/ZnFe2O4 nanocomposite was prepared by coprecipitation of Zn(NO3)2, Cu(NO3)2 and Fe(NO3)3 with mixed NaOH and Na2CO3 followed by calcination of the coprecipitation product. The molar ratios of Zn2+ to Cu2+ and (Zn2+ + Cu2+) to Fe3+ were both controlled to be 3:1, enabling the incorporation of suitable amounts of copper oxide and ferrite after calcination. Meanwhile, the formation of oxide and spinel (ferrite) phases depends on the calcination temperature, which consequently affects the photoFenton catalytic activity of the as-prepared catalyst. The crystal phases for the calcination products derived from various temperatures were first analyzed using the X-ray diffraction (XRD) technique. The photoFenton-like catalytic activity was analyzed by comparing the removal rate of methyl orange (MO) in the presence of H2O2 under light irradiation with that due to adsorption in the dark without H2O2.
Figure 1 shows the XRD patterns and photoFenton-like catalytic activities for the calcination products. As shown in Figure 1, the XRD pattern of the calcination product derived from 200 °C shows only a very weak distinguishable peak assigned to the (003) plane of an HLC, indicating that no oxide phase is formed at 200 °C. However, this calcination product removes a significant fraction of MO in the dark, with a MO removal rate under either UV, visible or NIR light irradiation higher than that in the dark, indicative of good absorption of MO and considerable full-spectrum photoFenton-like catalytic activity. By further increasing the calcination temperature to 400 °C, the hexagonal phase ZnO (JCPDS card no. 36–1451) is observed to appear, while the absorption capacity and photoFenton-like catalytic activity of the calcination product do not change significantly. As the calcination temperature increases to 600 °C, hexagonal phase ZnO and monoclinic phase CuO (JCPDS card no. 45–0937) are found to appear in the calcination product. The MO absorption capacity is significantly reduced while the photoFenton-like catalytic activity is significantly improved especially under irradiation by UV and NIR light. This observation demonstrates that the occurrence of CuO favors the development of NIR photoFenton-like catalytic activity for the calcination product.
When the calcination temperature reaches 800 °C, the spinel cubic phase ZnFe2O4 (JCPDS card no. 22–1012) is formed in the calcination product, with ZnO and CuO crystal size significantly increased, judging from the XRD patterns. However, no spinel phase assigned to CuFe2O4 was found in the corresponding calcination product. Meanwhile, the adsorption capacity of the calcination product is further remarkably reduced while its photoFenton-like catalytic activity is further improved probably due to the formation of ZnFe2O4. The crystal phases of the calcination product do not change significantly after further increasing the temperature to 1000 °C. Nevertheless, the adsorption capacity of the calcination product is slightly decreased, and its photoFenton-like catalytic activity is greatly lowered, suggesting the importance of adsorption for the photoFenton catalytic degradation of MO. Therefore, for the production of the ZnO/CuO/ZnFe2O4 nanocomposite, the calcination temperature was chosen to be 800 °C.

2.2. Characterization

The morphology and microstructure of the ZnO/CuO/ZnFe2O4 nanocomposite derived from calcination at 800 °C were analyzed using field emission scanning electron microscopy (FE-SEM) and high-resolution transmission electron microscopy (HR-TEM), respectively. Figure 2 shows the FE-SEM images of the ZnO/CuO/ZnFe2O4 nanocomposite and its precursor, as well as the TEM and HR-TEM images of the ZnO/CuO/ZnFe2O4 nanocomposite. The insets in Figure 2d show the magnified images and Fourier transform electron diffraction (FTED) patterns for the corresponding components.
The FE-SEM images shown in Figure 2a,b reveal that the precursor of the ZnO/CuO/ZnFe2O4 nanocomposite, i.e., the coprecipitation product of Zn2+, Cu2+ and Fe3+, consists of irregular nanoparticles with particle sizes ranging from 10 to 30 nm. The ZnO/CuO/ZnFe2O4 nanocomposite derived from calcination of the precursor at 800 °C also consists of irregular particles. The particle size falls in the range of 100–200 nm, which is much greater than that the size of the precursor due to the intergrowth of various crystal phases.
The TEM and HR-TEM images as well as the FTED patterns shown in Figure 2c,d indicate that the irregular particles in the ZnO/CuO/ZnFe2O4 nanocomposite are composed of well-crystallized monoclinic CuO, cubic ZnFe2O4 and hexagonal ZnO, corroborating the results from the XRD analysis. The CuO, ZnO and ZnFe2O4 crystals grow together. Transitions and distortions of lattice fringes are observed among these crystals, yet no clear interface is formed. This observation indicates that the ZnO/CuO/ZnFe2O4 nanocomposite is a symbiotic crystal of ZnO, CuO and ZnFe2O4, which formed the crystal simultaneously from the coprecipitation product of Zn2+, Cu2+ and Fe3+ without being well-separated.
The chemical states of the ZnO/CuOZnFe2O4 nanocomposite were analyzed using X-ray photoelectron spectroscopy (XPS). Figure 3 shows the core-level XPS spectra measured for Zn 2p, Cu 2p, Fe 2p and O 1s for the ZnO/CuOZnFe2O4 nanocomposite. As shown in Figure 3a, the Zn 2p core-level XPS spectrum shows two peaks at binding energies of 1045.0 eV and 1021.9 eV, which are assigned to Zn 2p1/2 and Zn 2p3/2 of tetrahedral Zn2+, respectively [29], confirming that the oxidation state of Zn is +2 in the ZnO/CuOZnFe2O4 nanocomposite.
In the Cu 2p spectrum for the ZnO/CuOZnFe2O4 nanocomposite, as shown in Figure 3b, two main peaks are observed assigned to Cu 2p1/2 and Cu 2p3/2, located at 953.7 eV and 933.8 eV, respectively. In addition, two satellite peaks due Cu 2p3/2 are found, corroborating the element Cu occurring as CuO in the ZnO/CuOZnFe2O4 nanocomposite [30]. In the Fe 2p spectrum ranging from 730 to 705 eV (Figure 3c), two main peaks belonging to Fe 2p1/2 and Fe 2p3/2 appear at 725.3 eV and 711.6 eV, while the satellite peak of Fe 2p3/2 is located at 719.3 eV, indicative of the presence of Fe as Fe3+ [29,31]. The Fe 2p2/3 binding energy value of 711.6 eV matches well with that for Fe 2p3/2 in ZnFe2O4 [32]. Meanwhile, the Fe 2p3/2 peak can be deconvoluted into two peaks located at 712.5 eV and 711.2 eV, which are assigned to octahedral and tetrahedral Fe3+, respectively [31]. This observation indicates that the spinel ZnFe2O4 in the ZnO/CuOZnFe2O4 nanocomposite has a partially inverse spinel structure [29].
The O 1s spectrum shown in Figure 3d displays an asymmetric peak, which can be deconvoluted into three peaks: a low bonding energy peak at 529.8 eV, a middle binding energy peak at 530.7 eV and a high binding energy peak at 531.7 eV. The low binding energy peak is assigned to lattice oxygen O2− from the Zn–O, Cu–O and Fe–O linkages [28,29]. It is much stronger than the other two binding energy peaks, indicating that the majority of oxygen occurs as lattice oxygen O2− in the ZnO/CuO/ZnFe2O4 nanocomposite [33]. The middle binding energy peak is associated with O2− in the oxygen-deficient regions, indicative of the presence of oxygen vacancies in the ZnO/CuO/ZnFe2O4 nanocomposite [29]. The high binding energy peak is attributed to the absorbed oxygen species such as O2, H2O and CO2 [33]; this peak is weaker than the other two binding energy peaks due to the high crystallinity of the ZnO/CuO/ZnFe2O4 nanocomposite.

2.3. Photo-Fenton-like Catalytic Activity

The photoFention-like catalytic activities for the ZnO/CuO/ZnFe2O4 nanocomposite under UV, visible and NIR light irradiation were further investigated by varying the amount of H2O2. To determine if the catalytic degradation of MO in the presence of the ZnO/CuO/ZnFe2O4 nanocomposite and H2O2 is initiated by light, H2O2 or their simultaneous action, the removal rates for MO were also analyzed in the presence of the ZnO/CuO/ZnFe2O4 nanocomposite in the dark, in the presence of the ZnO/CuO/ZnFe2O4 nanocomposite and H2O2 in the dark as well in the presence of the ZnO/CuO/ZnFe2O4 nanocomposite under light irradiation. The stability of the photoFenton-like catalyst reported in this work, ZnO/CuO/ZnFe2O4, was evaluated by performing cycle analyses, with the results shown in Figure 4.
As shown in Figure 4a,c,e, the combination of the ZnO/CuO/ZnFe2O4 nanocomposite with H2O2 (1.2 mL) without light irradiation cannot significantly improve the MO removal rate, compared to the adsorption of the ZnO/CuO/ZnFe2O4 nanocomposite in the dark, indicating that the ZnO/CuO/ZnFe2O4 nanocomposite is not a competent Fenton catalyst. Exposure of the ZnO/CuO/ZnFe2O4 nanocomposite to UV light only slightly improves the MO removal rate, while exposure to visible or NIR light does not grant any significant improvement for the removal of MO, implying that the ZnO/CuO/ZnFe2O4 nanocomposite is not a broad-spectrum photocatalyst. However, the introduction of H2O2 does not only significantly improve the UV photocatalytic activity of the ZnO/CuO/ZnFe2O4 nanocomposite but also produces visible and NIR photodegradation effects on MO, confirming the full-spectrum photocatalytic activity of ZnO/CuO/ZnFe2O4 nanocomposite as a photoFenton-like catalyst. Meanwhile, the photocatalytic activities of the ZnO/CuO/ZnFe2O4 nanocomposite under UV, visible and NIR light are further improved by increasing the amount of H2O2. When the amount of H2O2 is increased from 0.3 to 1.2 mL, the MO removal rate can be further increased from 60.1% to 93.1% under UV light irradiation, from 32.3% to 67.8 under visible light irradiation, and from 64.3% to 93.4% under NIR light irradiation, respectively.
As shown in Figure 4b,d,f, the ZnO/CuO/ZnFe2O4 nanocomposite is very stable under UV, visible and NIR light irradiation as a photoFenton-like catalyst. The MO removal rates are only reduced by 3.6%, 5.3% and 1.8% under UV, visible and NIR light irradiation, respectively, after the ZnO/CuO/ZnFe2O4 nanocomposite is repeatedly used three times. The slight reduction of the photocatalytic efficiency after repeated use is probably due to contamination and loss of the catalyst during separation.

2.4. Optical Properties and Photocurrent Response

To explain the origin for the full-spectrum photoFenton-like catalytic activity of ZnO/CuO/ZnFe2O4 nanocomposite, especially in the NIR light region, its optical absorption properties were analyzed using UV-vis-NIR diffuse reflectance spectroscopy, with its transient photocurrent response in the presence of H2O2 under NIR light irradiation analyzed using an electrochemical workstation. Figure 5a shows the UV-vis-NIR diffuse reflectance spectrum (DRS) for the ZnO/CuO/ZnFe2O4 nanocomposite compared to that for ZnO and the ZnO/ZnFe2O4 nanocomposite. ZnO only shows absorption in the UV region with an absorption edge of 386 nm. The ZnO/ZnFe2O4 nanocomposite extends the optical absorption to the visible light region with an absorption edge of 788 nm due to the incorporation of ZnFe2O4. The ZnO/CuO/ZnFe2O4 nanocomposite further extends the optical absorption to the NIR light region with an absorption edge of 968 nm due to the introduction of CuO. This full-spectrum light absorption for the ZnO/CuO/ZnFe2O4 nanocomposite provides the prerequisite for the ZnO/CuO/ZnFe2O4 nanocomposite to produce full-spectrum photoFenton-like catalytic activity.
From the equation Eg = 1240/λg and the absorption edges (λg) acquired from DRS, the band gaps (Eg) for ZnO, ZnO/ZnFe2O4, and the ZnO/CuO/ZnFe2O4 nanocomposite can be estimated to be 3.21 eV, 1.57 eV and 1.28 eV, respectively. The Eg values for ZnO/ZnFe2O4 and the ZnO/CuO/ZnFe2O4 nanocomposite are very similar to that of ZnFe2O4 [29] and CuO [28]. Therefore, the energy bands for CuO and ZnFe2O4 in the ZnO/CuO/ZnFe2O4 nanocomposite must determine the same in the ZnO/CuO/ZnFe2O4 nanocomposite. The two bands can be estimated to be 1.57 eV and 1.28 eV using the absorption edge for ZnO/CuO/ZnFe2O4 and the ZnO/ZnFe2O4 nanocomposite, respectively.
In a photoFenton-like process, H2O2 acts as an electron acceptor for reacting with the photogenerated electrons, providing hydroxyl radicals (•OH) for the degradation of organic pollutant [29,34]. To produce •OH in the presence of H2O2, the photogenerated electrons in the ZnO/CuO/ZnFe2O4 nanocomposite should have a potential lower than the redox potential of H2O2/•OH (+0.38 eV), i.e., the conduction band (CB) potential of either ZnO, CuO or ZnFe2O4 should be lower than +0.38 eV. The CB potentials for ZnO, ZnFe2O4 and CuO were calculated to be −0.32 eV, +0.58 eV and +0.67 eV, respectively, according to their band gap energies and the concept of Mulliken’s electronegativity [28]. The ZnO/CuO/ZnFe2O4 nanocomposite cannot provide •OH in the presence of H2O2 under visible light or NIR light irradiation if the ZnO, ZnFe2O4 and CuO components occur as a physical mixture, since neither ZnFe2O4 nor CuO can generate electrons acceptable to H2O2.
Figure 5b shows the transient photocurrent response of the ZnO/CuO/ZnFe2O4 nanocomposite under NIR light irradiation in the presence of H2O2 with on–off switch times of 100 s. As shown in Figure 5b, the ZnO/CuO/ZnFe2O4 nanocomposite produces a strong transient photocurrent in the presence of H2O2 under NIR irradiation, while the fluorine-doped tin oxide (FTO) glass hardly shows any photocurrent response under the same conditions. This observation implies that the ZnO/CuO/ZnFe2O4 nanocomposite possesses high transfer efficiency for photogenerated electrons and good separation of photogenerated electron–hole pairs in the presence of H2O2 under NIR light irradiation [17,29].

2.5. Suggested Photocatalytic Mechanism

Based on the above investigations, the excitation and transfer processes for the charge carriers under light irradiation, as well as the production of •OH, which are the active species in the photoFenton-like process, are shown in Figure 6.
The formation of p-n-p heterojunctions shown in Figure 6 is reasonably expected, as is evident from the transitions and distortions of the lattice fringes observed among the n-ZnO, p-CuO and p-ZnFe2O4 nanocrystals (Figure 2d), and the diffusion of electrons (e) from n-ZnO to p-CuO and p-ZnFe2O4 as well as holes (h+) from p-CuO and p-ZnFe2O4 to n-ZnO in the ZnO/CuO/ZnFe2O4 nanocomposite. The formation of the p-n-p junction causes the energy bands in n-ZnO to bend downwards and enables the energy bands in p-CuO and p-ZnFe2O4 to bend upwards to form a unified Fermi level among the ZnO, CuO and ZnFe2O4 nanocrystals. Consequently, the CB potentials for CuO and ZnFe2O4 become even lower than that for ZnO. The photogenerated electrons in the CuO and ZnFe2O4 components under visible and NIR light irradiation can, therefore, either be captured by H2O2 to form •OH, which can oxidize organic pollutants such as MO, or transfer to the CB of the ZnO component, leading to the separation of photogenerated electrons and holes.

3. Experimental

3.1. Preparation of the ZnO/CuO/ZnFe2O4 Nanocomposite

The precursor for the ZnO/CuO/ZnFe2O4 nanocomposite was synthesized using a coprecipitation method [35]. A mixed salt aqueous solution containing 0.045 M Zn(NO3)2·6H2O, 0.015 M Cu(NO3)2∙3H2O and 0.02 M Fe(NO3)3·9H2O and a mixed alkali aqueous solution containing 0.35 M NaOH and 0.05 M Na2CO3 were simultaneously added dropwise to vigorously stirred distilled water at a rate where the pH remained at approximately 10.5. After this addition, the resulting slurry was mixed for 30 min and aged at 65 °C for 3 h. This slurry was then filtered, thoroughly washed with deionized water, dried at 50 °C for 24 h, and milled to obtain the precursor. The precursor was heated to 800 °C in an electric tubular furnace at a rate of 2 °C min−1 before being calcined for 3 h. After being cooled naturally to room temperature, the product was milled to obtain the ZnO/CuO/ZnFe2O4 nanocomposite.
ZnO nanoparticles and the ZnO/ZnFe2O4 nanocomposite were also prepared using similar coprecipitation and calcination procedures, except that the Cu(NO3)2∙3H2O and Fe(NO3)3·9H2O were replaced by the same mole of Zn(NO3)2·6H2O, and Cu(NO3)2∙3H2O was replaced by the same mole of Zn(NO3)2·6H2O.

3.2. Characterization

The XRD patterns for the various calcination products were recorded using a Bruke D8 Advance powder X-ray diffractometer (Karlsruhe, Germany) with Cu Kα radiation (λ = 0.15406 nm). The FE-SEM and HR-TEM images of the samples were obtained with a Hitachi S-4800 field emission scanning electron microscope (Hitachi, Tokyo, Japan) and a JOEL JEM-2100 high-resolution transmission electron microscope (Akishima, Tokyo, Japan), respectively. The XPS spectra were collected by an ESCALAB 250 spectrometer (VG Scientific Ltd., United Kingdom) equipped with a monochromatized Al Kα X-ray source. All of the binding energies were calibrated C1s peak at a binding energy of 284.6 eV. The UV–Vis-NIR DRS were recorded on a spectrophotometer (Cary 5000, Varian, Inc., Palo Alto, CA, USA) with an integrating sphere attachment in the wavelength range of 200–1200 nm. The transient photocurrents were measured using a CHI660E electrochemical workstation (Chenhua Ins. Inc., Shanghai, China) with a standard three-electrode assembly, where Ag/AgCl was used as the reference electrode, a Pt wire was used as the counter electrode, and ZnO/CuO/ZnFe2O4 nanocomposite-coated FTO conductive glass (Zhuhai Kaivo Optoelectronic Technology Co., Ltd., Zhuhai, China) was used as the working electrode. The electrolyte was a 0.5 M Na2SO4 aqueous solution [29].

3.3. Photocatalytic Activity Test

Photocatalytic experiments under UV and visible light irradiation were carried out in a photocatalytic reactor at 25 °C. A 500 W mercury lamp with a primary wavelength of 365 nm and a 350 W Xe arc lamp equipped with an UV optical filter with a cutoff wavelength of 380 nm were used as UV light and visible light sources, respectively. Experiments under NIR light irradiation were performed at a temperature below 30 °C in a self-assembled experimental device, as reported previously [17,29]. A 200 W infrared lamp with a cutoff filter for cutting off the light below 800 nm was employed as the near-infrared source. 20 mg of the nanocomposite sample was suspended in 20 mL of 20 mg/L MO aqueous solution followed by sonication for several seconds; then, a predetermined amount of H2O2 (10 wt. %) was added. The above suspension either remained in the dark or was irradiated under UV, visible or NIR light. At given time intervals, 3 mL of suspension was collected and immediately centrifuged to remove the solid catalyst at 10000 rpm for 5 min. The residual concentration of MO solution was determined by measuring its absorbance at 465 nm with a UV-Vis spectrophotometer (Hitachi UV-300, Tokyo, Japan).

4. Conclusions

By calcinating the coprecipitation product of Zn2+, Cu2+ and Fe3+ at different temperatures, a series of nanocomposites with UV, visible and NIR photocatalytic activities in the presence of H2O2 were prepared. Among the various nanocomposites, the ZnO/CuO/ZnFe2O4 nanocomposite derived from calcination at 800 °C composed of ZnO, CuO and ZnFe2O4 nanocrystal heterojunctions shows the highest photoFenton-like activity. The ZnFe2O4 and CuO components extend the optical absorption of the ZnO/CuO/ZnFe2O4 nanocomposite to the visible and NIR regions, respectively. The presence of H2O2 promotes the separation of photogenerated electrons and holes and confers the ZnO/CuO/ZnFe2O4 nanocomposite with full-spectrum photocatalytic activity. The photocatalytic activity of the ZnO/CuO/ZnFe2O4 nanocomposite is improved by increasing the amount of H2O2 and remains mostly unchanged after undergoing three cycles of repeated use. The ZnO/CuO/ZnFe2O4 nanocomposite is a promising full-spectrum photoFenton-like catalyst for the degradation of organic pollutants.

Author Contributions

Conceptualization, W.L.; Investigation, Z.L. and H.C; Funding acquisition, W.L.; Validation: Z.L, H.C. and W.L.

Acknowledgments

The project was funded by the National Natural Science Foundation of China (Grant No. 31270625).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Q.; O’Hare, D. Recent advances in the synthesis and application of layered double hydroxide (LDH) nanosheets. Chem. Rev. 2012, 112, 4124–4155. [Google Scholar] [CrossRef] [PubMed]
  2. Meng, Y.; Luo, W.; Xia, S.; Ni, Z. Preparation of Salen–metal complexes (metal = Co or Ni) intercalated ZnCr-LDHs and their photocatalytic degradation of Rhodamine B. Catalysts 2017, 7, 143. [Google Scholar] [CrossRef]
  3. Prevot, V.; Tokudome, Y. 3D hierarchical and porous layered double hydroxide structures: An overview of synthesis methods and applications. J. Mater. Sci. 2017, 52, 11229–11250. [Google Scholar] [CrossRef]
  4. Carja, G.; Grosu, E.F.; Mureseanu, M.; Lutic, D. A family of solar light responsive photocatalysts obtained using Zn2+ Me3+ (Me = Al/Ga) LDHs doped with Ga2O3 and In2O3 and their derived mixed oxides: A case study of phenol/4-nitrophenol decomposition. Catal. Sci. Technol. 2017, 7, 5402–5412. [Google Scholar] [CrossRef]
  5. Wang, J.; Zhang, T.; Li, K.; Cao, Y.; Zeng, Y. Dehydrogenation catalysts for synthesis of O-phenylphenol via Cu/Ni/Mg/Al hydrotalcite-like compounds as precursors. Catalysts 2018, 8, 186. [Google Scholar] [CrossRef]
  6. Chowdhury, P.R.; Bhattacharyya, K.G. Ni/Ti layered double hydroxide: Synthesis, characterization and application as a photocatalyst for visible light degradation of aqueous methylene blue. Dalton Trans. 2015, 44, 6809–6824. [Google Scholar] [CrossRef] [PubMed]
  7. Xiang, X.; Xie, L.; Li, Z.; Li, F. Ternary MgO/ZnO/In2O3 heterostructured photocatalysts derived from a layered precursor and visible-light-induced hotocatalytic activity. Chem. Eng. J. 2013, 221, 222–229. [Google Scholar] [CrossRef]
  8. Huo, R.; Kuang, Y.; Zhao, Z.; Zhang, F.; Xu, S. Enhanced photocatalytic performances of hierarchical ZnO/ZnAl2O4 microsphere derived from layered double hydroxide precursor spray-dried microsphere. J. Colloid Interface Sci. 2013, 407, 17–21. [Google Scholar] [CrossRef] [PubMed]
  9. Mikami, G.; Grosu, F.; Kawamura, S.; Yoshida, Y.; Carja, G.; Izumi, Y. Harnessing self-supported Au nanoparticles on layered double hydroxides comprising Zn and Al for enhanced phenol decomposition under solar light. Appl. Catal. B 2016, 199, 260–271. [Google Scholar] [CrossRef]
  10. Nayak, S.; Pradhan, A.C.; Parida, K.M. Topotactic Transformation of solvated MgCr-LDH nanosheets to highly efficient porous MgO/MgCr2O4 nanocomposite for photocatalytic H2 evolution. Inorg. Chem. 2018, 57, 8646–8661. [Google Scholar] [CrossRef] [PubMed]
  11. Kim, S.; Durand, P.; André, E.; Carteret, C. Enhanced photocatalytic ability of Cu, Co doped ZnAl based mixed metal oxides derived from layered double hydroxides. Colloids Surf. A 2017, 524, 43–52. [Google Scholar] [CrossRef]
  12. Pan, D.; Ge, S.; Zhao, J.; Shao, Q.; Guo, L.; Zhang, X.; Lin, J.; Xu, G.; Guo, Z. Synthesis, characterization and photocatalytic activity of mixed-metal oxides derived from NiCoFe ternary layered double hydroxides. Dalton Trans. 2018, 47, 9765–9778. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, Q.; Wang, S.; Chen, F.; Luo, K.; Sun, J.; Gong, C.; Yao, F.; Wang, X.; Wu, J.; Li, X.; et al. Enhanced visible-light-driven photocatalytic removal of refractory pollutants by Zn/Fe mixed metal oxide derived from layered double hydroxide. Catal. Commun. 2017, 99, 15–19. [Google Scholar] [CrossRef]
  14. Chen, M.; Wu, P.; Wei, Q.; Zhu, Y.; Yang, S.; Ju, L.; Zhu, N.; Lin, Z. The role of oxygen vacancy over ZnCr-layered double oxide in enhancing solar light-driven photocatalytic degradation of bisphenol. A. Environ. Chem. 2018, 15, 226–235. [Google Scholar] [CrossRef]
  15. Zhang, G.; Hu, L.; Zhao, R.; Su, R.; Wang, Q.; Wang, P. Microwave-assisted synthesis of ZnNiAl-layered double hydroxides with calcination treatment for enhanced PNP photo-degradation under visible-light irradiation. J. Photochem. Photobiol. A 2018, 356, 633–641. [Google Scholar] [CrossRef]
  16. Goswami, K.; Ananthakrishnan, R.; Mandal, S. Facile synthesis of cation doped ZnO-ZnCo2O4 hetero-nanocomposites for photocatalytic decomposition of aqueous organics under visible light. Mater. Chem. Phys. 2018, 206, 174–185. [Google Scholar] [CrossRef]
  17. Gao, W.; Liu, W.; Leng, Y.; Wang, X.; Wang, X.; Hu, B.; Yu, D.; Sang, Y.; Liu, H. In2S3 nanomaterial as a broadband spectrum photocatalyst to display significant activity. Appl. Catal. B 2015, 176, 83–90. [Google Scholar] [CrossRef]
  18. Zhou, Y.; Liu, P.; Jiang, F.; Tian, J.; Cui, H.; Yang, J. Vanadium sulfide sub-microspheres: A new near-infrared-driven photocatalyst. J. Colloid Interface Sci. 2017, 498, 442–448. [Google Scholar] [CrossRef] [PubMed]
  19. Feng, R.; Lei, W.; Sui, X.; Liu, X.; Qi, X.; Tang, K.; Liu, G.; Liu, M. Anchoring black phosphorus quantum dots on molybdenum disulfide nanosheets: A 0D/2D nanohybrid with enhanced visible−and NIR −light photoactivity. Appl. Catal. B 2018, 238, 444–453. [Google Scholar] [CrossRef]
  20. Wang, X.; Wang, F.; Sang, Y.; Liu, H. Full-spectrum solar-light-activated photocatalysts for light–chemical energy conversion. Adv. Energy Mater. 2017, 7, 1700473. [Google Scholar] [CrossRef]
  21. Chen, J.; Liu, W.; Li, Z.; Liu, H. Thermally-assisted photodegradation of lignin by TiO2/H2O2 under visible/near-infrared light irradiation. Sci. China. Mater. 2018, 61, 382–390. [Google Scholar] [CrossRef]
  22. Yang, M.-Q.; Gao, M.; Hong, M.; Ho, G.W. Visible-to-NIR photon harvesting: Progressive engineering of catalysts for solar-powered environmental purification and fuel production. Adv. Mater. 2018, 1802894. [Google Scholar] [CrossRef] [PubMed]
  23. Acosta-Mora, P.; Domen, K.; Hisatomi, T.; Lyu, H.; Méndez-Ramos, J.; Ruiz-Morales, J.C.; Khaidukov, N.M. Shifting the NIR into the UV-blue: Up-conversion boosted photocatalysis. Opt. Mater. 2018, 83, 315–320. [Google Scholar] [CrossRef]
  24. Li, L.; Zhang, M.; Zhao, Z.; Sun, B.; Zhang, X. Visible/near-IR-light-driven TNFePc/BiOCl organic–inorganic heterostructures with enhanced photocatalytic activity. Dalton Trans. 2016, 45, 9497–9505. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, J.; Liu, W.; Gao, W. Tuning photocatalytic activity of In2S3 broadband spectrum photocatalyst based on morphology. Appl. Surf. Sci. 2016, 368, 288–297. [Google Scholar] [CrossRef]
  26. Tian, J.; Sang, Y.; Yu, G.; Jiang, H.; Mu, X.; Liu, H. A Bi2WO6-based hybrid photocatalyst with broad spectrum photocatalytic properties under UV, visible, and near-infrared irradiation. Adv. Mater. 2013, 25, 5075–5080. [Google Scholar] [CrossRef] [PubMed]
  27. Qin, Z.; Liu, W.; Li, Z.; Chen, H.; Li, G.; Yu, D. Er3+-doped ZnO/ZnAl2O4 multi-phase oxides acting as near-infrared active photocatalyst. J. Mater. Sci. - Mater. Electron. 2018, 29, 8293–8302. [Google Scholar] [CrossRef]
  28. Chen, H.; Liu, W.; Hu, B.; Qin, Z.; Liu, H. A full-spectrum photocatalyst with strong near-infrared photoactivity derived from synergy of nano-heterostructured Er3+-doped multi-phase oxides. Nanoscale 2017, 9, 18940–18950. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, H.; Liu, W.; Qin, Z. ZnO/ZnFe2O4 nanocomposite as a broadspectrum photo-Fenton-like photocatalyst with near-infrared activity. Catal. Sci. Technol. 2017, 7, 2236–2244. [Google Scholar] [CrossRef]
  30. Tougaard, S.M. Determination of the Cu 2p primary excitation spectra for Cu, Cu2O and CuO. Surf. Sci. 2014, 620, 17–22. [Google Scholar] [CrossRef]
  31. Grosvenor, A.P.; Kobe, B.A.; Biesinger, M.C.; McIntyre, N.S. Investigation of multiplet splitting of Fe 2p XPS spectra and bonding in iron compounds. Surf. Interface Anal. 2004, 36, 1564–1574. [Google Scholar] [CrossRef]
  32. Pan, F.; Guo, Y.; Cheng, F.; Fa, T.; Yao, S. Synthesis of ZnFe2O4 nanomagnets by Fe-ion implantation into ZnO and post-annealing. Chin. Phys. B 2011, 20, 127501. [Google Scholar] [CrossRef]
  33. Chen, Y.; Xu, X.L.; Zhang, G.H.; Xue, H.; Ma, S.Y. Blue shift of optical band gap in Er-doped ZnO thin films deposited by direct current reactive magnetron sputtering technique. Physica E 2010, 42, 1713–1716. [Google Scholar] [CrossRef]
  34. Bokare, A.D.; Choi, W. Review of iron-free Fenton-like systems for activating H2O2 in advanced oxidation processes. J. Hazard. Mater. 2014, 275, 121–135. [Google Scholar] [CrossRef] [PubMed]
  35. Hu, B.; Liu, W.; Gao, W.; Han, J.; Liu, H.; Lucia, L.A. Pseudo-Janus Zn/Al-based nanocomposites for Cr(VI) sorption/remediation and evolved photocatalytic functionality. Chem. Eng. J. 2015, 277, 150–158. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray diffraction (XRD) patterns for the ZnO/CuO/ZnFe2O4 nanocomposite derived from calcination for 3 h at different temperatures; photocatalytic activity for various ZnO/CuO/ZnFe2O4 nanocomposites in the presence of 1.0 mL of 10 wt. % H2O2 under (b) UV, (c) visible and (d) NIR light.
Figure 1. (a) X-ray diffraction (XRD) patterns for the ZnO/CuO/ZnFe2O4 nanocomposite derived from calcination for 3 h at different temperatures; photocatalytic activity for various ZnO/CuO/ZnFe2O4 nanocomposites in the presence of 1.0 mL of 10 wt. % H2O2 under (b) UV, (c) visible and (d) NIR light.
Catalysts 08 00557 g001
Figure 2. FE-SEM images of (a) the precursor of ZnO/CuO/ZnFe2O4 nanocomposite and (b) ZnO/CuO/ZnFe2O4 nanocomposite; (c) TEM and (d) HR-TEM images of the ZnO/CuO/ZnFe2O4 nanocomposite. The insets in (d) show the magnified images and Fourier transform electron diffraction (FTED) patterns for the corresponding components.
Figure 2. FE-SEM images of (a) the precursor of ZnO/CuO/ZnFe2O4 nanocomposite and (b) ZnO/CuO/ZnFe2O4 nanocomposite; (c) TEM and (d) HR-TEM images of the ZnO/CuO/ZnFe2O4 nanocomposite. The insets in (d) show the magnified images and Fourier transform electron diffraction (FTED) patterns for the corresponding components.
Catalysts 08 00557 g002
Figure 3. Core-level XPS spectra for (a) Zn 2p, (b) Cu 2p, (c) Fe 2p and (d) O 1s for the ZnO/CuOZnFe2O4 nanocomposite.
Figure 3. Core-level XPS spectra for (a) Zn 2p, (b) Cu 2p, (c) Fe 2p and (d) O 1s for the ZnO/CuOZnFe2O4 nanocomposite.
Catalysts 08 00557 g003
Figure 4. Photocatalytic degradation of MO using ZnO/CuO/ZnFe2O4 nanocomposite (left) and recycled ZnO/CuO/ZnFe2O4 nanocomposite (right) in the presence of H2O2 under (a,b) UV, (c,d) visible and (e,f) NIR light irradiation. The concentration of H2O2 is 10 wt. %. The experiments using recycled ZnO/ZnFe2O4 nanocomposite were carried out in the presence of 1.2 mL of H2O2.
Figure 4. Photocatalytic degradation of MO using ZnO/CuO/ZnFe2O4 nanocomposite (left) and recycled ZnO/CuO/ZnFe2O4 nanocomposite (right) in the presence of H2O2 under (a,b) UV, (c,d) visible and (e,f) NIR light irradiation. The concentration of H2O2 is 10 wt. %. The experiments using recycled ZnO/ZnFe2O4 nanocomposite were carried out in the presence of 1.2 mL of H2O2.
Catalysts 08 00557 g004
Figure 5. (a) UV-vis-NIR diffuse reflectance spectra (DRS) for ZnO, ZnO/ZnFe2O4 and the ZnO/CuO/ZnFe2O4 nanocomposite; (b) photocurrent generation for the ZnO/CuO/ZnFe2O4 nanocomposite in the presence of H2O2 under NIR light. The inset in plane b shows a photograph of the ZnO/CuO/ZnFe2O4 nanocomposite.
Figure 5. (a) UV-vis-NIR diffuse reflectance spectra (DRS) for ZnO, ZnO/ZnFe2O4 and the ZnO/CuO/ZnFe2O4 nanocomposite; (b) photocurrent generation for the ZnO/CuO/ZnFe2O4 nanocomposite in the presence of H2O2 under NIR light. The inset in plane b shows a photograph of the ZnO/CuO/ZnFe2O4 nanocomposite.
Catalysts 08 00557 g005
Figure 6. Schematic diagram for p-n-p heterojunction formation in the ZnO/CuO/ZnFe2O4 nanocomposite and the excitation and transfer process for charge carriers in the ZnO/CuO/ZnFe2O4 nanocomposite under light irradiation.
Figure 6. Schematic diagram for p-n-p heterojunction formation in the ZnO/CuO/ZnFe2O4 nanocomposite and the excitation and transfer process for charge carriers in the ZnO/CuO/ZnFe2O4 nanocomposite under light irradiation.
Catalysts 08 00557 g006

Share and Cite

MDPI and ACS Style

Li, Z.; Chen, H.; Liu, W. Full-Spectrum Photocatalytic Activity of ZnO/CuO/ZnFe2O4 Nanocomposite as a PhotoFenton-Like Catalyst. Catalysts 2018, 8, 557. https://doi.org/10.3390/catal8110557

AMA Style

Li Z, Chen H, Liu W. Full-Spectrum Photocatalytic Activity of ZnO/CuO/ZnFe2O4 Nanocomposite as a PhotoFenton-Like Catalyst. Catalysts. 2018; 8(11):557. https://doi.org/10.3390/catal8110557

Chicago/Turabian Style

Li, Zhenzhen, Huabin Chen, and Wenxia Liu. 2018. "Full-Spectrum Photocatalytic Activity of ZnO/CuO/ZnFe2O4 Nanocomposite as a PhotoFenton-Like Catalyst" Catalysts 8, no. 11: 557. https://doi.org/10.3390/catal8110557

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